05.10.2013 Views

Experimentelle Untersuchungen zur Transgenese bei Nutztieren

Experimentelle Untersuchungen zur Transgenese bei Nutztieren

Experimentelle Untersuchungen zur Transgenese bei Nutztieren

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Aus dem<br />

Institut für Nutztiergenetik, Mariensee<br />

im Friedrich-Loeffler-Institut<br />

<strong>Experimentelle</strong> <strong>Untersuchungen</strong> <strong>zur</strong> <strong>Transgenese</strong><br />

<strong>bei</strong> <strong>Nutztieren</strong><br />

Habilitationsschrift<br />

<strong>zur</strong> Erlangung der Venia legendi<br />

an der Stiftung Tierärztliche Hochschule Hannover<br />

Vorgelegt von<br />

Dr. rer. nat. Wilfried August Kues<br />

Hannover 2008<br />

1


2<br />

Für und mit Brigitte


Inhaltsverzeichnis<br />

1.0 Eigene Publikationen <strong>zur</strong> Thematik dieser Habilitationsschrift 4<br />

1.1 Charakterisierung primärer Zellen als Donorzellen für den Kerntransfer 4<br />

1.2 DNA-Mikroinjektion und somatischer Kerntransfer 4<br />

1.3 Übersichtsartikel: Transgene Nutztiere, Stammzellen und DNA-Array Analysen 5<br />

2.0 Einleitung 6<br />

2.1 Die DNA als Träger der genetischen Information und reverse Genetik 6<br />

2.2 Erstellung transgener Nutztiere 8<br />

2.3 Transgene Schweine als Zell- und Organdonoren für die Xenotransplantation 10<br />

2.4 Pharming: Produktion rekombinanter Proteine in der Milchdrüse 12<br />

2.5 Transgene Tiere für landwirtschaftliche Nutzungen 13<br />

2.6 Neuere Methoden der <strong>Transgenese</strong> 14<br />

2.6.1 Klonen durch somatischen Kerntransfer 14<br />

2.6.2 Gene knock-down in <strong>Nutztieren</strong> durch RNA-Interferenz 16<br />

2.6.3 Virus-vermittelte <strong>Transgenese</strong> 16<br />

2.6.4 Konditionale Mutagenese 17<br />

2.6.5 Induzierbare Promotoren und konditionale Expressionskontrolle 18<br />

2.7 Genomprojekte und ihre Bedeutung für die <strong>Transgenese</strong> 19<br />

2.8 Epigenetik 22<br />

3.0 Fragestellungen 24<br />

4.0 Ergebnisse und Diskussion 25<br />

4.1 Primäre somatische und Stamm-Zellen als Donorzellen für den Kerntransfer 25<br />

4.2 Gene knock-down-Versuche in transgenen Embryonen 29<br />

4.3 Generierung transgener Schweine mit konditionaler Transgenregulation 32<br />

5.0 Zusammenfassung 39<br />

6.0 Literaturverzeichnis 40<br />

7.0 Danksagungen 54<br />

8.0 Darstellung des eigenen Anteils an den wissenschaftlichen Ar<strong>bei</strong>ten 55<br />

9.0 Abkürzungsverzeichnis 59<br />

10.0 Anhang: Nachdruck der eigenen Publikationen 60<br />

3


1.0 Eigene Publikationen <strong>zur</strong> Thematik dieser Habilitationsschrift<br />

1.1 Charakterisierung primärer Zellen als Donorzellen für den Kerntransfer<br />

Anger, M., Kues, W.A., Klima, J., Mielenz, M., Motlik, J., Carnwath, J.W. and Niemann, H.<br />

2003. Cloning and cell cycle-specific regulation of Polo-like kinase in porcine fetal<br />

fibroblasts.<br />

Molecular Reproduction and Development 65, 245-253.<br />

Kues, W. A., Petersen, B., Mysegades, W., Carnwath, J.W. and Niemann, H. 2005. Isolation<br />

of fetal stem cells from murine and porcine somatic tissue.<br />

Biology of Reproduction 72, 1020-1028<br />

Yadav, P., Kues, W.A., Herrmann, D., Carnwath, J.W., Niemann, H. 2005. Bovine ICMs<br />

express the Oct4 ortholog.<br />

Molecular Reproduction and Development 72, 182-190.<br />

Dieckhoff, B., Karlas, A., Hofmann, A., Kues, W.A., Petersen, B., Pfeifer, A., Niemann, H.,<br />

Kurth, R., Denner, J. 2007. Inhibition of porcine endogenous retroviruses (PERVs) in primary<br />

porcine cells by RNA interference using lentiviral vectors.<br />

Archives of Virology 152, 629-34.<br />

1.2 DNA-Mikroinjektion und somatischer Kerntransfer<br />

Schätzlein S., Lucas-Hahn, A., Lemme, E., Kues, W.A., Dorsch, M., Manns, M., Niemann,<br />

H., Rudolph, K.H. 2004. Telomere length is reset during early embryogenesis.<br />

Proceedings of the National Academy of Sciences of the USA 101, 8034-8038.<br />

Hölker, M., Petersen, B., Hassel, P., Kues, W.A., Lemme, E., Lucas-Hahn, A., Niemann, H.<br />

2005. Duration of in vitro maturation of recipient oocytes affects blastocyst development of<br />

cloned porcine embryos.<br />

Cloning and Stem Cells 7, 34-43.<br />

Kues, W.A., Schwinzer, R., Wirth, D., Verhoeyen, E., Lemme, E., Herrmann, D., Barg-Kues,<br />

B., Hauser, H., Wonigeit, K., Niemann H. 2006. Epigenetic silencing and tissue independent<br />

expression of a novel tetracycline inducible system in double-transgenic pigs.<br />

FASEB Journal 20, E1-E10, doi: 10.1096/fj.05-5415fje.<br />

Deppenmeier, S., Bock, O., Mengel, M., Niemann, H., Kues, W., Lemme, E., Wirth, D,<br />

Wonigeit, K., Kreipe, H. 2006. Health status of transgenic pig lines expressing human<br />

complement regulator protein CD59.<br />

Xenotransplantation 13, 345-356.<br />

Hornen, N., Kues, W.A., Carnwath, J.W., Lucas-Hahn, A., Petersen, B., Hassel, P., Niemann,<br />

H. 2007. Production of viable pigs from fetal somatic stem cells.<br />

Cloning and Stem Cells 9, 364-73.<br />

Iqbal, K, Kues, W.A., Niemann, H. 2007. Parent of origin dependent gene-specific knock<br />

down in mouse embryos.<br />

Biochemical and Biophysical Research Communications 358, 727-732.<br />

4


Dieckhoff, B., Petersen, B., Kues, W.A., Kurth, R., Niemann, H., Denner, J. 2008.<br />

Knockdown of porcine endogenous retrovirus (PERV) expression by PERV-specific shRNA<br />

in transgenic pigs.<br />

Xenotransplantation 15, 36-45.<br />

1.3 Übersichtsartikel: Transgene Nutztiere, Stammzellen und DNA-Array Analysen<br />

Niemann, H., und Kues, W.A. 2003. Application of transgenesis in livestock for agriculture<br />

and biomedicine.<br />

Animal Reproduction Science 79, 291-317.<br />

Kues, W.A. und Niemann, H. 2004. The contribution of farm animals to human health.<br />

Trends in Biotechnology 22, 286-294.<br />

Kues, W.A., Carnwath, J.W., Niemann, H. 2005. From fibroblasts and stem cells:<br />

Implications for cell therapies and somatic cloning.<br />

Reproduction, Fertility and Development 17, 125-134.<br />

Niemann H, Kues W, Carnwath JW. 2005. Transgenic farm animals: present and future.<br />

OIE Scientific and Technical Reviews (Rev. Sci. Tech. Off. Int. Epiz.) 24, 284-298.<br />

Niemann, H. und Kues, W.A. 2007. Transgenic farm animals – an update.<br />

Reproduction, Fertility and Development 19, 762-770.<br />

Niemann, H., Carnwath, J.W., Kues, W.A. 2007. Application of array technology to<br />

mammalian embryos.<br />

Theriogenology 68S, S165-S177.<br />

Kues, W.A., Rath, D., Niemann, H. 2008. Reproductive biotechnology goes genomics.<br />

CAB Reviews: Perspectives in Agriculture, Veterinary Science, Nutrition and Natural<br />

Resources 3, No. 036, 1-18.<br />

5


2.0 Einleitung<br />

2.1 DNA als Träger der genetischen Information und reverse Genetik<br />

Die Desoxyribonucleinsäure (DNS, bzw. DNA) ist der Träger der genetischen Information.<br />

DNA wurde 1869 erstmals von Friedrich Miescher aus Eiterzellen isoliert, ihre biologische<br />

Funktion als Informationsträger wurde aber erst 1944 durch Transformation von<br />

Pneumokokken gezeigt (Avery et al., 1944). Die Aufklärung der DNA-Molekülstruktur als<br />

Doppelhelix (Watson und Crick, 1953) erlaubte erstmals einen Einblick, wie biologische<br />

Information gespeichert, repliziert und abgelesen wird und bildete eine wichtige Grundlage<br />

für die Entwicklung der Molekularbiologie.<br />

Seit ca. 1970 wurde es möglich, in kurzen DNA-Stücken die Basenabfolge zu bestimmen<br />

(DNA-Sequenzierung), sowie DNA-Stücke gezielt neu zusammen zu setzen und die<br />

Eigenschaften der rekombinanten DNA in vitro und in verschiedenen einfachen<br />

Modellorganismen, vorwiegend Escherichia coli (E. coli), oder in eukaryotischen Zell-Linien,<br />

zu testen (Sanger et al., 1973; Maxam und Gilbert, 1977; Sanger et al., 1977; Jackson et al.,<br />

1972; Johnson, 1983; Watson et al., 1983, Mulligan et al., 1979). So konnten grundlegende<br />

molekulare Eigenschaften der DNA wie der Aufbau von Genen, regulatorischen Sequenzen<br />

(Promotor, Enhancer, Silencer) und die Intron-Exon-Organisation eukaryotischer Gene<br />

aufgeklärt werden (Glover, 1985; Sambrook et al., 1989). Ein Standardverfahren in E.coli ist<br />

die Transformation mit ringförmigen DNA-Molekülen (Plasmide), die eine Antibiotika-<br />

Resistenz und einen Startpunkt für die Replikation (origin of replication) in ihrer Sequenz<br />

tragen. Zusätzlich kann Fremd-DNA in ein Plasmid ligiert werden, typischerweise 1 000 – 20<br />

000 Basenpaare (1-20 kBp). Die Plasmide liegen episomal in E. coli vor und werden <strong>bei</strong> jeder<br />

Zellteilung repliziert (Glover, 1985; Sambrook et al., 1989). Falls die Fremd-DNA<br />

prokaryotische Promotorelemente und eine protein-kodierende Sequenz enthält, wird diese<br />

Information abgelesen und das rekombinante Protein exprimiert. Zahlreiche Medikamente<br />

werden mittlerweile rekombinant in Bakterien hergestellt. Humanes Insulin,<br />

Wachtumshormon, Erythropoietin und Interferon sind die umsatzstärksten Medikamente<br />

(Blockbuster) mit jeweils mehr als 1 Milliarde $ Umsatz pro Jahr (Nightingale und Martin,<br />

2004). In Prokaryoten erfolgt im wesentlichen eine direkte Ablesung der genetischen<br />

Information, im Eukaryoten liegt eine wesentlich komplexere Regulation der Genaktivität<br />

vor, die bisher erst teilweise entschlüsselt ist.<br />

6


Bereits 1971 wurden durch Inkubation von Spermien mit Fremd-DNA transgene Embryonen<br />

erstellt (Brackett et al, 1971), bzw. durch Infektion präimplantatorischer Maus-Embryonen<br />

mit Simian Virus 40 (Jaenisch und Mintz, 1974) oder Moloney Leukemia-Viren (Jaenisch,<br />

1976) transgene Mäuse erzeugt, die die Fremd-DNA in ihr Genom integriert hatten, und an<br />

ihre Nachkommen weitergaben. Trotz Veränderung des Genotyps (reverse genetics) dieser<br />

Tiere, wurden keine phänotypischen Veränderungen gezeigt.<br />

Der erste Nachweis, dass eine rekombinante DNA-Sequenz eine phänotypische Änderung im<br />

Säuger <strong>zur</strong> Folge hatte, wurde 1982 publiziert (Palmiter et al. 1982, Abb.1). Diese Ar<strong>bei</strong>t ist<br />

ein Meilenstein in der Genetik und die Grundlage für ein vertieftes molekulares Verständnis<br />

der Gen-Phänotyp-Beziehung und für transgene Forschungsansätze, die darauf abzielen, die<br />

komplexe Genregulation in Eukaryoten und speziell in Säugern zu verstehen. In dieser Ar<strong>bei</strong>t<br />

wurde ein Expressionskonstrukt mit dem Metallothionein-Promoter und der cDNA des<br />

Rattenwachstumshormons (rGH) verwendet. Das aufgereinigte DNA-Expressionskonstrukt<br />

wurde dann direkt in den männlichen Vorkern von Mauszygoten injiziert, die anschließend in<br />

den Eileiter von scheinschwangeren Empfängertieren übertragen wurden.<br />

Abb. 1. Verstärktes Wachstum in transgenen Mäusen.<br />

Transgene Maus (links), in der die Expression des Ratten-Wachstumshormongens zu verstärktem Wachstum<br />

führt; zum Vergleich ein nicht-transgenes Geschwistertier (rechts) (Palmiter et al., 1982).<br />

Die Integration der rGH-Sequenz in das Mausgenom wurde in den Nachkommen durch<br />

Southern blot-Nachweis mit einer markierten komplementären Sonde gezeigt. Die transgenen<br />

7


Tiere zeigten phänotypisch ein verstärktes Wachstum, das durch die erhöhten Serumspiegel<br />

des Wachstumshormons hervorgerufen wurde.<br />

Transgene Mausmodelle sind mittlerweile ein Standardwerkzeug der molekularbiologischen<br />

Forschung geworden, sie werden für das Verständnis genetischer Erkrankungen, wie<br />

Parkinson, Alzheimer, Trisomie 21, sowie die Entwicklung von Therapiemöglichkeiten in der<br />

Immunologie, der Krebsforschung und der Grundlagenforschung verwendet (Dodart und<br />

May, 2005; Yang und Gong, 2005; Scharschmidt und Segre, 2008). Insbesondere die<br />

Möglichkeit über murine embryonale Stammzellen (ES) gezielt spezifische Gene<br />

auszuschalten (Gene knock-out) hat die Anwendungsfelder der reversen Genetik vergrößert<br />

(Mansour et al., 1988; Capecchi, 1989; Götz et al., 1998; Müller, 1999). Weitere<br />

Verfeinerungen des Methodenrepertoires sind die konditionale Expression von Transgenen<br />

über Rekombinasen (z.B. Cre/loxP-System), die konditionale Expression von binären<br />

Expressionskassetten (tet-on, tet-off) oder die RNA-Interferenz (RNAi) mittels kurzer<br />

doppelsträngiger RNAs (siRNA oder shRNA) (Orban et al., 1992; Gu et al., 1994; Furth et<br />

al., 1994; Zhou et al., 2006).<br />

Allerdings hat das Mausmodell neben zahlreichen Vorteilen auch Beschränkungen, die sich<br />

insbesondere aus den gravierenden anatomischen und physiologischen Unterschieden<br />

zwischen Maus und Mensch ergeben (Habermann et al., 2007; Wall und Shani, 2008; Taft,<br />

2008). Daher ist das Mausmodell für spezifische Fragestellungen, z.B. in<br />

Transplantationsmedizin, Kreislauf/Bluthochdruckforschung oder Gerontologie aufgrund der<br />

geringen Größe und kurzen Lebensdauer nur sehr bedingt aussagekräftig.<br />

2.2 Erstellung transgener Nutztiere<br />

Transgene Nutztiere stellen für spezifische biomedizinische Fragestellungen eine interessante<br />

Alternative zum Mausmodell dar. Hierzu zählen insbesondere Xenotransplantation, Pharming<br />

(Produktion rekombinanter Proteine in der Milchdrüse), Bluthochdruck- und Altersforschung.<br />

Auch in der Landwirtschaft kann eine gezielte genetische Veränderung von <strong>Nutztieren</strong><br />

wesentliche Vorteile für eine effizientere und ressourcenschonendere Produktion haben (Wall,<br />

1996).<br />

Die ersten genetisch veränderten Schweine, Schafe und Kaninchen wurden 1985 durch DNA-<br />

Mikroinjektion in Zygotenstadien erstellt (Hammer et al., 1985, Brem et al., 1985). In der<br />

Folgezeit wurde über transgene Ansätze versucht, verschiedene Eigenschaften von <strong>Nutztieren</strong><br />

zu verbessern: Wachstum und Futterverwertung (Pursel, 1998; Golovan et al., 2001),<br />

Milchleistung (Zuelke, 1998; Brophy et al., 2003), Wolleigenschaften (Damak et al., 1996),<br />

8


sowie Reproduktion und Krankheitsresistenz (Müller et al., 1992; Clements et al., 1994;<br />

Müller und Brem, 1994; Wolf et al., 2000). Als wesentlicher Nachteil dieses Verfahrens<br />

erwies sich, dass die Erfolgsrate transgener Nachkommen mit durchschnittlich 5-10% der<br />

geborenen Tiere sehr gering ist (Abb.2). Bezogen auf die mikroinjizierten Embryonen liegen<br />

die Erfolgsraten <strong>bei</strong> 1-3% (Abb.2). Im Gegensatz <strong>zur</strong> Maus, weisen Zygoten von Rind und<br />

Schwein einen sehr hohen Anteil von Lipiden auf, die die DNA-Mikroinjektion in<br />

Zygotenvorkerne erschweren. Nur durch zusätzliche Schritte, wie Zentrifugation der Zygoten,<br />

können die Vorkerne dargestellt werden (Abb.3). Erschwerend kommt hinzu, dass von den<br />

transgenen Trägertieren (founder animals) nur ein Teil das Transgen vererbt<br />

(Keimbahntransmission) und die gewünschte Expression aufweist (Hogan et al., 1996). Da<br />

die Fremd-DNA zufallsmäßig in das Genom integriert, können Mosaik-Integration,<br />

insertionale Mutationen, variable Kopienzahlen des Transgens und variable Expression durch<br />

Positionseffekte benachbarter DNA-Sequenzen auftreten (Hogan et al., 1986; Deppenmeier et<br />

al., 2006).<br />

Maus<br />

Kaninchen<br />

Schwein<br />

Schaf<br />

Rind<br />

0 5 10 15 [%]<br />

Abb. 2. Effizienz der DNA-Vorkerninjektion in verschiedenen Säugerspezies.<br />

Die <strong>Transgenese</strong>-Effizienz ist als relativer Anteil der transgenen Tieren an den geborenen Nachkommen<br />

(gestrichelte Säulen) und an den DNA-injizierten und auf Empfängertier übertragenen Embryonen (schwarze,<br />

gefüllte Säulen) angegeben (verändert nach Wall, 1996).<br />

In den letzten Jahren wurde deutlich, dass an der komplexen Regulation der Genomaktivität<br />

in Eukaryoten auch zahlreiche sogenannte epigenetische Mechanismen (Epi-Genetik =<br />

„Über“-Genetik) beteiligt sind (Reik, 2007). Als Epigenetik bezeichnet man vererbbare<br />

Eigenschaften, die nicht direkt in der DNA-Sequenz verankert sind. Relativ gut untersucht<br />

sind Methylierungen der Cytosin-Base in CG-Dinukleotiden (CpG) <strong>bei</strong> Säugern, sowie<br />

9


Methylierungen, Azetylierungen und Phosphorylierungen der Histon-Proteine (Chandler und<br />

Jones, 1988; Bird, 1992; Li, 2002; Maatouk et al., 2006).<br />

Bedingt durch die hohen Kosten für die Erstellung transgener Nutztiere, die erforderliche<br />

umfangreiche Logistik und die langen Reproduktionszeiten, fokussierte sich die Forschung im<br />

wesentlichen auf biomedizinisch relevante Gebiete wie Xenotransplantation und Pharming.<br />

Insbesondere <strong>bei</strong> <strong>Nutztieren</strong>, aber auch in der Maus sind die Probleme der <strong>Transgenese</strong> mit<br />

häufig ungenügender Expression, oder ektopischer Expression nur teilweise verstanden.<br />

Gegenwärtig ist es aufgrund der un<strong>zur</strong>eichenden Vorhersagbarkeit der transgenen Expression,<br />

immer noch notwendig, zahlreiche Linien auf die gewünschten Expressionseigenschaften zu<br />

untersuchen. Bei der Maus stellt das aufgrund der kurzen Generationszeiten und der relativ<br />

geringen Kosten kein grundlegendes Problem dar, <strong>bei</strong> <strong>Nutztieren</strong> ergeben sich enorme Kosten<br />

<strong>bei</strong> der Erstellung und Charakterisierung. Es wird geschätzt, dass die Kosten für die<br />

Erstellung eines transgenen Rindes ca. 100 000 $ betragen, <strong>bei</strong> Schweinen werden Kosten<br />

von ca. 25 000 $ genannt (Wall, 1996).<br />

A B<br />

Promotor Transgen-cDNA Poly(A)<br />

Abb. 3. DNA-Mikroinjektion in einen Vorkern einer porcinen Zygote.<br />

A) Nach Zentrifugation <strong>bei</strong> 10 000 g ist die undurchsichtige Lipidfraktion des Zytoplasmas in einer Zellhälfte<br />

konzentriert. In einen der dadurch sichtbar gewordenen Vorkerne wird mit Hilfe einer Mikrokapillare (rechte<br />

Bildhälfte) die DNA-Lösung injiziert. B) Schematische Darstellung eines Minigen-DNA-Konstrukts bestehend<br />

aus einem Promotor, einer protein-kodierenden cDNA und einer Polyadenylierungssequenz, die die Stabilität der<br />

abgelesenen mRNA erhöht.<br />

Daraus ergibt sich, dass für einen breiteren Einsatz der <strong>Transgenese</strong> <strong>bei</strong> <strong>Nutztieren</strong>, die<br />

Effizienz der <strong>Transgenese</strong> erhöht und die Kosten deutlich sinken müssen. Trotz dieser<br />

Schwierigkeiten sind in den letzten Jahren wesentliche Fortschritte in der Erstellung<br />

transgener Nutztiere erzielt worden.<br />

2.3 Transgene Schweine als Zell- und Organdonoren für die Xenotransplantation<br />

Die Erfolge der Allotransplantation, d.h. die Transplantation humaner Organe in Patienten mit<br />

Organausfall, hat zu einer dramatischen Verknappung der verfügbaren Organe geführt. Selbst<br />

10


eine verbesserte Allokation und die Lebendspende von Organen können den Bedarf <strong>bei</strong><br />

weitem nicht decken (Deutsche Stiftung Organtransplantation: www.dso.de). In Deutschland<br />

stehen ca. 12 000 Patienten auf der Warteliste für eine Nierentransplantation, pro Jahr sind<br />

aber nur knapp 4 000 Spendernieren verfügbar. Für andere Organe, wie Herz, Leber und<br />

Lunge können mit den verfügbaren Spenderorganen jeweils nur ein Drittel bis die Hälfte der<br />

notwendigen Transplantationen durchgeführt werden. Aufgrund der Organmangelsituation<br />

versterben in Deutschland pro Tag im Durchschnitt drei Patienten auf der Warteliste an<br />

akutem Organversagen.<br />

Eine mögliche Alternative stellt die Transplantation von Tierorganen (Xenotransplantation)<br />

dar (White, 1996; Bach, 1998, Niemann und Kues, 2000). Wesentliche Voraussetzungen für<br />

klinische Studien mit Xenotransplantaten sind, dass die immunologischen<br />

Abstoßungsreaktionen vermindert oder überwunden werden können, die Übertragung von<br />

Tierpathogenen ausgeschlossen werden kann, und eine weitgehende physiologische<br />

Kompatibilität zu dem entsprechenden Humanorgan besteht. Das Schwein stellt da<strong>bei</strong> die am<br />

besten geeignete Spezies dar, da die technischen Möglichkeiten <strong>zur</strong> genetischen Modifikation<br />

des Schweinegenoms vorhanden sind, die Haltung unter kontrollierten Bedingungen in<br />

spezifisch pathogenfreien (SPF) oder gnotobiotischen Anlagen etabliert ist, aufgrund der<br />

multiparen Reproduktion sehr schnell zahlreiche Nachkommen erzeugt werden können und<br />

eine grosse Ähnlichkeit in der Physiologie und Anatomie zum Menschen bestehen (Niemann<br />

und Kues, 2000).<br />

Bisher sind im wesentlichen transgene Schweine mit Expression humaner<br />

Komplementregulatoren (regulators of complement activation – RCA) (McCurry et al., 1995;<br />

Cozzi et al., 1997; Zaidi et al., 1998; Niemann et al., 2001), bzw. mit einem Gene knock-out<br />

der α-1,3-Galactosyltransferase (Dai et al. 2002; Lai et al., 2002) für Xeno-<br />

transplantationsmodelle produziert worden. Beide methodischen Ansätze setzen auf<br />

molekularer Ebene auf der Unterdrückung der hyperakut ablaufenden Komplement-<br />

vermittelten Abstoßung an (Yang-Guang und Sykes 2007; Zhong, 2007). Nach<br />

Xenotransplantation von RCA-transgenen, bzw Gal-knock-out Nieren und Herzen in nicht-<br />

humane Primaten wurden maximale Überlebenszeit von 179 Tagen erreicht, die medianen<br />

Überlebenszeiten liegen aber deutlich unter 60 Tagen (McGregor et al., 2004; Kuwaki et al.,<br />

2005; Yamada et al., 2005). Obwohl die geringen Transplantationszahlen keine statistisch<br />

signifikanten Aussagen erlauben, wird angenommen, dass zu den bisherigen genetischen<br />

Modifikationen weitere Transgene in das Schweinegenome eingeführt werden müssen, um die<br />

nachfolgenden Abstoßungsreaktionen, wie die akut vaskuläre Abstoßung (AVR) und die<br />

11


zelluläre Abstoßung kontrollieren können und dadurch längere Überlebenszeiten zu<br />

ermöglichen.<br />

2.4 Pharming: Produktion rekombinanter Proteine in der Milchdrüse<br />

2006 wurde von der European Medicine Agency (EMEA) erstmalig ein rekombinantes<br />

Protein aus der Milchdrüse als Medikament zugelassen (Niemann und Kues, 2007).<br />

Rekombinantes humanes Antithrombin III wird in der Milch transgener Ziegen produziert<br />

und ist zunächst für die prophylaktische Behandlung von Patienten mit Antithrombindefizienz<br />

vorgesehen. Inzwischen ist die Entwicklung fortgeschritten und zahlreiche weitere<br />

rekombinante Proteine aus der Milchdrüse von Kaninchen, Ziegen, Schafen oder Kühen<br />

befinden sich in der klinischen Prüfung (Tab.1).<br />

Tab. 1: Entwicklungsstand in der Produktion rekombinanter Proteine aus der Milchdrüse von <strong>Nutztieren</strong><br />

Rekombinantes Protein Hersteller Tierart Klinische Prüfung<br />

humanes Antithrombin III GTC Biotherapeutics Ziege EU: zugelassen<br />

(Produktname: ATryn) USA: Phase 3<br />

C1 Estera se Inhibitor Pharming Kaninchen Phase 3<br />

MM-093 (AFp) Merrimack und GTC Ziege Phase 2<br />

α-Glucosidase Pharming Kaninchen Phase 2 (gestoppt)<br />

humanes Wach stumshormon BioSidus Rind Vorklinisch<br />

Albumin GTC Biotherapeutics Rind Vorklinisch<br />

Fibrinogen Pharming Rind Vorklinisch<br />

Collagen Pharming Rind Vorklinisch<br />

Lactoferrin Pharming Rind Vorklinisch<br />

α-1 Antitrypsin GTC Biotherapeutics Ziege Vorklinisch<br />

Malaria-Vakzin GTC Biotherapeutics Ziege Vorklinisch<br />

CD137 monoklonaler AK GTC Biotherapeutics Ziege Vorklinisch<br />

(verändert nach Echelard et al., 2006)<br />

Die Milchdrüse weist für die rekombinante Proteinproduktion eine Reihe von Vorteilen auf.<br />

So können Prokaryoten keine posttranslatorischen Glykosylierungen durchführen, die für die<br />

Aktivität zahlreicher eukaryotischer Proteine notwendig sind. Die Produktion komplexer<br />

Proteine in eukaryotischen Zellkulturen ist häufig nur mit geringer Syntheseleistung und<br />

Ausbeute möglich und entsprechend kostenaufwendig. Die Milchdrüse ist für hohe<br />

metabolische Syntheseraten von Proteinen ausgelegt, Milch kann durch Melken einfach<br />

gewonnen und für die Aufreinigung der rekombinanten Proteine verwendet werden (Rudolph,<br />

1999; Echelard et al., 2006).<br />

12


2.5 Transgene Tiere für landwirtschaftliche Nutzungen<br />

Tabelle 2 zeigt eine Zusammenfassung der wichtigsten transgenen Ansätze <strong>zur</strong> Verbesserung<br />

von <strong>Nutztieren</strong> für die landwirtschaftliche Produktion. So sind transgene Schweine erstellt<br />

worden, die ein Desaturase-Gen im Fettgewebe produzieren und dadurch ein höheres<br />

Verhältnis von vielfach ungesättigten zu gesättigten Fettsäuren im Muskelfleisch aufweisen<br />

(Saeki et al., 2004; Lai et al., 2006). Insbesondere Omega-3-Fettsäuren wurden vermehrt<br />

gebildet. Eine Diät mit hohem Anteil ungesättigter Fettsäuren wird mit einem verminderten<br />

Risiko für Herzkreislauferkrankungen korreliert, und daher von den Gesundheitsbehörden<br />

empfohlen.<br />

Die Expression von bovinem Lactalbumin in der Milchdrüse von Schweinen führte zu einem<br />

höheren Laktose-Gehalt und erhöhter Milchproduktion, was wiederum zu besseren<br />

Entwicklungsraten der neugeborenen Ferkel führte (Wheeler et al., 2001). Durch die<br />

Expression bakterieller Phytase in der Speicheldrüse von Schweinen können diese Tiere sonst<br />

unverdaubares pflanzliches Phytat metabolisieren (Golovan et al., 2001). Dadurch wird die<br />

Phosphat-Belastung in der Gülle signifikant vermindert.<br />

Tab. 2 Transgene Nutztiere für landwirtschaftlich wichtige Merkmale<br />

Tierart Transgen/Konstrukt Phänotyp Methode Referenz<br />

Schwein Wachstumshormon/hMT-pGH verbessertes Wachstum, weniger Körperfett Mikroinjektion Nottle et al., 1999<br />

Schwein Insulin-like Wachstumsfaktor verbessertes Wachstum Mikroinjektion Pursel et al., 1999<br />

Schwein Desaturase/maP2-FAD erhöhte Werte für mehrfach ungesättigte Fettsäuren Mikroinjektion Saeki et al., 2004<br />

Schwein Desaturase/CAGGS-hfat-1 erhöhte Werte für mehrfach ungesättigte Fettsäuren somatischer Kerntransfer Lai et al., 2006<br />

Schwein Phytase/PSP-APPA Phytat-Metabolisierung Mikroinjektion Golovan et al., 2001<br />

Schwein Laktalbumin/genomische DNA erhöhter Laktosegehalt der Milch Mikroinjektion Wheeler et al. 2001<br />

Schwein Mx-Protein/mMx1-Mx Influenza-Resistenz Mikroinjektion Müller et al., 1992<br />

Schwein/Schaf IgA/a,K-a,K Pathogen-Resistenz Mikroinjektion Lo et al., 199<br />

Schaf Insulin-like Wachstumsfaktor Wollwachstum Mikroinjektion Damak et al., 1996<br />

Schaf PRNP knock-out-Konstrukt Scrapie-Resistenz (Tiere neonatal verstorben) somatischer Kerntransfer Denning et al. 2001<br />

Ziege Stearoyl-CoA Desaturase veränderte Milchzusammensetzung Mikroinjektion Reh et al., 2004<br />

Rind Kaseine/genomische DNA veränderte Milchzusammensetzung somatischer Kerntransfer Brophy et al., 2003<br />

Rind Lysostaphin Staphylococcus aureus-Resistenz somatischer Kerntransfer Wall et al., 2005<br />

(verändert aus Niemann und Kues, 2007)<br />

Der Einsatz dieser Tiere in der kommerziellen Fleischproduktion würde den kostspieligen<br />

Futterzusatz von rekombinanten Phytasen oder organischem Phosphor vermeiden und einen<br />

wesentlichen Beitrag zu verminderten Phosphat-Emissionen liefern.<br />

Über Gene knock-out und Gene knock-down-Techniken wurden Rinder mit einer Deletion<br />

des Prionprotein-Gens (PRNP), bzw. mit signifikant verminderten PRNP-mRNA-Werten<br />

erstellt (Kuroiwa et al., 2004; Golding et al., 2006; Richt et al., 2007). Diese sollten genetisch<br />

13


esistent gegen die bovine spongiforme Enzephalitis (BSE) sein, erste Ergebnisse aus in vitro<br />

Versuchen scheinen diese Resistenz zu bestätigen (Richt et al., 2007). BSE-resistente Rinder<br />

sind insbesondere für das Pharming von Interesse, da hier alle Produktionsschritte nach Good<br />

Manufacturing Practices (GMP) durchgeführt werden müssen, und Freiheit von möglichen<br />

Pathogenen gewährleistet sein muss.<br />

Bisher sind keine transgenen Tiere für die landwirtschaftliche Produktion zugelassen.<br />

Mittlerweile werden transgene Zierfische kommerziell in den USA, Hong Kong, Taiwan und<br />

Malaysia gehandelt (Niemann und Kues, 2007). Bei den Medaka (Oryzias latipes) und<br />

Zebrabärblingen (Danio rerio) werden verschiedene fluoreszente Proteine im Muskelgewebe<br />

exprimiert (Chou et al., 2003; Gong et al., 2003), die die Eigenfärbung der Fische<br />

überstrahlen. Ein Bericht der Food and Drug Administration der USA (FDA, 2003) stellte<br />

fest, dass diese transgenen Fische unbedenklich sind. Die transgenen Zierfische können<br />

möglicherweise zu einer veränderten Einstellung der Öffenlichkeit gegenüber gentechnisch<br />

veränderten Tieren <strong>bei</strong>tragen.<br />

2.6 Neuere Methoden der <strong>Transgenese</strong><br />

2.6.1 Klonen durch somatischen Kerntransfer<br />

1997 wurde erstmals das erfolgreiche Klonen eines Säugers über den Kerntransfer einer<br />

somatischen Zelle in eine entkernte Oozyte beschrieben (Wilmut et al., 1997). Damit wurde<br />

gezeigt, dass somatische Zellen grundsätzlich reprogrammierbar sind und nach<br />

Reprogrammierung eine komplette Ontogenese durchlaufen können. In der Folge wurde das<br />

somatische Klonen zu einer Schlüsseltechnik für die Grundlagenforschung, um Prozesse der<br />

Reprogrammierung, der Pluripotenz und der zellulären Differenzierung untersuchen zu<br />

können. Daneben wird das Klonen von genetisch wertvollen Tieren durch somatischen<br />

Kerntransfer mittlerweile kommerziell angeboten (Panarace et al., 2006).<br />

Ein weiterer wichtiger Aspekt ist, dass über das Klonen auch transgene Ansätze ermöglicht<br />

wurden, die zuvor im Nutztier nicht verfügbar waren (Cibelli et al., 1998). So können in der<br />

Maus über etablierte ES-Zellen Gene knock-out-Mutanten erhalten werden. Für Nutztiere sind<br />

bis heute keine echten ES-Zellen isoliert worden, und die klassische Knock-out-Technik ist<br />

daher nicht verfügbar. Mit dem somatischen Klonen eröffnet sich die Möglichkeit, zunächst<br />

einen Gene knock-out in primären somatischen Zellen durchzuführen (Schnieke et al., 1997;<br />

Dai et al., 2002; Lai et al., 2002), und vorselektionierte Zellen mit Knock-out dann als<br />

Kernspenderzellen einzusetzen (Kolber-Simonds et al., 2004; Abb. 4).<br />

14


Über die homologe Rekombination in primären Fibroblasten wurden Schafe mit einem Gen-<br />

knock-out für das Prion Protein (PRNP) geklont, die allerdings kurz nach der Geburt<br />

verstarben (Denning et al., 2001). Im Rind konnten vitale Tiere mit PRNP-knock-out erhalten<br />

werden (Kuroiwa et al., 2004; Richt et al., 2007). Im Schwein wurden Knock-out Tiere für die<br />

α-1,3- Galaktosyltransferase erstellt (Dai et al., 2002; Lai et al., 2002), die in der<br />

Xenotransplantation eine wesentliche Rolle spielt (siehe 2.3).<br />

Targeting-Konstrukt<br />

Elektroporation<br />

homologe Rekombination<br />

Zweite<br />

Durchmusterung<br />

auf ein „loss of<br />

heterozygosity“-<br />

Ereignis<br />

Isolation rekombinierter<br />

Zellklone<br />

Gen knock-out Tier<br />

Somatischer Kerntransfer<br />

Embryotransfer auf<br />

synchronisiertes<br />

Empfängertier<br />

15<br />

-/-<br />

entkernte Oozyte<br />

Abb.4 Erstellung von Gene knock-out Schweinen durch somatischen Kerntransfer (mod. aus Niemann u. Kues,<br />

2003)<br />

Über einen sogenannten Gene knock-in wurde der Blutgerinnungsfaktor IX gezielt in einen<br />

aktiven Genbereich des ovinen Genoms integriert und anschließend über den somatischen<br />

Kerntransfer vitale Schafe geklont (Schnieke et al., 1997).<br />

Das somatische Klonen hat entscheidende Vorteile für die <strong>Transgenese</strong> von <strong>Nutztieren</strong>.<br />

Allerdings sind neben dem hohen mikromanipulatorischen Aufwands für den Kerntransfer<br />

(Du et al., 2005) auch neue Probleme (Young et al., 1998; DeSousa et al., 2001) und Kosten<br />

entstanden. So war die ursprüngliche Ansicht, dass nur Zellen in der G0 oder G1-Phase des<br />

Zellzyklus als Kernspender geeignet seien (Wilmut et al., 1997). Dies scheint aber aus<br />

heutiger Sicht für den Klonerfolg nicht entscheidend zu sein. Auch die Frage, ob es


Unterschiede verschiedener somatischer Zellen in Bezug auf Differenzierungsgrad,<br />

Ausgangsgewebe, oder Passagenzahl in vitro gibt, ist nicht abschließend geklärt (Kues et al.,<br />

2000; Kues et al., 2005b). Bei der Maus konnte gezeigt werden, dass auch terminal<br />

ausdifferenzierte Zellen aus dem Nerven- und Blutsystem zu Klonnachkommen führen<br />

können (Hochedlinger und Jaenisch, 2002; Eggan et al., 2004), allerdings mit stark reduzierter<br />

Effizienz verglichen zu Fibroblasten. Porcine fetale Stammzellen und Fibroblasten zeigten<br />

ähnliche Kloneffizienzen (Hornen et al., 2007). Fibroblasten dienen in den meisten<br />

Klonversuchen als Kerndonoren. Fibroblasten sind relativ leicht zu kultivieren und benötigen<br />

keine Supplementation mit spezifischen Wachstumsfaktoren in der in vitro Kultur. Allerdings<br />

handelt es sich hier stets um heterogene Zellpopulationen mit beschränkter Proliferation in<br />

vitro (Hayflick, 1965).<br />

2.6.2 Knock-down in <strong>Nutztieren</strong> durch RNA interference<br />

Eine Alternative zum klassischen Gene knock-out ist der Gene knock-down, <strong>bei</strong> dem über<br />

RNA-Interferenz eine Degradation von spezifischen mRNAs erreicht wird (Fire et al., 1998;<br />

Jorgensen et al., 1998; Sharp, 2001). Bei Säugern werden da<strong>bei</strong> kurze doppelsträngige RNAs<br />

(17-25 Bp), sogenannte short interfering RNA (siRNA), bzw. short hairpin RNA (shRNA),<br />

vom RNA-Induced Silencing Complex (RISC) gebunden, und mRNAs mit komplementären<br />

Sequenzbereichen selektiv degradiert (Elbashir et al., 2001). Man vermutet, dass der RNAi-<br />

Mechnismus neben der Genregulation vor allem eine Funktion in der Virusresistenz hat.<br />

Durch das zelluläre Dicer-Enzym werden doppelsträngige RNAs, wie sie in RNA-Viren als<br />

Genom vorkommen, erkannt und zu siRNAs prozessiert. Diese „maßgeschneiderten“<br />

virusspezifischen siRNAs führen dann zu einer Degradation von viralen mRNAs.<br />

Für die <strong>Transgenese</strong> können siRNA-Expressionskonstrukte in das Genom eingebracht werden<br />

und den funktionellen Verlust eines spezifischen Transkriptes bewirken. Da keine<br />

Veränderung auf der Ziel-DNA-Sequenz notwendig ist, ist eine Zucht auf Homozygotie, wie<br />

<strong>bei</strong> Gene knock-out Ansätzen nicht notwendig (Tenenhaus-Dann et al., 2006). In<br />

Kombination mit neuartigen Vektoren für die <strong>Transgenese</strong>, wie Lentiviren, ist damit im<br />

Prinzip eine direkte genetische Veränderung von Produktionslinien von <strong>Nutztieren</strong> möglich.<br />

2.6.3 Virus-vermittelte <strong>Transgenese</strong><br />

Die ersten viralen Ansätze <strong>zur</strong> <strong>Transgenese</strong> <strong>bei</strong> Mäusen mit SV40 oder Moloney Leukemia-<br />

Virus (Jaenisch und Mintz, 1974; Jaenisch, 1976) und mit Retroviren <strong>bei</strong> Rindern (Haskell<br />

and Bowen, 1995; Chan et al., 1998) führten zu Tieren, die die Fremd-DNA im Genom<br />

16


integriert trugen; die Fremd-DNA war jedoch stillgelegt (DNA silencing) und die Tiere damit<br />

phänotypisch unverändert. In den letzten Jahren zeigte sich, dass Lentiviren, eine Unterklasse<br />

der Retroviren, deutlich bessere Vektoren für die Erstellung transgener Mäuse und Nutztiere<br />

darstellen (Pfeifer et al., 2002; Hofmann et al., 2003 u. 2004; Whitelaw et al., 2004). So<br />

können transgene Tiere direkt durch Injektion der Lentiviren in den perivitellinen Raum von<br />

Zygotenstadien erzeugt werden. Da<strong>bei</strong> wurde eine sehr hohe Rate transgener Nachkommen<br />

beobachtet, in denen eine aktive Expression der Transgene festgestellt wurde. Offenbar<br />

kommt es durch die veränderte Zusammensetzung der terminalen Repeat-Bereiche <strong>bei</strong> den<br />

Lentiviren nicht zu einem DNA-Silencing im Säugergenom. Ein weiterer Faktor ist, dass nach<br />

lentiviraler Infektion multiple Integrationen auf verschiedenen Chromosomen stattfinden.<br />

Inwiefern dies zu einer höheren Rate von Integrationsmutationen führt, muss noch abgeklärt<br />

werden.<br />

2.6.4 Konditionale Mutagenese<br />

Eine weitere Verfeinerung des Methodenrepertoires stellt die konditionale Mutagenese dar.<br />

So kann ein konventioneller Gene knock-out zunächst einmal nicht räumlich oder zeitlich<br />

beschränkt werden. Für viele <strong>Untersuchungen</strong> ist es aber wünschenswert, ein Gen zu einem<br />

definierten Zeitpunk aus- oder anzuschalten, um z.B. kompensatorische Mechanismen, die<br />

sich <strong>bei</strong> einem permanenten Knock-out entwickeln können, zu vermeiden. Auch können so<br />

Gene, die essentiell für die Embryogenese sind und Letalmutationen darstellen, gezielt nach<br />

der Geburt ausgeschaltet werden. Die Methoden für die konditionale Mutagenese sind im<br />

wesentlichen in der Maus entwickelt worden. So ist die Cre-spezifische DNA-Rekombinase<br />

des Bakteriophagen P1 auch in Säugerzellen aktiv und kann spezifische Rekombinationen von<br />

DNA-Sequenzen durchführen, die von als „loxP“ bezeichneten Erkennungssequenzen<br />

flankiert sind (Sauer, 1998). Das 38kD Cre Protein benötigt da<strong>bei</strong> keine weiteren Cofaktoren.<br />

Je nach Orientierung der loxP-Erkennungssequenzen zueinander, kommt es da<strong>bei</strong> <strong>zur</strong><br />

Exzision der flankierten DNA-Region oder <strong>zur</strong> Inversion (Sauer und Henderson, 1988).<br />

Durch das Einführen der loxP-Erkennungssequenzen in ein Zielgen und die zusätzliche<br />

Präsenz des Cre-Gens unter der Kontrolle eines gewebespezifischen Promoters kann ein<br />

beschränkter gewebespezifischer Knock-out erreicht werden (Orban et al., 1992; Gu et al.,<br />

1994; St-Onge et al., 1996). Für diesen Ansatz sind zwei Linien transgener Mäuse notwendig.<br />

Die erste Linie trägt da<strong>bei</strong> die Cre-Rekombinase unter der Kontrolle eines Promoters mit dem<br />

gewünschten Expressionsprofil. In der zweiten Linie wird die DNA-Zielregion über<br />

homologe Rekombination in ES-Zellen mit loxP-Sequenzen flankiert. Durch die loxP-<br />

17


Sequenzen wird die normale Expression nicht beeinträchtigt. Durch Linien-Verpaarung<br />

werden <strong>bei</strong>de Modifikationen in einem Genom kombiniert. So zeigen DNA-Polymerase ß-<br />

defiziente Mäuse einen embryonal-lethalen Phenotyp. Durch konditionale Mutagenese war es<br />

möglich, lebensfähige Nachkommen mit loxP-flankiertem DNA-Polymerase ß-Gen zu<br />

erhalten und selektiv einen Knock-out in T-Zellen zu generieren (Gu et al., 1994). Eine<br />

wichtige Voraussetzung, um mögliche embryonal-letalen Phenotypen vorhersagen zu können,<br />

ist die Analyse der Genexpression in frühen Entwicklungsstadien, z.B. über in situ-<br />

Hybridisierungstechniken (Kues und Wunder, 1992; Kues et al., 1995a, b). Der Nachteil des<br />

Cre/loxP-System und anderer Rekombinasen ist, dass die Methodik relativ aufwendig ist, und<br />

in der Regel nur einmal eine Rekombination erfolgt.<br />

2.6.5 Induzierbare Promotoren und konditionale Expressionskontrolle<br />

Eine Möglichkeit für die konditionale Expressionkontrolle sind sogenannte induzierbare<br />

Promotoren, die mittels exogener Substanzen reguliert werden können. Hierzu gehören der<br />

Metallothioneinpromoter (Nottle et al., 1999), sowie Hitzeschock- und Steroid induzierbare<br />

Promotoren. Jedoch hat sich gezeigt, dass diese induzierbaren Promotoren in der <strong>Transgenese</strong><br />

nur beschränkt nutzbar sind, da sie teilweise eine geringe Induktion zeigten, bzw. die<br />

Induktoren unspezifische physiologische Effekte zeigten (Yarranton, 1992). Ein wesentlicher<br />

Fortschritt wurde mit dem Tetracyclin-regulierten System erzielt (Gossen und Bujard, 1992).<br />

Hier<strong>bei</strong> werden wiederum zwei Transgene in das Genom eingeführt (Abb.5). Ein Konstrukt<br />

wird da<strong>bei</strong> von einem tetracycylin-sensitiven Promoter (ptTA) kontrolliert. Sobald der<br />

rekombinante Transaktivator (tTA) an den ptTA bindet, wird das Transgen exprimiert. Der tTA<br />

wird von einer zweiten Kassette kodiert; er ist aus zwei Domänen zusammengesetzt: einem<br />

Virusprotein 22 (VP22)-Anteil, mit transaktivierender Funktion und einem Protein-Anteil, der<br />

Tetracyclin binden kann (Gossen et al., 1995). Durch die Bindung von Tetracyclin kommt es<br />

zu einer Konformationsänderung des Gesamtproteins und die DNA-Bindungsfähigkeit geht<br />

verloren.<br />

In dieser Situation wird der pTA nicht mehr aktiviert und das Zieltransgen abgestellt (tet-off-<br />

System). Durch Mutagenese wurden auch tTA-Varianten entwickelt, die erst durch<br />

Tetracyclin-Bindung aktiv an den Promoter binden (Tet-on-system, Furth et al., 1994). In<br />

<strong>bei</strong>den Fällen werden zwei unabhängige Transgene benötigt, was für den Nutztierbereich ein<br />

wesentliches Hindernis darstellt. Es gibt aber auch erfolgversprechende Ansätze, <strong>bei</strong>de<br />

Transgen-Kassetten auf einem Konstrukt zu verbinden (Schultze et al., 1996) und damit eine<br />

vereinfachte konditionale Expression zu erzielen.<br />

18


Abb.5. Konditionale Expressionskontrolle im tet-off-System<br />

Für die konditionale Expressionskontrolle werden zwei Konstrukte benötigt. Das tetracycline responsive element<br />

(TRE) ist vor einem minimalen Promoter (pminCMV) lokalisiert, dieser wird ohne Aktivierung nicht exprimiert.<br />

Der Transaktivator (tTA) wird von dem zweiten Konstrukt kodiert, der tTA bindet an die TRE-Sequenz und<br />

aktiviert die Expression. In Gegenwart von Tetracyclin, bzw. Doxycyclin (DOX) wird der tTA durch eine<br />

Konformationsänderung inaktiviert, und die Expression des TRE-Konstruktes wird abgestellt. (mod. aus Fa.<br />

Clontech Handbuch).<br />

2.7 Genomprojekte und ihre Bedeutung für die <strong>Transgenese</strong><br />

Seit ca. 1970 werden Methoden <strong>zur</strong> schnellen Sequenzierung kurzer DNA-Fragmenten<br />

entwickelt (Sanger et al., 1973; Maxam und Gilbert 1977; Sanger et al., 1977). Pro Reaktion<br />

konnten maximal 400-800 Basen gelesen werden. Aufgrund des notwendigen zeitlichen und<br />

technischen Aufwands wurden nur ausgesuchte DNA-Bereiche sequenziert. Vollständig<br />

wurden Plasmide und kleine virale Genome mit bis zu 50 Kilobasenpaaren (kBp) sequenziert.<br />

Ab ca. 1990 wurde die Komplettsequenzierung des humanen Genoms (HUGO-Projekt)<br />

initiiert, zu diesem Zeitpunkt wurden die Kosten für die Sequenzierung einer Base mit ca. 1.-<br />

US $ angenommen, hochgerechnet für ein haploides Humangenom mit 3 x 10 9 Bp ergibt dies<br />

3 Milliarden US $. Aufgrund der extrem hohen Kosten und der verfügbaren<br />

Sequenzierverfahren, die zu dieser Zeit im wesentlichen in „Handar<strong>bei</strong>t“ durchgeführt<br />

19


wurden, wurde das HUGO-Projekt von vielen Wissenschaftlern zunächst als illusorisch<br />

angesehen. Die Entwicklung automatisierter Sequenziermaschinen erlaubte innerhalb weniger<br />

Jahre deutlich leistungsfähigere Hochdurchsatzsequenziermethoden und wesentlich reduzierte<br />

Kosten (Bennett et al., 2005). Gegenwärtig liegen die Kosten für eine Sequenzierung eines<br />

Säugergenoms mit einfacher Abdeckung in der Größenordnung von ca. 100 000 US $. Im<br />

Jahr 2001 wurden die ersten Rohentwürfe der humanen Genomsequenzierung publiziert<br />

(Lander et al., 2001). Aufgrund der Humangenomsequenz wird die Anzahl der protein-<br />

kodierenden Gene heute mit ca. 25 000 annotiert (Goodstadt und Ponting, 2006). Einem<br />

Drittel der Gene kann da<strong>bei</strong> eine eindeutige Funktion zugeordnet, <strong>bei</strong> einem weiteren Drittel<br />

gibt es Vermutungen über die Funktion, und für ein Drittel liegt keine Annotationen vor.<br />

Die durch das HUGO-Projekt begonnene Entwicklung führte <strong>zur</strong> Genomsequenzierung von<br />

zahlreichen Organismen. So sind gegenwärtig > 1000 Prokaryoten-Genome und > 49<br />

Eukaryoten-Genome sequenziert. Unter anderem auch die Genome von Rind (Adelson, 2008),<br />

Hund (Lindblad-Toh et al., 2005) und Pferd (www.broad.mit.edu.mammals.horse). Für das<br />

Schwein liegen Teilgenomsequenzierungen vor. Ein freier Zugang zu umfassenden und<br />

annotierten Genominformationen ist über die Internet-Seiten von ENSEMBL<br />

(www.ensembl.org), National Center for Biotechnology Information (www.pubmed.org) und<br />

der Universität von Californien, Santa Cruz (http://genome.ucsc.edu) möglich. So sind z.B.<br />

auf der ENSEMBL-Seite umfassende Genominformationen von 49 Eukaryoten-Spezies,<br />

einschließlich von kompletten und Teilgenomsequenzen von <strong>Nutztieren</strong> aufgear<strong>bei</strong>tet und<br />

direkt zugänglich. Zahlreiche Funktionen können benutzerdefiniert dargestellt und extrahiert<br />

werden, z.B. Chromosomenlokalisation annotierter Gene, deren Exon-Intron-Struktur,<br />

bekannte Splice-Varianten, Einzelnukleotidpolymorphismen (SNP), die Lokalisation von<br />

CpG-reichen-Regionen (CpG-islands), der GC-Gehalt, die Position von repetitiven<br />

Elementen. Neben zahlreichen weiteren Informationen kann direkt auf die DNA-Sequenz<br />

zugegriffen werden. Noch nicht sequenzierte Bereiche, bzw. Sequenzdaten mit geringer<br />

Qualität von Eukaryoten erstrecken sich im wesentlichen auf hochrepetitive<br />

Genomabschnitte, z.B. Centromeren, Telomeren, sowie auf das Y-Chromosom.<br />

Beispielhaft zeigt Abb. 6 ein 100 kBp-Abschnitt auf dem Rinderchromosom 2, auf dem das<br />

Myostatin-Gen (GDF8) lokalisiert ist. Myostatin ist ein negativer Regulator des<br />

Muskelwachstums, und eine Reduktion oder der Verlust der Myostatinaktivität führt <strong>bei</strong><br />

Rindern zum Doppellender-Phänotyp mit Muskelhypertrophie (Grobet et al., 1997; Kambadur<br />

et al., 1997; McPherron und Lee, 1997). Bei zahlreichen Fleischrinderrassen, wie Blaue<br />

20


Belgier und Piedmonteser, sind durch konventionelle Züchtung Mutationen und Deletionen<br />

im Myostatin-Gen selektioniert worden.<br />

Abb. 6. Schematische Darstellung des Genlocus für das bovine Myostatin-Gen (GDF8). Ein 100 kBp-Bereich<br />

wurde aus Ensembl (www.ensembl.org) extrahiert.<br />

Die Regulation des Muskelwachstums ist <strong>bei</strong> Säugern offenbar hochkonserviert, so wurde <strong>bei</strong><br />

Texel-Schafen mit hypertrophen Muskelwachstum eine Mutation im GDF8-Gen entdeckt, die<br />

keine Veränderung des Leserasters bedingt, aber im GDF8-Transkript eine falsche<br />

Zielsequenz für eine microRNA darstellt. Durch die RNAi-vermittelte Degradation des<br />

Transkripts wird eine signifikante Verringerung der Myostatin-Spiegel mit entsprechendem<br />

Phänotyp hervorgerufen (Clop et al., 2006). Beim Menschen ist eine Mutation beschrieben<br />

worden, die das Spleißen der Prä-mRNA und den Leseraster verändert und ebenfalls zu einem<br />

Phänotyp mit Muskelhypertrophie führt (Schuelke et al., 2004). Über ENSEMBL können die<br />

entsprechenden syntenen Genomabschnitte dieser Spezies direkt miteinander verglichen<br />

21


werden, bzw. in das Design von genetischen Veränderungen einfließen. Zusammen mit der<br />

Verfügbarkeit von cDNA- und genomischen DNA-Klonen zahlreicher Spezies in öffentlichen<br />

Genbanken (z.B. Ressourcen Zentrum Berlin, jetzt ImaGenes) ergeben sich für die<br />

<strong>Transgenese</strong> bisher ungeahnte Möglichkeiten, so dass auch komplexe Veränderungen des<br />

Genoms (Kuroiwa et al., 2002; Grosse-Hovest et al., 2004) wesentlich vereinfacht werden.<br />

2.8 Epigenetik<br />

In der Epigenetik geht es um das Verständnis, wie Information über die Genregulation, die<br />

nicht direkt in der DNA-Sequenz kodiert ist, von einer Zell- oder Organismen-Generation in<br />

die nächste gelangt (Reik, 2002). Spezifische epigenetische Prozesse umfassen komplexe<br />

Genregulations-Mechanismen, unter anderem das Imprinting, das Gen-Silencing und die<br />

Reprogrammierung. Die griechische Vorsilbe epi hat mehrere Bedeutungen, wie „nach“,<br />

„hinterher“ oder „zusätzlich, über“. So beschreibt die Epigenetik alle Prozesse in einer Zelle,<br />

die als „zusätzlich zu“ den Prozessen und Informationen der Genetik gelten. Die Gesamtheit<br />

aller epigenetischen Mechanismen wird als Epigenom bezeichnet.<br />

Es gibt verschiedene Mechanismen, die in der Epigenetik wirksam werden. Die<br />

bestuntersuchten in Säugern sind DNA-Methylierung und Histon-Modifikationen. Bei der<br />

DNA-Methylierung wird die Cytosin-Base in CG-Dinukleotiden an der 5-Position methyliert<br />

(Abb. 7). Somit wird eine „fünfte“ Base eingeführt, da sich Cytosin und 5-Methyl-Cytosin<br />

funktionell unterscheiden. Die Promoterbereiche von vielen Genen umfassen CpG-reiche<br />

Regionen, die als „CpG islands“ bezeichnet werden. In aktiven Genen sind die CpG islands<br />

unmethyliert, in inaktiven Genen sind die CpG islands und weitere Promoterbereiche häufig<br />

Cytosin-methyliert. Auch <strong>bei</strong>m Imprinting ist in der Regel das inaktive Allel an den CpG-<br />

Positionen Cytosin-methyliert. Während der frühen Embryogenese kommt es zu einem<br />

Neusetzen des größten Teils der DNA-Methylierung. So wird in der befruchteten Oozyte<br />

zunächst die DNA des paternalen und dann die DNA des maternalen Vorkerns demethyliert.<br />

Aufgrund der zeitlichen Aufeinanderabfolge geht man davon aus, dass die Demethylierung<br />

des paternalen Vorkerns aktiv und die des maternalen Vorkerns passiv erfolgt (Mayer et al.,<br />

2000 a, b). Bis zum Blastozystenstadium erfolgt dann eine aktive Remethylierung. Diese<br />

Neusetzung der DNA-Methylierungen wird heute als ein Prozess der Reprogrammierung der<br />

Gameten-DNA verstanden, um ein funktionelles Embryonenepigenom zu formen.<br />

Ausgenommen von diesen DNA-Methylierungsänderungen im frühen Embryo sind geprägte<br />

Gene (imprinted genes), die erst in der Gametogenese neu methyliert werden.<br />

22


Abb. 7. Die DNA-Methyltransferase (DNA-MTase) katalysiert die Methylierung von Cytosin-Basen in CpG-<br />

Dinukleotiden zu 5-Methyl-Cytosin (m 5 CpG). Da<strong>bei</strong> dient S-Adenosylmethionin (AdoMet) als Methylgruppen-<br />

Donor (aus Heby, 1995).<br />

Wichtige Histonmodifikationen sind Azetylierungen, Methylierungen oder<br />

Phosphorylierungen an spezifischen Aminosäuren der Seitenkette des Histons H3 (Koipally et<br />

al., 1999; Senawong et al., 2003; Eilertsen et al., 2008; Nandy et al., 2008). Je nach Histon-<br />

Modifikation kann so ein komprimierter, oder ein aufgelockerter Packungszustand des<br />

Chromatins (DNA/Histon-Komplex) erreicht werden (Yao et al., 2001). Im komprimierten<br />

Zustand ist die DNA unzugänglich und Transkriptionsfaktoren können nicht oder nur<br />

vermindert binden; im aufgelockerten Zustand können Transkriptionsfaktoren leichter binden<br />

und die Transkription starten. Aktive DNA-Bereiche mit aufgelockerter DNA werden auch<br />

als Euchromatin bezeichnet, komprimierte inaktive DNA-Bereiche als Heterochromatin.<br />

Es ist noch nicht eindeutig geklärt, inwieweit CpG-Methylierung und Histon-Modifikationen<br />

miteinander interagieren und in welcher Reihenfolge die DNA und Histonmodifikationen<br />

erfolgen. Es gilt aber als gesichert, dass <strong>bei</strong>de Mechanismen zum regionalen Packungszustand<br />

der DNA (Eu- und Heterochromatin) und <strong>zur</strong> differentiellen Genaktivität <strong>bei</strong>tragen (Reik,<br />

2002). Dementsprechend sind sie für die <strong>Transgenese</strong> und die Expressionsprofile in<br />

transgenen Tieren von essentieller Bedeutung.<br />

23


3.0 Fragestellungen<br />

Die Zielsetzung der vorliegenden Ar<strong>bei</strong>t war es, verbesserte Methoden <strong>zur</strong> <strong>Transgenese</strong> von<br />

<strong>Nutztieren</strong> zu etablieren. Da<strong>bei</strong> wurde insbesondere auf die Spezies Schwein fokussiert, da im<br />

Rahmen von BMBF und DFG geförderten Projekten Fördermittel für die Generierung<br />

transgener Schweine als Xenotransplantationsmodell eingeworben werden konnten.<br />

Methodisch wurden da<strong>bei</strong> zwei Ansätze verfolgt. Die „klassische“ Mikroinjektion von DNA-<br />

Expressionskonstrukten in den Vorkern von Zygoten und der somatische Kerntransfer<br />

(Klonen) von Kernspenderzellen in entkernte Metaphase II-Oocyten. Für den somatischen<br />

Kerntransfer wurden insbesondere die Gewinnung und Kultur der Kernspenderzellen aus<br />

adulten und fetalen Geweben untersucht und die Zellkulturen detailliert charakterisiert.<br />

Um eine konditionale Expressionskontrolle der eingesetzten DNA-Transgene zu erreichen,<br />

wurden neuartige tetracyclin-regulierte Expressionskassetten (tet-off) eingesetzt. Damit sollte<br />

die Möglichkeit einer von aussen-regulierbaren Transgenexpression im Schwein evaluiert<br />

werden. Bisher war dieser Ansatz in keiner Nutztierspezies untersucht worden.<br />

24


4.0 Ergebnisse und Diskussion<br />

Im Ergebnisteil wird auf folgende Aspekte eingegangen:<br />

1) Charakterisierung von primären Kernspenderzellen und Isolierung somatischer Stamm-<br />

zellen (Anger et al., 2003; Kues et al., 2005a, b),<br />

2) Gene knock-down-Versuche in transgenen Embryonen (Iqbal et al., 2007),<br />

3) Generierung und detaillierte Untersuchung von transgenen Schweinen mit konditioneller<br />

Genexpression (tet-off: Kues et al., 2006a; Deppenmeier et al., 2006; Niemann u. Kues, 2003;<br />

Niemann u. Kues, 2007).<br />

Die anderen Aspekte dieser wissenschaftlichen Thematik können in den Nachdrucken der<br />

Orginalpublikationen (in Kapitel 10.0) nachgelesen werden (Schätzlein et al., 2004; Hoelker<br />

et al., 2005; Yadav et al., 2005; Hornen et al., 2007; Dieckhoff et al., 2007, 2008; Kues u.<br />

Niemann, 2004; Niemann et al., 2005; Niemann et al., 2007; Kues et al., 2008).<br />

4.1 Primäre somatische und Stamm-Zellen als Donorzellen für den Kerntransfer<br />

Nach der erstmaligen Beschreibung eines erfolgreichen Klonexperiments mit somatischen<br />

Zellen (Wilmut et al., 1997), ging man davon aus, dass die Zellzyklusphase der<br />

Kernspenderzelle entscheidend für die erfolgreiche Weiterentwicklung der rekonstruierten<br />

Embryonen sei. Insbesondere wurde die Ansicht favorisiert, dass primäre Zellen in der G0-<br />

Phase des Zellzyklus am besten geeignet seien (Wilmut et al., 1997; Wells et al., 1999).<br />

Fibroblasten, die direkt aus adultem Gewebe isoliert werden, liegen überwiegend in der G0-<br />

Phase vor. In der Kultur primärer Zellen wird durch den Zusatz von Serum und<br />

Wachstumsfaktoren zum Kulturmedium eine starke Proliferationsstimulation erzeugt<br />

(Connell-Crowley et al., 1998), so dass in der Regel maximale Zellwachstums- und<br />

Teilungsraten vorliegen. Primäre humane Fibroblasten können in Kultur bis zu 50<br />

Zellteilungen durchführen, bevor sie in Zellseneszenz übergehen (Hayflick, 1965). Für die<br />

Anreicherung von Zellkulturen in bestimmten Zellzyklusphasen gibt es eine Reihe von<br />

Methoden, die aber meistens für transformierte Zellen optimiert sind (Holley et al., 1968;<br />

Reed, 1997; Zetterberg et al., 1995). Für die Zellzyklussynchronisierung primärer<br />

Fibroblasten von Nutztierspezies waren keine etablierten Protokolle vorhanden; dies gilt<br />

insbesondere für unerwünschte oder toxische Effekte der möglichen Verfahren.<br />

Als Kerndonorzellen für den somatischen Kerntransfer wurden hier überwiegend primäre<br />

Fibroblastenkulturen angelegt und detailliert charakterisiert. Da<strong>bei</strong> wurden insbesondere die<br />

Homogenität der Fibroblasten, die Proliferationsfähigkeit, die Transfizierbarkeit mit DNA-<br />

25


Konstrukten und die Zellzyklussynchronisierung untersucht (Kues et al., 2002; Anger et al.,<br />

2003; Kues et al., 2005a).<br />

Es wurden Methoden für die Isolation primärer Fibroblasten aus fetalen und adulten Gewebe<br />

von <strong>Nutztieren</strong> entwickelt und die etablierten Zellkulturen hinsichtlich der<br />

Proliferationsfähigkeit in Kultur und der Zellzyklussynchronisierung detailliert untersucht.<br />

Die primären Zellen konnten genetisch verändert werden (Dieckhoff et al., 2007) und wurden<br />

erfolgreich für Klonversuche verwendet (Schaetzlein et al., 2004; Hornen et al., 2007;<br />

Dieckhoff et al., 2008). Es konnte gezeigt werden, dass porcine fetale Fibroblasten<br />

grundsätzlich Zellzyklus synchronisierbar sind. Maximal wurden Anreicherungen in der<br />

G0/G1-Phase des Zellzyklus von 80-90 % der Zellen erreicht, da<strong>bei</strong> wurden sowohl<br />

Serumreduktion als auch chemische Inhibitoren des Zellzyklus (Pedrail et al., 1980; Kitagawa<br />

et al., 1994), wie Aphidicolin und Butyrolactone eingesetzt (Kues et al., 2000). Die optimale<br />

Zellzyklussynchronisierung wurde nach 48 Stunden Kultur in serumreduzierten<br />

Zellkulturmedien erreicht; damit sind wesentlich kürzere Synchronisationszeiten möglich, als<br />

zunächst beschrieben (Wilmut et al., 1997; Wells et al., 1999). Bei längeren<br />

Serumentzugszeiten (als 48 Stunden) kam es zunehmend zum apoptotischen Zelltod, wo<strong>bei</strong><br />

eine direkte Korrelation mit der Serumreduktion und der Zeitdauer der Behandlung<br />

festgestellt werden konnte. In den primären Fibroblasten wurde da<strong>bei</strong> eine unkonventionelle<br />

Form der Apoptose beobachtet (Kues et al., 2002). Die Messung der Zellzyklusphasen durch<br />

Fluorescence activated cell sorting (FACS) wurde durch Reverse transcriptase-polymerase<br />

chain reaction (RT-PCR)-Bestimmungen der Transkripte des zellzyklus-spezifisch regulierten<br />

Gens Polo-ähnlicher Kinase (Polo-like kinase) auf molekularer Ebene untermauert (Kues et<br />

al., 2000; Kues et al., 2002).<br />

In den Primärkulturen porciner Fibroblasten wurde ein neuer Stammzelltyp gefunden, der mit<br />

Hilfe der Explantatmethode und der Kultur unter hohen Zelldichten angereichert werden<br />

konnte (Abb. 8). Diese fetalen somatischen Stammzellen (FSSCs) konnten aus fetalem<br />

Bindegewebe von Schwein, Maus, Rind und Schaf isoliert werden (Kues et al., 2005a), in<br />

adultem Gewebe kommen sie offenbar in wesentlich geringerer Frequenz vor.<br />

26


eseeding of<br />

outgrowing cells<br />

Tissue explant<br />

Standard culture<br />

Fibroblast monolayer<br />

confluent,<br />

mitotic inactive<br />

High density culture<br />

Colonies of FSSCs<br />

colony growth on monolayer,<br />

mitotic active,<br />

Oct-4, Stat3, TNAP positive<br />

27<br />

Fetus<br />

Abb. 8. Isolation fetaler somatischer Zellen (FSSC, aus Kues et al., 2005b).<br />

Suspension culture<br />

Anchorage independent growth<br />

reaggregation,<br />

growth as spheroids<br />

Die FSSCs wachsen als Kolonien auf einem autologem Nährzellrasen aus Fibroblasten, sie<br />

exprimieren Stammzellmarker (Yeom et al., 1996), wie OCT4, alkaline Phosphatase (AP) und<br />

STAT3, zeigen kein Hayflick-Limit und nach Injektion muriner FSSCs in Maus-Blastozysten<br />

konnten chimäre Feten erhalten werden, in denen die injizierten Zellen zu somatischen<br />

Geweben und der Keimbahn <strong>bei</strong>trugen (Abb.9). Inwieweit der beobachtete Chimärismus<br />

durch Zellfusion (Ying et al., 2002; Wang et al., 2003) bedingt ist, muss in weitergehenden<br />

<strong>Untersuchungen</strong> abgeklärt werden. Porcine FSSCs wurden zudem erfolgreich im Kerntransfer<br />

eingesetzt (Hornen et al., 2007). Aufgrund ihrer höheren Potenz sind FSSCs möglicherweise<br />

bessere Kernspender im Klonverfahren. Durch ihre längere Proliferationsfähigkeit in Kultur<br />

vereinfachen sie genetische Modifikationen und die Isolierung von Zellklonen.<br />

Zur Zeit wird davon ausgegangen, dass FSSCs einen gewebe-spezifischen Stammzelltyp<br />

darstellen. Fibroblasten stellen einen Hauptzelltyp in Bindegewebe dar, wo sie eine Reihe von<br />

Funktionen erfüllen (Moulin et al., 1999; Chang et al., 2002; Han et al., 2002), unter anderem<br />

sind sie an der Regeneration beteiligt. In Fetalstadien von Säugern verheilen<br />

Hautverletzungen ohne Narbenbildung (Moulin et al., 1999), was als ein indirekter Hinweis<br />

auf eine höhere Zellplastizität von Bindegewebszellen in diesem Stadium gewertet werden<br />

kann.


A Wt chimeric fetus day 15.5 p.c.<br />

B<br />

Chimeric fetuses control OG2/Rosa26<br />

Abb.9. Zellchimärismus in Mausfeten nach Injektion von Rosa26/OG2-transgenen FSSCs in Blastozysten.<br />

A) Durch histologischen Nachweis der LacZ-Expression (Pfeile) kann die Beteiligung der injizierten Zellen an<br />

den Fetalorganen gezeigt werden (rechter Embryo). Links ein Kontrollembryo.<br />

B) Nachweis von GFP-positiven Zellen (OG2) in chimärer Keimleiste. Rechts als Positivkontrolle die<br />

Keimleiste eines OG2 transgenen Fetuses (aus Kues et al., 2005a).<br />

28


4.2 Gene knock-down Versuche in transgenen Embryonen<br />

Das Transkriptom in frühen Säugerembryonen wird zunächst durch maternale mRNAs<br />

geprägt, die in der Oozyte angereichert vorliegen (Schultz, 2002). Im Zygotenstadium erfolgt<br />

keine Transkription des embryonalen Genoms, die Proteinsynthese ist in diese Phase von den<br />

maternalen Transkripten abhängig. Die Aktivierung des embryonalen Genoms (embryonic<br />

genome activation) erfolgt spezies-spezifisch <strong>bei</strong> murinen Embryonen am Ende des<br />

Zygotenstadiums, <strong>bei</strong> humanen und porcinen Embryonen im 4-Zellstadium und <strong>bei</strong> bovinen<br />

Embryonen im 8-Zellstadium (Telford et al., 1990).<br />

In der Literatur war gezeigt worden, dass der RNAi-Mechanismus in Säugerembryonen<br />

funktionell ist, da<strong>bei</strong> wurden überwiegend lange doppelsträngige RNA (500-1000 bp)<br />

injiziert, und verschiedene Gentranskripte, wie E-Cadherin, Mos, Plat, Oct4, Coxa, Cox5b<br />

und Cox6b selektiv degradiert (Svoboda et al., 2000; Wianny et al., 2000; Plusa et al., 2004;<br />

Paradis et al., 2005; Nganvongpanit et al., 2007; Cui et al., 2006). Die doppelsträngigen<br />

RNAs werden durch das zelleigene Enzym Dicer in siRNAs prozessiert, die dann in dem<br />

eigentlichen RNAi-Mechanismus den Abbau von komplementären Transkripten verursachen<br />

(Svoboda et al., 2000). Bei diesem Verfahren ist also eine relativ hohe Beladung der<br />

Embryonen mit siRNAs notwendig, von denen nur ein Teil funktionell ist, zusätzlich kann es<br />

zu unspezifischen Effekten kommen. Zudem kann <strong>bei</strong> Genen, die ein biphasisches<br />

Expressionsmuster zeigen, d.h. die sowohl im maternalen Transkriptom vorliegen, als auch<br />

im embryonalen Genom transkribiert werden, nicht ausgeschlossen werden, dass maternale<br />

und embryonale Transkripte unterschiedlich sensitiv für den RNAi-Mechanismus sind. So ist<br />

bekannt, dass maternale Transkripte eine Halbwertzeit im Bereich von Tagen, embryonale<br />

Transkripte dagegem im Bereich von Minuten besitzen (Karlson et al., 2005).<br />

In transgenen Mäusen wurde getestet, ob im RNAi-Mechanismus eine differentielle<br />

Degradierung von maternalen und embryonalen Transkripten auftritt. Dazu wurde eine<br />

funktionelle synthetische siRNA (22 bp) gegen GFP in transgene Mausembryonen injiziert.<br />

(Iqbal et al., 2007). Um die Degradation von maternalen und embryonalen Transkripten<br />

verfolgen zu können, wurden transgene Mäuse (OG2), die das Green Fluorescent Protein<br />

(GFP) unter Kontrolle des keimbahnspezifischen Promoters Oct4 exprimieren (Szabo et al.,<br />

2002), verwendet. In transgen-homozygoten Embryonen wird GFP in allen Blastomeren<br />

exprimiert, also sowohl maternal als auch embryonal. Bei Verpaarungen mit Wildtypmäusen<br />

entstehen hemizygot-transgene Embryonen, <strong>bei</strong> denen die GFP-Expression davon abhängt, ob<br />

sie das Transgen-Allel von der mütterlichen (OG2 mat/- ) oder der väterlichen (OG2 pat/- ) Seite<br />

29


geerbt haben. OG mat/- Embryonen zeigen eine kontinuierliche GFP-Expression von der Oocyte<br />

bis <strong>zur</strong> Blastozyste, während OG pat/- -Embryonen erst ab dem 4-8-Zellstadium GFP-positiv<br />

werden (Abb.10). Die RNAi der GFP-Expression kann somit an lebenden Embryonen<br />

selektiv für maternale und embryonale Transkripte evaluiert werden.<br />

Abb. 10. Vererbungsabhängige Expression von Oct4-GFP in hemizygoten Mausembryonen.<br />

A) Embryonen mit paternal vererbten Transgen (OG2 pat/- ) zeigen GFP-Expression ab dem 8-Zellstadium,<br />

Embryonen mit maternal vererbten Transgen (OG2 mat/- ) zeigen in allen Stadium von Zygote bis Blastocyste<br />

GFP-Fluoreszenz. B) Schematischer Aufbau der siRNA, die verwendete siRNA ist mit dem Fluorochrom<br />

Rhodamine konjugiert. C) Injektion der siRNA in OG2 mat/- -Zygote unter Hellfeldbedingungen (oben),<br />

Fluoreszenzanregung für GFP (mitte) und Fluoreszenzanregung für Rhodamin (unten) (aus Iqbal et al., 2007).<br />

Durch Mikroinjektion einer short interfering RNA (siRNA), die gegen die GFP mRNA<br />

gerichtet ist, wurden die unterschiedlichen RNAi-Kinetiken <strong>bei</strong> Embryonen mit maternaler,<br />

bzw. embryonaler GFP-Expression ermittelt.<br />

Da<strong>bei</strong> zeigte sich, dass die Expression vom paternalen Allel nahezu komplett inhibiert werden<br />

kann; es war keine GFP-Fluoreszenz nachweisbar. In den OG mat/- Embryonen konnte eine<br />

Reduktion der GFP-mRNA um ca. 95% nachgewiesen werden, aufgrund der hohen<br />

Proteinstabilität von GFP war aber erst in Blastozysten eine relative Verringerung der GFP-<br />

Fluoreszenz feststellbar (Abb.11). Diese Untersuchung zeigt, dass eine einmalige Injektion<br />

von siRNA im Zygotenstadium ausreichend ist, um bis zum Blastozystenstadium eine RNA-<br />

Interferenz zu erreichen. Durch die Verwendung einer synthetischen siRNA, können<br />

unspezifische Effekte, die <strong>bei</strong> langen doppelsträngigen RNAs beobachtet werden, weitgehend<br />

30


vermieden werden. Um zu bestimmen, wie lange der Effekt einer einmaligen siRNA-<br />

Injektion, anhält, müssten in Folgeuntersuchungen Postimplantations-Stadien untersucht<br />

werden.<br />

(OG2 pat /-) (OG2 mat /-)<br />

Abb.11. RNAi in hemizygoten Mausembryonen.<br />

Nach Injektion der GFP-spezifischen siRNA (Fig.10) in hemizygote OG2 transgene Zygoten wurden diese in<br />

vitro kultiviert. 2-Zell, Morula- und Blastozystenstadien von OG2 pat/- (links) und OG2 mat/- (rechts) sind im<br />

Hellfeld und unter GFP-Fluoreszenzanregung dargestellt (aus Iqbal et al., 2007).<br />

Im Gegensatz zu den relativ aufwendigen Verfahren eine RNA-Interferenz in<br />

postimplantatorischen Embryonen zu induzieren (Mellitzer et al., 2002; Caligari et al., 2004;<br />

Soares et al., 2005), ist das siRNA-Injektionsverfahren wesentlich einfach durchzuführen und<br />

zeitlich mindestens bis zum Blastozystenstadium effektiv.<br />

31


4.3. Generierung transgener Schwein mit konditioneller Transgenexpression<br />

Über die DNA-Mikroinjektion in einen Vorkern von Schweinezygoten wurden transgene<br />

Linien mit Expression des humanen Komplementregulators CD59 erstellt (Niemann und<br />

Kues, 2003). Als regulative Sequenz wurde der Cytomegalovirus (CMV) Immediate Early-<br />

Promotor gewählt, der als einer der stärksten Promotoren in der Zellkultur gilt.<br />

Abb. 12. Generation transgener Schwein mit Expression von hCD59<br />

A) Minigen-Konstrukt mit hCD59 cDNA, B) Injektion des aufgereinigten Konstruktes in einen Vorkern einer<br />

Schweinezygote, C) Transgenes F0-Tier, D) Organspezifische Expression von hCD59 in einem F1-Tier, E)<br />

Nachweis von hCD59 auf der Zelloberfläche von porcinen Fibroblasten und Endothelzellen, F) funktioneller<br />

Nachweis von hCD59 in Schweinefibroblasten durch verminderte Lyse durch Humankomplement, G)<br />

funktioneller Nachweis von hCD59 in einem Nierenperfusionsystem mit Humanblut (aus Niemann und Kues,<br />

2003).<br />

Es wurden 5 Linien transgener Schweine mit diesem Konstrukt etabliert. Abb. 12 fasst die<br />

Ergebnisse aus der Linie 5 zusammen. Bei Expressionsuntersuchungen zeigte sich, dass eine<br />

32


starke Expression von hCD59 in Pankreas, Niere, Herz und Haut vorlag. Auch primäre<br />

Endothelzellen, die aus der Aorta isoliert wurden, wiesen eine, wenn auch geringe Expression<br />

von hCD59 auf. In Perfusionen von Schweinenieren mit Humanblut, die in Zusammenar<strong>bei</strong>t<br />

mit der Medizinischen Hochschule Hannover durchgeführt wurden, konnte gezeigt werden,<br />

dass die vorhandene hCD59-Expression ausreichte, um wesentlich verbesserte<br />

Perfusionszeiten zu erreichen und die hyperaktute Abstoßung zu unterdrücken (Niemann et<br />

al., 2001).<br />

Feingewebliche <strong>Untersuchungen</strong> zeigten, dass eine starke Variegation der hCD59-Expression<br />

in allen CMV-transgenen Linien vorlag (Abb. 13). So waren in der Niere einzelne Zellen<br />

stark anfärbbar, während benachbarte Zellen keine nachweisbare hCD59-Expression<br />

aufwiesen.<br />

Abb. 13. Variegierte Transgenexpression in CMV-hCD59 transgenen Schweinen (L5).<br />

A), C), E) immunhistologischer Nachweis von hCD59 in Herz, Pankreas und Nierengewebe eines Linie 5-<br />

Tieres. B), D) und F) Kontrollgewebe eines wt-Tieres (aus Niemann et al., 2001). H) Vergrößerter Ausschnitt<br />

einer immunhistologischen Anfärbung für hCD59 in der Niere.<br />

In einem weitergehenden Ansatz wurde versucht, diese Limitationen zu überwinden, dazu<br />

wurde eine neuartige Tetracyclin-regulierbare Expressionskassette (NTA) zusammen mit der<br />

Ar<strong>bei</strong>tsgruppe von Dr. Hauser (GBF/jetzt Helmholtz-Zentrum für Infektionsforschung)<br />

eingesetzt. Tetracyclin-regulierbare Expressionssysteme (tet-off, bzw. tet-off) wurden für die<br />

konditionale Genregulation in Zell-Linien und transgenen Mäusen entwickelt (Gossen und<br />

Bujard, 1992; Furth et al., 1994; Lewandowski, 2001; Corbel und Rossi, 2002), um die<br />

Limitationen zu überwinden, die mit der konditionalen Genregulation auf der Basis von<br />

Metall- oder Steroid-sensitiven Promotoren bestanden (Mayo et al., 19982; Lee et al., 1981;<br />

Miller et al., 1989; Pursel et al., 1989). Das originale Tet-System besteht aus zwei<br />

Expressionskassetten, einem Transaktivator-Konstrukt und einem Transaktivator-regulierten<br />

33<br />

H


Konstrukt, die in zwei Mauslinien etabliert und dann durch Verpaarung in einem Genom<br />

kombiniert werden (Lewandowski, 2001). Da der Zeitaufwand für die Verkreuzung von zwei<br />

<strong>Nutztieren</strong>linen wesentlich höher ist als <strong>bei</strong> der Maus, wurde hier ein Ansatz gewählt, der<br />

<strong>bei</strong>de funktionellen Elemente auf einem Konstrukt vereinigt und somit nur die Erstellung<br />

einer transgene Linie erfordert (Kues et al., 2006a). Dazu wurde ein bicistronisches Konstrukt<br />

entworfen (Abb. 14), <strong>bei</strong> dem eine mRNA gebildet wird, die einen Komplementregulator<br />

(hCD59 oder hDAF) und den Tet-Transaktivator (NTA-RCA) kodiert. Durch die basale<br />

Expression des Konstruktes kommt es dann zu einer autoregulativen Hochregulation der<br />

Expression; die Transgenexpression ist durch Zugabe von Tetracyclin oder Doxycyclin<br />

abstellbar, bzw. regulierbar (Abb. 14). Als eigentliche Transgene wurden die humanen<br />

Komplentregulatoren CD59, bzw. DAF jeweils in das erste Cistron kloniert. Es wurden also<br />

zwei NTA-Konstrukte mit den Kombinationen CD59-tTA und DAF-tTA verwendet. Der<br />

Vorteil dieser Konstrukte gegenüber den Orginalsystem (Gossen und Bujahr, 1992) besteht<br />

darin, dass nur ein Transgenkonstrukt in das Genom eingebracht wird und die Verpaarung<br />

von zwei Linien entfällt.<br />

+<br />

Tet-O<br />

p tTA<br />

TATA box<br />

RCA Transactivator<br />

RCA IRES tTA pA<br />

34<br />

-<br />

Dox<br />

Inactivated<br />

exogenous Dox<br />

Abb. 14. Autoregulative bicistronische Expressionskassette (NTA).<br />

Der tetracyclin-regulierte Promoter besteht aus sieben Tet-Operator-Sequenzen und einem minimalen Promoter<br />

mit einer TATA-Box. Die bicistronische Kassette wird durch eine IRES-Sequenz getrennt, das erste Cistron<br />

kodiert einen RCA (entweder CD59 oder DAF) und das zweite Cistron den Transaktivator. Nach Transkription<br />

führt der tTA durch Bindung an den eigenen Promoter zu einer autoregulativen Hochregulation, durch Zugabe<br />

von Doxycyclin wird der tTA inaktiviert (verändert aus Kues et al., 2006a).<br />

Die Funktionalität der NTA-Kassette wurde durch Transfektion in murine NIH-3T3 Zellen<br />

und in porcine Swine testis epithelium (STE)-Zellen getestet Da<strong>bei</strong> konnte sowohl in<br />

(AAA) n


Zellpopulationen als auch in Einzelzellklonen eine volle Aktivität der NTA-Kassette<br />

beobachtet werden, dass heisst durch basale Transkription konnte eine Hochregulation der<br />

humanen Komplementregulatoren CD59, bzw. DAF erzielt werden. Die Transgenexpression<br />

war durch Tetracyclin-Supplementation des Zellkulturmediums reversibel abstellbar (Kues et<br />

al., 2006a).<br />

Insgesamt wurden 10 transgene Linien (hemizygot) von Schweinen mit diesem Konstrukt<br />

über Injektion der linearisierten DNA in Zygotenvorkerne erstellt und charakterisiert (Tab. 3;<br />

Kues et al., 2006a, Niemann und Kues, 2007). Da<strong>bei</strong> zeigte sich, das in den hemizygoten<br />

Linien die NTA-Kassette nahezu komplett stillgelegt vorlag (gene silencing); also trotz<br />

Vorhandenseins des Transgens im Genom keine Transkription erfolgte. Gene silencing wird<br />

auch in Mauslinien beobachtet, dass allerdings 10 unabhängige Linien mit unterschiedlichen<br />

Kopienzahlen des Transgens (1-3) und unterschiedlichen Integrationsorten eine (nahezu)<br />

komplettes gene silencing aufwiesen, deutet daraufhin, dass entweder das Konstrukt selbst<br />

diesen Effekt hervorrief, oder in Schweinen deutlich sensitivere Mechanismen zum Erkennen<br />

und Stilllegen von Fremd-DNA als in Mäusen aktiv sind. Die Expression der NTA-Kassette<br />

konnte in hemizygoten Tieren nur in der Skelettmuskulatur gefunden werden, und zwar waren<br />

offenbar zufallsmäßig einzelne Fasern in der feingeweblichen Untersuchung Transgen-positiv<br />

(Abb. 15B). Auffallend war, dass dieses Muster <strong>bei</strong> 9 von den untersuchten Linien vorlag<br />

(Tab.3), linienabhängig waren 5-20% der Muskelfasern hCD59 oder hDAF positiv. Die<br />

positiven Fasern waren da<strong>bei</strong> durchgehend stark positiv angefärbt (Kues et al., 2006a).<br />

Tab. 3: Transgenexpression in Schweinen mit einer und zwei NTA-Kassetten<br />

Transgene Schweine mit einer NTA-Kassette Transgene Schweine mit zwei NTA-Kassetten<br />

DAF-NTA CD59-NTA DAF-NTA x CD59-NTA<br />

L12 L13 L14 L15 L16 L17 L19 L20 L21 L22 L13 x L20 (n = 7) L15 x L20 (n = 2)<br />

DAF CD59 DAF CD59<br />

Herz - - - - - - - - - - - + - -<br />

Leber - - - - - - - - - - - ++ - -<br />

Lunge - - - - - - - - - - - + - -<br />

Niere - - - - - - - - - - - + - -<br />

Pankreas - - - - - - - - - - - ++ - -<br />

Skel. Muskel - ++ + + + + + ++ + + ++ ++++ + ++<br />

Thymus - - - - - - - - - - n.d. n.d. n.d. n.d.<br />

Lymphozyten - - n.d. n.d. n.d. n.d. n.d. - n.d. n.d +* ++* - -<br />

Fibroblasten n.d. - n.d. n.d. n.d. n.d. n.d. - n.d. n.d. - + n.d. n.d.<br />

Zusammengefasste Expressionsdaten aus Northern blot, Immunhistologie und FACS-Messungen; (+)-Zeichen zeigen Expressionshöhe an, (-) keine<br />

Expression detektierbar, (n.d.) nicht untersucht und (*) nach Mitogenstimulation.<br />

Primäre Fibroblasten aus hemizygoten Tieren exprimierten kein Transgen, auch nach<br />

Transfektion mit einem tTA-Konstrukt wurde keine Expression gefunden. Entsprechende<br />

35


Befunde wurden nach Injektion des tTA-Plasmids in die Skelettmuskulatur gemacht. Dagegen<br />

konnte in Mitogen-stimulierten Lymphozyten die Expression von hCD59 und hDAF<br />

eindeutig gezeigt werden (Kues et al., 2006a).<br />

Durch Fütterung mit Doxycyclin war die Expression in den Muskelfasern abgestellbar. Durch<br />

Resequenzierung der Konstrukte aus genomischer DNA der transgenen Tiere, konnte<br />

ausgeschlossen werden, dass die beobachteteten Expressionsmuster durch Deletion von<br />

regulativen Sequenzen hervorgerufen wurden.<br />

Ansätze, die transgenen Linien auf Transgen-Homozygotie zu verpaaren, waren nicht<br />

erfolgreich. Offenbar aufgrund eines bereits hohen Inzuchtgrades im vorhandenen Bestand<br />

führten Geschwisterverpaarungen zu sehr kleinen Würfen (0-3 Ferkel) mit lebensschwachen<br />

Ferkeln. Daher wurden Tiere aus Linien mit NTA-CD59 und NTA-DAF-Transgenen<br />

verpaart. In den normal großen Würfen wurden in etwa die erwarteten Verhältnisse von 25 %<br />

Nachkommen mit <strong>bei</strong>den Kassetten, 50% mit einer Kassette und 25% Wildtyp gefunden<br />

(Kues et al., 2006a). Bei Tieren mit zwei NTA-Kassetten wurde linien-abhängig die<br />

Hochregulation von einer oder <strong>bei</strong>der Kassetten gefunden (Tab. 3; Kues et al., 2006a;<br />

Niemann und Kues, 2007). Da<strong>bei</strong> kam es einer Ausweitung der exprimierenden Zellen, so<br />

wurden regelmäßig bis zu 100 % positive Skelettmuskelfasern gefunden, und in zahlreichen<br />

anderen Organen, wie Leber, Haut, Herz und Niere lagen positive Zellen vor (Abb. 15A, C).<br />

In den NTA-doppelttransgenen Tieren wurde eine Doxycyclin-abhängige Transgenexpression<br />

gefunden (Abb. 16). Nach oraler Applikation von 3.3 mg Doxycyclin pro kg Lebendgewicht<br />

und Tag wurde bereits nach 2 Tagen eine massive Reduktion der bicistronischen Transkripte<br />

gefunden. In Abb. 16 wurden von drei doppelt-transgenen Geschwistertieren eines mit einem<br />

Placebo und zwei mit Doxycyclin gefüttert. Die gewählte Doxycyclin-Dosis von 3.3 mg/<br />

Kilogramm Lebendgewicht und Tag wurde aufgrund der Literaturangaben für die<br />

Doxycyclin-Transgenregulation <strong>bei</strong> Mäusen (Furth et al., 1994; Kistner et al., 1996; Zhu et<br />

al., 2001) extrapoliert; sie liegt deutlich unter der für Schweine empfohlenen antimikrobiell<br />

wirksamen Dosis von 12 mg/kg und Tag. Die Transgenexpression in transgenen Schweinen<br />

blieb nach dem Absetzen der Doxycyclin-Gabe für mindestens 50 Tage abgeschaltet, bevor<br />

eine Reexpression messbar ist (Abb. 16). Eine mögliche Erklärung ist, dass Doxycyclin<br />

offenbar effektiv in Knochen zwischengespeichert und dann zeitverzögert in den Kreislauf<br />

<strong>zur</strong>ückgeführt werden kann (Stepensky et al., 2003). In transgenen Mäusen wurden<br />

unterschiedliche Wirksamkeiten von Doxycyclin in Abhängigkeit vom Mausstamm<br />

beschrieben (Robertson et al., 2002). In nachfolgenden Dosis-Wirksamkeitsuntersuchungen<br />

36


ei doppelttransgenen Schweine wurde die zweimalige Gabe einer Doxycyclin-Dosis von 0,6<br />

mg/kg als ausreichend für eine sichere Expressionsregulation gefunden (nicht gezeigt).<br />

Abb. 15. Transgen-Expression in einfach und doppelttransgenen Schweinen (NTA).<br />

A) Northern blot mit CD59-Sonde; (1, 2, 3), Muskelbiopsien von NTA-CD59/NTA-DAF doppelttransgenen<br />

Tieren; (4,5,6), Muskelbiopsien von NTA-CD59 einzeltransgenen Tieren; (7, 8) wt-Kontrollen; (9, 10) Positiv-<br />

Kontrollen von CMV-CD59 transgenen Tieren. In den Proben aus einzeltransgenen NTA-Tieren (4, 5, 6) ist die<br />

CD59-Expression erst nach längerer Expositionszeit sichtbar.<br />

B) Immunhistologischer Nachweis von CD59 und DAF in einzeltransgenen Tieren (NTA) im Skelettmuskel.<br />

Einzelne Muskelfasern sind angefärbt. Oberer Reihe Muskellängsschitte, untere Reihe Muskelquerschnitte. In<br />

anderen Organen wurden keine Zellen mit Transgenexpression gefunden.<br />

C) Immunhistologischer Nachweis von CD59 und DAF in doppeltransgenen Tieren. In doppeltransgenen Tieren<br />

aus Linie 13 mit Linie 20-Verpaarungen ist die Muskulatur komplett positiv für CD59 gefärbt (M.) in Leber<br />

(Le.), Lunge (Lu.) und Pankreas (P.) sind einzelne Zellen positiv. Dagegen ist DAF nur in einzelnen<br />

Muskelfasern nachweisbar. In doppelttransgenen Tieren aus der Verpaarung L15 mit L20 (unterster Reihe, M.)<br />

ist die Expression von CD59 und DAF auf einzelne Muskelfasern beschränkt, wo<strong>bei</strong> einige Faser nur CD59-<br />

Expression aufweisen (Pfeile, bzw. gestrichelte Kreise) (mod. aus Kues et al., 2006a).<br />

Die Analyse auf DNA-Methylierung durch Bisulfitsequenzierung in Promoterbereichen der<br />

NTA-Konstrukte ergab, das einzeltransgene Tiere eine nahezu komplette Methylierung an<br />

CpG-Dinukleotiden aufwiesen, während in den Promoterbereichen der doppeltransgenen<br />

Tieren eine ca. 50% Reduktion der Methylierung gefunden wurde. In Zellklonen, Maus 3T3<br />

und porcinen STE-Zelllinien, waren die Promotorsequenzen methylierungsfrei. Offenbar wird<br />

die NTA-Kassette während der Ontogenese stillgelegt (gene silencing); dieses manifestiert<br />

37


sich einer DNA-Methylierung der Promotorsequenzen. In den doppelttransgenen Tieren kann<br />

dieses Silencing unterdrückt werden; offenbar kann durch den autoregulativen Charakter der<br />

Expressionskontrolle das Silencing verhindert werden. Möglicherweise geschieht dies direkt<br />

durch die Bindung des tTA, was in Abhängigkeit von der basalen Expressionshöhe ein<br />

Silencing bestimmter Integranten dauerhaft verhindern würde. Gegenwärtig laufen Versuche,<br />

durch weitere Verpaarung mehr als zwei NTA-Kassetten in einem Tier zu kombinieren. Das<br />

hier gefundene bimodale Expressionsmuster wurde mittlerweile auch in Zell-Linien für<br />

bicistronische Konstrukte beschrieben (May et al., 2008).<br />

Abb. 16. Konditionale Transgenexpression im Schwein<br />

Drei NTA-doppelttransgenen Geschwistertieren wurde über 6 Tage ein Placebo oder Doxycyclin (3.3 mg/kg)<br />

gefüttert. Zu verschiedenen Zeitpunkten wurden Muskelbiopsien genommen und im Northern blot auf die<br />

Expression von hCD59, hDAF und des porcinen ß-Actin (ACTB)-Gens untersucht (aus Kues et al., 2006a).<br />

Die Bedeutung der NTA-transgenen Tiere liegt im wesentlichen in den neuen Einblicken in<br />

die Regulation von Gensequenzen im Säugergenom. In diesem Tiermodell konnte ein<br />

stillgelegtes Transgen erstmals wieder reaktiviert werden. Durch die Verpaarung zu<br />

doppelttransgenen Tieren wird offenbar in einem kritischen Entwicklungsfenster das<br />

Silencing der Transgene verhindert. Die detaillierte Identifikation der beteiligten<br />

Mechanismen und zeitlichen Abläufe in weitergehenden <strong>Untersuchungen</strong> wird für eine<br />

optimierte <strong>Transgenese</strong>, insbesondere in <strong>Nutztieren</strong> von entscheidender Bedeutung sein.<br />

38


5.0 Zusammenfassung<br />

In der vorliegenden Ar<strong>bei</strong>t wurden experimentelle Ansätze <strong>zur</strong> <strong>Transgenese</strong> von <strong>Nutztieren</strong><br />

durchgeführt. Aus primären Fibroblastenkulturen von <strong>Nutztieren</strong> wurden Stammzellen<br />

(FSSCs) abgeleitet, die wesentliche Vorteile für eine genetische Modifikation haben und die<br />

Effizienz des somatischen Kerntransfers erhöhen können. In transgenen Embryonen wurde<br />

mit einer synthetischen siRNA ein allel-spezifischer Gene knock-down gezeigt. Es wurden<br />

transgene Schweine mit konditionaler Expressionskontrolle (tet-off) erstellt und detailliert<br />

charakterisiert. Da<strong>bei</strong> zeigten Tiere aus den hemizygoten Linien ein nahezu komplettes<br />

Silencing der Transgen-Kassetten. Durch die Verpaarung zu doppelttransgenen Tieren<br />

konnten die Transgenkassetten reaktiviert werden. Da<strong>bei</strong> konnten die NTA-Kassetten von<br />

allen vier untersuchten Integrationsorten reaktiviert werden, die Reaktivierung war da<strong>bei</strong> in<br />

linien-spezifischen Kombinationen möglich. Damit wurde erstmalig eine wirksame<br />

konditionale Expressionskontrolle im Nutztier gezeigt; und es wurde nachgewiesen, dass<br />

epigenetische Mechanismen, insbesondere die DNA-Methylierung, mit der konditionalen<br />

Expression interferieren können. Dieser Befund ermöglicht ein wichtiges<br />

grundlagenwissenschaftliches Verständnis der Genomregulation in <strong>Nutztieren</strong> und fügt sich in<br />

das gegenwärtige Modell der Reprogrammierung und des Neusetzens von<br />

Methylierungsmustern in der frühen Embryogenese ein. Für verbesserte Ansätze <strong>zur</strong><br />

<strong>Transgenese</strong> sind die hier erhobenen Befunde von wichtiger Bedeutung.<br />

39


6.0 Literaturverzeichnis<br />

Adelson, D.L. 2008. Insights and applications from sequencing the bovine genome.<br />

Reproduction, Fertility and Development 20:54-60.<br />

Anger, M., Kues, W.A., Klima, J., Mielenz, M., Motlik, J., Carnwath, J.W. and Niemann, H.<br />

2003 Cloning and cell cycle-specific regulation of Polo-like kinase in porcine fetal<br />

fibroblasts. Molecular Reproduction and Development 65, 245-253.<br />

Avery, O.T., MacLeod, C.M., McCarty, M. 1944. Studies on the chemical nature of the<br />

substance inducing transformation of pneumococcal types. J. Exp. Med. 79, 137-158.<br />

Bach, F. H. 1998. Xenotransplantation: Problems and prospects. Annu. Rev. Med. 49, 301-<br />

310.<br />

Bennett, S.T., Barnes C., Cox, A., Davies, L, Brown, C. 2005. Toward the 1,000 dollars<br />

human genome. Pharmacogenomics 6, 373-82.<br />

Bird, A. 1992. The essentials of DNA methylation. Cell. 70, 5-8.<br />

Brackett, B.G., Baranska, W., Sawicki, W., Koprowski, H. 1971.Uptake of heterologous<br />

genome by mammalian spermatozoa and its transfer to ova through fertilization. Proc Natl<br />

Acad Sci U S A. 68, 353-7.<br />

Brem, G., Brenig, B., Goodman, H. M., Selden, R. C., Graf, F.,Kruff, B., Springmann, K.,<br />

Hondele, J., Meyer, J., Winnacker, E.L. & Kr.ausslich, H. 1985. Production of transgenic<br />

mice, rabbits and pigs by microinjection into pronuclei. Zuchthygiene 20, 251—252.<br />

Brophy, B., Smolenski, G., Wheeler, T., Wells, D., L'Huillier, P., and Laible, G. 2003. Cloned<br />

transgenic cattle produce milk with higher levels of b-casein and k-casein. Nat.<br />

Biotechnol. 21, 157–162.<br />

Caligari, F., Marzesco, A.M., Kittler, R. et al., 2004. Tissue-specific RNAi in postimplantation<br />

mouse embryos using directional electroporation and whole embryo culture.<br />

Differentiation 72, 92-102.<br />

Capecchi, M.R. 1989. The new mouse genetics: altering the genome by gene targeting.<br />

Trends Genetics 5, 70-76.<br />

Cibelli, J. B., Stice, S. L., Golueke, P. L., Kane, J. J., Jerry, J., Blackwell, C., Ponce de Leon,<br />

F. A., and Robl, J. M. 1998. Transgenic bovine chimeric offspring produced from somatic<br />

cell-derived stem like cells. Nat. Biotechnol. 16, 642-646.<br />

Chan, A. W., Homan, E. J., Ballou, L. U., Burns, J. C., and Bremel, R. D. 1998. Transgenic<br />

cattle produced by reverse-transcribed gene transfer in oocytes. Proc. Natl. Acad. Sci.<br />

USA 95, 14028-14033.<br />

Chandler, L.A., Jones, P.A. 1988. Hypomethylation of DNA in the regulation of gene<br />

expression. Dev Biol 5, 335-49.<br />

40


Chang, H.Y., Chi, J.T., Dudoit, S., et al., 2002. Diversity, topographic differentiation and<br />

positional memory in human fibroblasts. Proc. Natl. Acad. Sci. USA 99, 12877-11882.<br />

Chou, C.Y., Horng, L.S., Tsai, H.J., 2001. Uniform GFP-expression in transgenic medaka<br />

(Oryzias latipes) at the F0 generation. Transgenic Res 10, 303-315.<br />

Clements, J. E., Wall, R. J., Narayan, O., Hauer, D., Schoborg, R., Sheffer, D., Powell, A.,<br />

Carruth, L. M., Zink, M. C., and Rexroad, C. E. 1994. Development of transgenic sheep<br />

that express the visna virus envelope gene. Virology 200, 370-380.<br />

Clop, A., Marcq, F., Takeda, H., Pirottin, D., Tordoir, X., Bibé, B. 2006. A mutation creating<br />

a potential illegitimate microRNA target site in the myostatin gene affects muscularity in<br />

sheep. Nature Genetics 38, 813-8.<br />

Connell-Crowley, L., Elledge, S.J., Harper, J.W. 1998. G1 cyclin-dependent kinases are<br />

sufficient to initiate DNA synthesis in quiescent human fibroblasts. Curr. Biol 8, 65-68.<br />

Cozzi, E., and White, D. J. G. 1995. The generation of transgenic pigs as potential organ<br />

donors for humans. Nat. Med. 1, 964–966.<br />

Cozzi, E., Tucker, A.W., Langford, G.A. et al., 1997. Characterization of pigs transgenic for<br />

human decay –accelerating factor. Transplantation 64,1383-1390.<br />

Cui, X.S., Li, X.Y., Jeong, Y.J., Jun, J.H. Kim, N.H. 2006. Gene expression of cox5a, 5b or<br />

6b1 and their roles in preimplantation mouse embryos. Biol Reprod. 72, 601-610.<br />

Dai, Y., Vaught, T. D., Boone, J., Chen, S.H., Phelps, C. J., Ball, S., Monahan, J. A., Jobst,<br />

P.M., McCreath, K. J., Lamborn, A. E., Cowell-Lucero, J. L., Wells, K. D., Colman, A.,<br />

Polejaeva, I. A., and Ayares, D. L. 2002. Targeted disruption of the alpha1,3galactosyltransferase<br />

gene in cloned pigs. Nat. Biotechnol. 20, 251–255.<br />

Dallas, A. and Vlassow, A. 2006 RNAi: A novel antisense technology and its therapeutic<br />

potential. Med. Sci. Monit.12, RA67-74.<br />

Damak, S., Jay, N.P., Barrel, G.K., Bullock, D.W. 1996. Improved woll production in<br />

transgenic sheep expressing insulin-like growth factor 1. Biotechnology 14, 185-188.<br />

Denning, C., Burl, S., Ainslie, A., Bracken, J., Dinnyes, A., Fletcher, J., King, T., Ritchie, M.,<br />

Ritchie, W. A., Rollo, M., de Sousa, P., Travers, A., Wilmut, I., and Clark, A. J. 2001.<br />

Deletion of the alpha(1,3)galactosyl transferase (GGTA1) gene and the prion protein (PrP)<br />

gene in sheep. Nat. Biotechnol. 19, 559-562.<br />

Deppenmeier, S., Bock, O., Mengel, M., Niemann, H., Kues, W., Lemme, E., Wirth, D,<br />

Wonigeit, K., and Kreipe, H. 2006. Health status of transgenic pig lines expressing human<br />

complement regulator protein CD59. Xenotransplantation 13, 345-356.<br />

De Sousa, P.A., King, T., Harkness, L., Young, L.E., Walker, S.K., Wilmut, I. 2001.<br />

Evaluation of gestational deficiencies in cloned sheep fetuses and placentae. Biol Reprod.<br />

65, 23-30<br />

41


Dieckhoff, B., Karlas, A., Hofmann, A., Kues, W. A., Petersen, B., Pfeifer, A., Niemann, H.,<br />

Kurth, R., and Denner, J. 2006. Inhibition of porcine endogenous retroviruses (PERV) in<br />

primary porcine cells by RNA interference using lentiviral vectors. Arch. Virol. 152, 629-<br />

634.<br />

Dieckhoff, B., Petersen, B., Kues, W.A., Kurth, R., Niemann, H., Denner, J. 2008.<br />

Knockdown of procine endogenous retrovirus (PERV) expression by PERV-specific<br />

shRNA in transgenic pigs. Xenotransplantation 15, 36-45<br />

Dodart, J.C., May, P. Overview on rodent models of Alzheimer's disease. Curr Protoc<br />

Neurosci. 2005 Nov;Chapter 9:Unit 9.22.<br />

Du, Y., Kragh, P.M., Zhang, X., Purup, S., Yang, H., Bolund, L., Vajta, G. 2005. High overall<br />

in vitro efficiency of porcine handmade cloning (HMC) combining partial zona digestion<br />

and oocyte trisection with sequential culture. Cloning Stem Cells 7, 199-205.<br />

Echelard, Y., Ziomek, C., Meade, H. 2006. Production of recombinant therapeutic proteins in<br />

the milk of transgenic animals. BioPharm Intern. 2, 1-6.<br />

Eggan, K., Baldwin, K., Tackett, M., Osborne, J., Gogos, J., Chess, a., Axel, R., Jaenisch, R.,<br />

2004. Mice cloned from olfactory sensory neurons. Nature 428, 44-49.<br />

Eilertsen, K.J., Floyd, Z., Gimble, J.M..2008. The epigenetics of adult (somatic) stem cells.<br />

Crit Rev Eukaryot Gene Expr. 18, 189-206<br />

Elbashir, S.M., Harborth, J., Lendeckel, W., Yalcin, A., Weber, K. Tuschel, T. 2001.<br />

Duplexes of 21-nucleotide RNAs mediate RNA interference in cultured mammalian cells.<br />

Nature 411, 494-498.<br />

Fire, A., Xu , S., Montgomery, M.K., Kostas, S.A., Driver, S.E., Mello, C. 1998. Potent and<br />

specific genetic interference by double-stranded RNA in Caenorhabditis elegans. Nature<br />

391, 809-811.<br />

Food and Drug Administration USA 2003. FDA statement regarding Glofish.<br />

http://www.fda.gov/bbs/topics/NEWS/2003/NEW00994.html.<br />

Furth, P.A., Onge, L., Böger, H., Gruss, P., Gossen, M., Kistner, A., Bujard, H.,<br />

Hennighausen, L. 1994. Temporal control of gene expression in transgenic mice by a<br />

tetracycline-responsive promoter. Proc. Natl. Acad. Sci USA 91, 9302-9306.<br />

Glover, D.M. (Eds). 1985. DNA cloning. A practical approach. Vols. 1-2. IRL Press, 1985.<br />

Götz, J., Probst, A., Ehlers, E., Hemmings, B., Kues, W. 1998. Delayed embryonic lethality in<br />

mice lacking protein phosphatase 2a catalytic subunit C alpha. Proc. Natl. Acad. Sci. USA<br />

95, 12370-12375.<br />

Götz, J., Kues, W. 1999. The role of PP2A catalytic subunit C alpha in embryogenesis:<br />

Evidence from sequence analysis and localization studies. Biol Chem. 380, 1117-1120.<br />

42


Golding, M.C., Long, C.R., Carmell, M.A., Hannon, G.J., Westhusin, M.E. 2006. Suppression<br />

of prion protein in livestock by RNA interference. Proc. Natl. Acad. Sci. USA 103, 5285-<br />

5290.<br />

Gong, W., Wan, H., Tay, T.L., Wang, H., Chen, M., Yan, T. 2003 Development of transgenic<br />

fish for ornamental and bioreactor by strong expression of fluorescent proteins in the<br />

skeletal muscle. Biochemical and Biophysical Research Communications 308, 58-63.<br />

Golovan, S. P., Meidinger, R. G., Ajakaiye, A., Cottrill, M., Wiederkehr, M. Z., Barney, D. J.,<br />

Plante, C., Pollard, J. W., Fan, M. Z., Hayes, M. A., Laursen, J., Hjorth, J. P., Hacker, R.<br />

R., Phillips, J. P., Forsberg, C. W. 2001. Pigs expressing salivary phytase produce lowphosphorus<br />

manure. Nat. Biotechnol. 19, 741–745.<br />

Goodstadt, L., Ponting, C.P.. Phylogenetic reconstruction of orthology, paralogy, and<br />

conserved synteny for dog and human. PLoS Computational Biology 2006; 2:e133.<br />

Grosse-Hovest, L., Muller, S., Minoia, R., Wolf, E., Zakhartchenko, V., Wenigerkind, H.,<br />

Lassnig, C., Besenfelder, U., Müller, M., Lytton, S. D., Jung, G., and Brem, G. (2004).<br />

Cloned transgenic farm animals produce a bispecific antibody for T cell-mediated tumor<br />

cell killing. Proc. Natl. Acad. Sci. USA 101, 6858-6863.<br />

Gossen, M., Bujard, H. 1992. Tight control of gene expression in mammalian cells by<br />

tetracycline-responsive promoters. Proc Natl Acad Sci U S A. 89, 5547-51<br />

Gossen, M., Freundlieb, S., Bender, G., Müller, G., Hillen, W., Bujard, H..<br />

1995.Transcriptional activation by tetracyclines in mammalian cells. Science 268, 1766-9.<br />

Grobet, L., Martin, L.J., Poncelet, D., Pirottin, D., Brouwers, B., Riquet, J., et al. 1997. A<br />

deletion in the bovine myostatin gene causes the double-muscled phenotype in cattle.<br />

Nature Genetics 17, 71-4.<br />

Gu, H., Marth, J.D., Orban, P.C., Mossmann, R., Rajewsky, K. 1994. Deletion of a DNA<br />

polymerase ß gene segment in T cells using cell type-specific gene targeting. Science 265,<br />

103-106.<br />

Habermann, F.A. Wuensch, A., Sinowatz, F. Wolf, E. 2007. Reporter genes for<br />

embryogenesis research in livestock species. Theriogenology 68, S116-S124.<br />

Hammer, R. E., Pursel, V. G., Rexroad, C. E. Jr., Wall, R. J., Bolt, D. J., Ebert, K. M.,<br />

Palmiter, R. D., Brinster, R. L. 1985. Production of transgenic rabbits, sheep and pigs by<br />

microinjection. Nature 315, 680–683.<br />

Han, X., Amar, S., 2002. Identification of genes differentially expressed in cultured human<br />

periodontal ligament fibrobasts vs. human gingival fibroblast by DNA microarray<br />

analysis. J. Dent. Res. 81, 399-405.<br />

Haskell, R. E., and Bowen, R. A. 1995. Efficient production of transgenic cattle by retroviral<br />

infection of early embryos. Mol. Reprod. Dev. 40, 386-390.<br />

Hayflick, L. 1965. The limited in vitro proliferation of human diploid cells. Exp. Cell Res 37;<br />

614-636.<br />

43


Heby, O. 1995. DNA methylation and polyamines in embryonic development and cancer. Int<br />

J Dev Biol. 39, 737-57.<br />

Hochedlinger, K., Jaenisch, R. 2002. Monoclonal mice generated by nuclear transfer from<br />

mature B and T donor cells. Nature 415, 1035-1038.<br />

Hölker, M., Petersen, B., Hassel, P., Kues, W.A., Lemme, E., Lucas-Hahn, A., Niemann, H.<br />

2005. Duration of in vitro maturation of recipient oocytes affects blastocyst development of<br />

cloned porcine embryos. Cloning and Stem Cells 7, 34-43.<br />

Hofmann, A., Kessler, B., Ewerling, S., Weppert, M., Vogg, B., Ludwig, H., Stojkovic, M.,<br />

Boelhauve, M., Brem, G., Wolf, E., and Pfeifer, A. 2003. Efficient transgenesis in farm<br />

animals by lentiviral vectors. EMBO Rep. 4, 1054-1060.<br />

Hofmann, A., Zakhartchenko, V., Weppert, M.,, Sebald, H., Wenigerkind, H., Brem, G.,<br />

Wolf, E., and Pfeifer, A. 2004. Generation of transgenic cattle by lentiviral gene transfer<br />

into oocytes. Biol. Reprod. 71, 405-409.<br />

Hogan, B., Constantini, F., Lacy, E. 1986. Manipulating the mouse embryo. A laboratory<br />

manual. Cold Spring Habor Laboratory.<br />

Holley, R.W., Kiernan, J.A., 1968. “Contact inhibition” of cell division in 3T3 cells. Proc.<br />

Natl. Acad. Sci. USA 60, 300-304.<br />

Hornen, N., Kues, W.A., Carnwath, J.W., Lucas-Huhn, A., Petersen, B., Hassel, P., Niemann,<br />

H. 2007. Production of viable pigs from fetal somatic stem cells. Cloning and Stem Cells 9,<br />

364-73.<br />

Iqbal, K. Kues, W.A., and Niemann, H. 2007. Gene knock down of maternal and embryonic<br />

expressed GFP in murine embryos by the injection of a short interfering RNA (siRNA).<br />

Biochem Biophys Res Commun 358, 727-732.<br />

Jackson, D.A., Symons, R.H., Berg, P. 1972. Biochemical method for inserting new genetic<br />

information into DNA of Simian Virus 40: circular SV40 DNA molecules containing<br />

lambda phage genes and the galactose operon of Escherichia coli. Proc Natl Acad Sci U S<br />

A. 69, 2904-9.<br />

Jaenisch, R., Mintz, B. 1974. Simian virus 40 DNA sequences in healty mice derived from<br />

preimplantation blastocysts injected with viral DNA. Proc. Natl. Acad. Sci. USA 71, 1250-<br />

1254.<br />

Jaenisch, R. 1976. Germ line integration and Mendelian transmission of the exogenous<br />

Moloney leukemia virus. Proc Natl Acad Sci U S A. 73, 1260-4.<br />

Johnson, I.S.. Human insulin from recombinant DNA technology. Science 219, 632-637.<br />

Jorgensen, R.A., Atkinson, R.G., Forster, R.L., Lucas, W.J. 1998. An RNA-based information<br />

superhighway in plants. Science 279, 1486-1487.<br />

44


Kambadur, R., Sharma, M., Smith, T.P., Bass, J.J. 1997. Mutations in myostatin (GDF8) in<br />

double-muscled Belgian Blue and Piedmontese cattle. Genome Research 7, 910-6.<br />

Karlson, P., Doenecke, D., Koolman, J. 2005. Kurzes Lehrbuch der Biochemie für Mediziner<br />

und Naturwissenschaftler. Georg Thieme Verlag Stuttgart, New York.<br />

Kitagawa, M., Higashi, H., Suzuki, I.T., et al., 1994. A cylin-dependent kinase inhibitor,<br />

butyrolactone I, inhibits phosphorylation of RB protein and cell cycle progression.<br />

Oncogene 9, 2549-2557.<br />

Koipally, J., Renold, A., Kim, J., Georgopoulos, K. 1999. Repression by Ikaros and Aiolos is<br />

mediated through histone deacetylase complexes. EMBO J. 18, 3090-100.<br />

Kolber-Simonds, D., Lai, L., Watt, S.R. et al. 2004. Production of alpha-1,3-<br />

galactosyltransferase null pigs by means of nuclear transfer with fibroblasts bearing loss<br />

of heterozygosity mutations. Proc. Natl. Acad. Sci. USA 101, 7335-7340.<br />

Kues, W.A., Wunder, F. 1992. Heterogeneous expression patterns of mammalian potassium<br />

channel genes in developing and adult rat brain. Eur. J. Neurosci. 4, 1296-1308.<br />

Kues, W.A., Sakmann, B., Witzemann, V. 1995a. Differential expression patterns of five<br />

AChR subunit genes in rat muscle during development. Eur. J. Neurosci. 7, 1376-1385.<br />

Kues, W.A., Brenner, H., Sakmann, B., Witzemann, V. 1995b. Local neurotrophic repression<br />

of gene transcripts encoding fetal AChRs at rat neuromuscular synapses. J. Cell Biol. 130,<br />

949-957.<br />

Kues, W.A., Anger, M., Carnwath, J.W., Paul, D., Motlik, J., Niemann, H. 2000. Cell cycle<br />

synchronisation of porcine fetal fibroblasts. Effects of by serum deprivation and reversible<br />

cell cycle inhibitors. Biol. Reprod. 62, 412-419.<br />

Kues, W.A., Carnwath, J. W., Paul, D. and Niemann, H. 2002. Cell cycle synchronization of<br />

porcine fetal fibroblasts by serum deprivation initiates a nonconventional form of<br />

apoptosis. Cloning Stem Cells 4: 231-244.<br />

Kues, W.A. and Niemann, H. 2004. The contribution of farm animals to human health.<br />

Trends in Biotechnology 22, 286-294.<br />

Kues, W. A., Petersen, B., Mysegades, W., Carnwath, J.W. and Niemann, H. 2005a. Isolation<br />

of fetal stem cells from murine and porcine somatic tissue. Biology of Reproduction 72,<br />

1020-1028.<br />

Kues, W.A., Carnwath, J.W., Niemann, H. 2005b. From fibroblasts and stem cells:<br />

Implications for cell therapies and somatic cloning. Reproduction, Fertility and<br />

Development 17, 125-134.<br />

Kues, W. A. Schwinzer, R., Wirth, D., Verhoeyen, E., Lemme, E., Herrmann, D., Barg-Kues,<br />

B., Hauser, H., Wonigeit, H., and Niemann, H. 2006a. Epigenetic silencing and tissue<br />

independent expression of a novel tetracycline inducible system in double-transgenic pigs<br />

FASEB J 20, E1-E10, doi: 10.1096/fj.05-5415fje.<br />

45


Kues, W.A., Hofmann, D.R., Carnwath, J.W., de Souza, F.P., Madeira, H.M.F. and Niemann,<br />

H. 2006b. High incidence of single nucleotide polymorphisms in prion protein gene of<br />

native Brazilian Caracu cattle. J. Anim. Breeding and Genetics 123, 326-330.<br />

Kues, W.A., Rath, D., Niemann, H. 2008. Reproductive biotechnology goes genomics. CAB<br />

Reviews: Perspectives in Agriculture, Veterinary Science, Nutrition and Natural Resources<br />

3, No. 036<br />

Kuroiwa, Y., Kasinathan, P., Choi, Y. J., Naeem, R., Tomizuka, K., Sullivan, E. J., Knott, J.<br />

G., Duteau, A., Goldsby, R. A., Osborne, B. A., Ishida, I., and Robl, J. M. 2002. Cloned<br />

transchromosomic calves producing human immunoglobulin. Nat. Biotechnol. 20, 889–<br />

894.<br />

Kuroiwa Y, Kasinathan P, Matsushita H, Sathiyaselan J, Sullivan EJ, Kakitani M, Tomizuka<br />

K, Ishida I, Robl JM. 2004. Sequential targeting of the genes encoding immunoglobulinmu<br />

and prion protein in cattle. Nat Genet. 36, 775-80.<br />

Kuwaki, K., Tseng, Y. L., Dor, F. J., Shimizu, A., House, S. L., Sanderson, T. M., Lanceros,<br />

C. J., Rabharasuth, D. D., Cheng, J., Moran, K., Hisashi, Y., Mueller, N., Yamadoa, K.,<br />

Greenstein, J. L., Hawley, R. J., Patience,C., Awwad, M., Fishman, J. A., Robson, S. C.,<br />

Schuurman, H. J, Sachs, D. H, and Cooper, D. K. 2005. Heart transplantation in baboons<br />

usion a1,3-glactosyltransferase knock out pigs as donors: initial experiments. Nat. Med.<br />

11, 29-31.<br />

Lai, L., Kolber-Simonds, D., Park, K. W., Cheong, H. T., Greenstein, J. L., Im, G. S., Samuel,<br />

M., Bonk, A., Rieke, A., Day, B. N., Murphy, C. N., Carter, D. B., Hawley, R. J., and<br />

Prather, R. S. 2002. Production of alpha-1,3-galactosyltransferase knockout pigs by<br />

nuclear transfer cloning. Science 295, 1089–1092.<br />

Lai, L., Kang, J.X., Li, R., Wang, J., Witt, W.T., Yong, H.Y., Hao, Y., Wax, D.M., Murphy,<br />

C.N., Rieke, A., Samuel, M., Linville, M.L., Korte, S.W., Evans, R.W., Starzl, T.E.,<br />

Prather, R.S., and Dai, Y. 2006. Generation of cloned transgenic pigs rich in omega-3<br />

fatty acids. Nature Biotechnol. 24, 435-436.<br />

Lander, E.S. et al. 2001. Initial sequence and analysis of the human genome. Nature 409, 860-<br />

921.<br />

Lee, F., Mulligan, R., Berg, P., Ringold, G. 1991. Glucocorticoids regulate expression of<br />

dihydrofolate reductase cDNA in mouse mammary tumour virus chimaeric plasmids.<br />

Nature 294, 228-232.<br />

Li, E. 2002. Chromatin modification and epigenetic reprogramming in mammalian<br />

development. Nat Rev Genet. 3, 662-73.<br />

Lindblad-Toh, K, Wade, CM, Mikkelsen, TS, Karlsson, EK, Jaffe, DB, Kamal, M. et al. 2005.<br />

Genome sequence, comparative analysis and haplotype of the domestic dog. Nature 438,<br />

803-819.<br />

Lo, D., Pursel, V., Linton, P. J., Sandgren, E., Behringer, R., Rexroad, C., Palmiter, R. D., and<br />

Brinster, R. L. 1991. Expression of mouse IgA by transgenic mice, pigs and sheep. Eur. J.<br />

Immunol. 21, 1001-1006<br />

46


Maatouk, D.M., Kellam, L.D., Mann, M.R., Lei, H., Li, E., Bartolomei, M.S., Resnick, J.L.<br />

2006. DNA methylation is a primary mechanism for silencing postmigratory primordial<br />

germ cell genes in both germ cell and somatic cell lineages. Development 133, 3411-8.<br />

Mansour, S.L., Thomas, K.R., Capecchi, M.R. 1988. Disruption of the proto-oncogene int-2<br />

in mouse embryo-derived stem cells: a general strategy for targeting mutations to nonselectable<br />

genes. Nature. 336, 348-52.<br />

Manzini, S., Vargiolu, A., Stehle, I.M., Bacci, M.L., Cerrito, M.G., Giovannoni, R., Zannoni,<br />

A., Bianco, M. R., Forni, M., Donini, P., Papa, M., Lipps, H., and Lavitrano, M. (2006).<br />

Genetically modified pigs produced with a nonviral episomal vector. Proc. Natl. Acad.<br />

Sci. (USA) 103, 17672-17677.<br />

May, T., Eccleston, L., Herrmann, S., Hauser, H., Goncalves, J., Wirth, D. 2008. Bimodal and<br />

hysteretic expression in mammalian cells from a synthetic gene circuit. PloS One 3,<br />

e2372.<br />

Mayer, W., Niveleau, A., Walter, J., Fundele, R., Haaf, T. 2000a. Demethylation of the<br />

zygotic paternal genome. Nature. 403, 501-2.<br />

Mayer, W., Smith, A., Fundele, R., Haaf, T. 2000b. Spatial separation of parental genomes in<br />

preimplantation mouse embryos. J Cell Biol. 148, 629-34.<br />

Mayo, K.E., Warren, R., Palmiter, R.D. 1982. The mouse metallothionein-I gene in<br />

transcriptionally regulated by cadmiun following transfection into human or mouse cells.<br />

Cell 29, 99-108.<br />

Maxam, A.M., Gilbert, W. 1977. A new method for sequencing DNA. Proc Natl Acad Sci U<br />

S A. 74, 560-4<br />

McCurry, KR, Kooyman, Kl, Diamond, LE et al., 1995. Transgenic expression of human<br />

complement regulatory proteins in mice results in diminished complement deposition<br />

during organ xenoperfusion. Transplantation 59, 1177-1183.<br />

McGregor, C.G., Teotia, S.S., Byrne, G.W. et al. 2004. Cardiac xenotransplantation: progress<br />

toward the clinic. Transplantation 78, 1569-1575.<br />

McPherron, A.C., Lee, S.J. 1997. Double muscling in cattle due to mutations in the myostatin<br />

gene. Proc Natl. Acad. Sci. USA 94, 12457-61.<br />

Mellitzer, G., Hallonet, M., Chen, L., Ang, S.L. 2002. Spatial and temporal knock downs of<br />

gene expression by electroporation of dsRNA and morpholinos into early periimplantation<br />

embryos. Mech. Dev. 118, 57-63.<br />

Miller, K.F., Bolt, D.J., Pursel, V.G., Hammer, R.E., Pinkert, C., Palmiter, R.D. 1989.<br />

Expression of human or bovine growth hormone gene with a mouse metallothionein-1<br />

promoter in transgenic swine alters the secretion of porcine growth hormone and insulilike<br />

growth factor-I. J. Endocrinol. 120, 481-488.<br />

47


Moulin, V., Garrel, D., Auger, F.A., et al., 1999. What´s new in human wound-healing<br />

myofibroblasts? Curr. Top. Pathol. 93, 123-133.<br />

Müller, M., Brenig, B., Winnacker, E. L., and Brem, G. 1992. Transgenic pigs carrying<br />

cDNA copies encoding the murine Mx1 protein which confers resistance to influenza<br />

virus infection. Gene 121, 263-270.<br />

Müller, M, Brem, G. 1994 Transgenic strategies to increase disease resistance in livestock.<br />

Reprod. Fert. Dev 6, 605-613.<br />

Müller, U.T. 1999. Ten years of gene targeting: targeted mouse mutants, from vector design<br />

to phenotype analysis. Mech Dev. 82, 3-21.<br />

Mulligan, R.C., Howard, B.H., Berg, P. 1979. Synthesis of rabbit beta-globin in cultured<br />

monkey kidney cells following infection with a SV40 beta-globin recombinant genome.<br />

Nature 277, 108-14.<br />

Nady, N., Min, J., Kareta, M.S., Chédin, F., Arrowsmith, C.H. 2008. A SPOT on the<br />

chromatin landscape? Histone peptide arrays as a tool for epigenetic research. Trends<br />

Biochem Sci. 2008 Jun 4.<br />

Nganvongpanit, K., Müller, H., Rings, B., Gilles, M., Jennen, D., Hoelker, M., Tholen E.,<br />

Schellander, D., Tesfaye, D. 2005. Mol. Reprod. Dev. 73, 153-163.<br />

Niemann, H., Kues, W.A. 2000. Transgenic livestock: premises and promises. Anim. Reprod.<br />

Sci. 60-61, 277-293.<br />

Niemann, H., Kues, W.A. 2003. Application of transgenesis on livestock for agriculture and<br />

biomedicine. Anim. Reprod. Sci. 79, 291–317.<br />

Niemann, H., Kues, W.A., Carnwath, J. W. 2005. Transgenic farm animals: present and past.<br />

Rev. Sci. Tech. OIE 24, 285-298.<br />

Niemann, H, Kues, W.A. 2007. Transgenic farm animals – an update. Reproduction, Fertility<br />

and Development 19, 762-770.<br />

Niemann, H., Carnwath, J.W., Kues, W.A. 2007. Application of array technology to<br />

mammalian embryos. Theriogenology 68S, S165-S177.<br />

Nightingale, P, Martin, P. 2004. The myth of the biotech revolution. Trends Biotechnol 22,<br />

564-569.<br />

Nottle, M. B., Nagashima, H., Verma, P. J., Du, Z. T., Grupen, C. G., McIlfatick, S. M.,<br />

Ashman, R. J., Harding, M. P., Giannakis, C., Wigley, P. L., Lyons, I. G., Harrison, D. T.,<br />

Luxford, B. G., Campbell, R. G., Crawford, R. J., Robins, A. J. 1999. Production and<br />

analysis of transgenic pigs containing a metallothionein porcine growth hormon gene<br />

construct. In ‘Transgenic Animals in Agriculture’. (Eds J. D. Murray, G. B. Anderson, A.<br />

M. Oberbauer, and M. M. McGloughlin) pp. 145-156. (CABI Publishing: New York, NY,<br />

USA.)<br />

48


Orban, PC, Chui, D, Marth, JD 1992. Tissue- and site-specific DNA recombination in<br />

transgenic mice. Proc. Natl. Acad. Sci. USA 89, 6861-6865.<br />

Palmiter, RD, Brinster, RL, Hammer, RE, Trumbauer, ME, Rosenfeld, MG, Birmberg, NC,<br />

Evans, RM, 1982. Dramatic growth of mice that develop from eggs microinjected wit h<br />

metallothionein-growth hormone fusion genes. Nature 300, 611-615.<br />

Panarace, M., Agüero, J.I., Garrote, M. et al., 2006. How healthy are clones and their<br />

progeny: 5 years of field experience. Theriogenology 67, 142-151.<br />

Paradis, F., Vigneault, C., Robert, M.A., Sirard, M. 2005. RNA interference as a tool to study<br />

gene fuction in bovine oocytes. Mol. Reprod. Dev. 70, 111-121.<br />

Pedrail-Noy, G., Spadari, S., Miller-Faures, A., Miller, A.O.A., Kruppa, J., Koch, G. 19980.<br />

Synchronsation of Hela by inhibition of DNA polymerase A with aphidicolin. Nucl. Acid<br />

Res. 8, 377-387.<br />

Pfeifer, A., Ikawa, M., Dayn, Y., Verma, I.M. 2002. Transgenesis by lentiviral vectors: lack<br />

of gene silencing in mammalian embryonic stem cells and preimplantation embryos. Proc<br />

Natl Acad Sci U S A. 99, 2140-5<br />

Phelps, C. J., Koike, C., Vaught, T. D., Boone, J., Wells, K. D., Chen, S. H., Ball, S., Specht,<br />

S. M., Polejaeva, I. A., Monahan, J. A., Jobst, P. M., Sharma, S. B., Lamborn, A. E.,<br />

Garst, A. S., Moore, M., Demetris, A. J., Rudert, W. A., Bottino, R., Bertera, S., Trucco,<br />

M., Starzl, T. E., Dai, Y., and Ayares, D. L. 2003. Production of alpha 1,3galactosyltransferase<br />

deficient pigs. Science 299, 411–414.<br />

Platenburg, G. J., Kootwijk, E. P., Kooiman, P. M., Woloshuk, S. L., Nuijens, J. H.,<br />

Krimpenfort, P. J., Pieper, F. R., de Boer, H. A., and Strijker, R. (1994). Expression of<br />

human lactoferrin in milk of transgenic mice. Transgenic Res. 3, 99-108.<br />

Plusa, B., Frankenberg, S., Chalmers, A., Hadjantonakis, C.A., Moore, N., Papalopulu, V.E.,<br />

Papaioannou, D.M., Glover, M., Zernicka-Goetz, M 2004. Downregulation of Par3 and<br />

aPKC function directs cells towards the ICM in the preimplantation mouse embryo. J.<br />

Cell Sci. 118, 505-515.<br />

Pursel, V.G., Pinkert C.A., Miller ,K.F., Boldt, D.J., Campbell, R.G., Palmiter ,R.D., Brinster,<br />

R.L., Hammer, R.E. 1989. Genetic engineering of livestock. Science 244, 1281-1288.<br />

Pursel, V. 1998. Modification of production traits. In AnimalBreeding — Technology for the<br />

21st Century, ed. Clark, A. J.,pp. 183—200, Overseas Publishers Association,<br />

Amsterdam.<br />

Pursel, V. G., Wall, R. J., Mitchell, A. D., Elsasser, T. H., Solomon, M. B., Coleman, M. E.,<br />

Mayo, F., and Schwartz, R. J. 1999. Expression of insulin-like growth factor-1 in skeletal<br />

muscle of transgenic pigs. In ‘Transgenic animals in agriculture’. (Eds J. D. Murray, G. B.<br />

Anderson, A. M. Oberbauer and M. M. McGloughlin.) pp. 131-144. (CABI Publishing:<br />

New York, NY, USA.)<br />

Reed, S.I., 1997. Control of the G1/S transition. Cancer Surv 29, 7-23.<br />

49


Reik, W. 2007. Stability and flexibility of epigenetic gene regulation in mammalian<br />

development. Nature. 447: 425-32<br />

Reh, W.A., Maga, E. A., Collette, N. M., Moyer, A., Conrad-Brink, J. S., Taylor, S. J.,<br />

DePeters, E. J., Oppenheim, S., Rowe, J. D., BonDurant, R. H., Anderson, G. B., and<br />

Murray, J. D. 2004. Hot topic: using a stearoyl-CoA desaturase transgene to alter milk<br />

fatty acid composition. J. Dairy Sci. 87, 3510-3514.<br />

Richt, J.A., Kasinathan, P., Hamir, A.N., Castilla, J., Sathiyaseelan, T. et al., 2007. Production<br />

of cattle lacking prion protein. Nat Biotechnol. 25, 132-8<br />

Robertson, A., Perea, J., Tomachova, T., Thomas, P.K., Huxley, C. 2002. Effects of mouse<br />

strain, position of integration and tetracycline analogue on the tetracycline conditonal<br />

system in transgenic mice. Gene 282, 65-72.<br />

Robl, J. M., Wang, Z., Kasinathan, P., and Kuroiwa, Y. 2007. Transgenic animal production<br />

and animal biotechnology. Theriogenology 67, 127-133<br />

Rudolph, N. S. 1999. Biopharmaceutical production in transgenic livestock. Trends<br />

Biotechnol. 17, 367–374.<br />

Saeki, K., Matsumoto, K., Kinoshita, M., Suzuki, I., Tasaka, Y., Kano, K., Taguchi, Y.,<br />

Mikami, K., Hirabayashi, M., Fashiwazaki, N., Hosoi, Y., Murata, N., and Iritani, A.<br />

2004. Functional expression of a Delta12 fatty acid desaturase gene from spinach in<br />

transgenic pigs. Proc. Natl. Acad. Sci. USA 101, 6361-6366.<br />

Sanger, F., Donelson, J.E., Coulson, A.R., Kössel, H., Fischer, D. 1973.Use of DNA<br />

polymerase I primed by a synthetic oligonucleotide to determine a nucleotide sequence in<br />

phage fl DNA. Proc Natl Acad Sci U S A. 70, 1209-13<br />

Sambrook, J, Frisch, EF, Maniatis, T. Molecular cloning. A laboratory manual. 2 nd Ed. Cold<br />

Spring Habor Laboratory 1989.<br />

Sanger, F., Nicklen, S., Coulson, A.R. 1977. DNA sequencing with chain-terminating<br />

inhibitors. Proc Natl Acad Sci U S A. 74, 5463-7.<br />

Sauer, B., Henderson, N. 1988. Site-specific DNA recombination in mammalian cells by the<br />

Cre recombinase of bacteriophage P1. Proc Natl Acad Sci U S A. 85, 5166-70.<br />

Sauer, B.1998. Inducible gene targeting in mice using the Cre/lox system. Methods 14, 381-<br />

92.<br />

Schätzlein, S., Lucas-Hahn, A., Lemme, E., Kues, W.A., Dorsch, M., et al., 2004. Telomere<br />

length is reset during early embryogenesis. Proc. Natl.l Acad. Sci. USA 101, 8034-8038.<br />

Scharschmidt, T.C., Segre, J.A.. 2008. Modeling atopic dermatitis with increasingly complex<br />

mouse models. J Invest Dermatol. 128, 1061-4.<br />

Schnieke, A. E., Kind, A. J., Ritchie, W. A., Mycock, K., Scott, A. R., Ritchie, M., Wilmut, I.,<br />

Colman, A., and Campbell, K. H. 1997. Human factor IX transgenic sheep produced by<br />

transfer of nuclei from transfected fetal fibroblasts. Science 278, 2130-2133.<br />

50


Schuelke, M., Wagner, K.R., Stolz, L.E., Hübner, C., Riebel, et al. 2004. Myostatin mutation<br />

associated with gross muscle hypertrophy in a child. New England Journal of Medicine<br />

350: 2682-8.<br />

Schultze, N., Burki, Y., Lang, Y., Certa, U., Bluethmann, H. 1996. Efficient control of gene<br />

expression by single step integration of the tetracycline system in transgenic mice. Nat<br />

Biotechnol. 14, 499-503.<br />

Schultz, R.M., 2002. The molecular foundations of the maternal to zygotic transition in the<br />

preimplantion embryo. Hum. Reprod. Update 8, 323-331.<br />

Senawong, T., Peterson, V.J., Avram, D., Shepherd, D.M., Frye, R.A., Minucci, S., Leid, M..<br />

2003. Involvement of the histone deacetylase SIRT1 in chicken ovalbumin upstream<br />

promoter transcription factor (COUP-TF)-interacting protein 2-mediated transcriptional<br />

repression. J Biol Chem. 278, 43041-50<br />

Sharp, PA, 2001. RNA interference-2001. Genes Development 15, 485-490.<br />

Soares, M.L., Haraguchi, S., Torres-Padilla et al., 2005. Functional studies of signalling<br />

pathways in peri-implantation development of the mouse embryo by RNAi. BMC Dev.<br />

Biol. 5, 1-11.<br />

Stepensky, D., Kleinberg, L., Hoffman, A., 2003. Bone as an effect compartment . Clin.<br />

Pharmacokinet, 42, 863-881.<br />

St-Onge, L., Furth, P.A., Gruss, P.1996. Temporal control of the Cre recombinase in<br />

transgenic mice by a tetracycline responsive promoter. Nucleic Acids Res. 24, 3875-7.<br />

Svoboda, P., Stein, P., Hayashi, H., Schultz, R.M. 2000. Selective reduction of dormant<br />

maternal mRNA in mouse oocytes by RNA interference. Development 127, 4147-4156.<br />

Szabo, P.E., Huebner, K., Scholer, H., Mann, J.R. 2002. Allele-specific expression of<br />

imprinted genes in mouse migratory primordial germ cell. Mech. Dev. 115, 157-160.<br />

Taft RA. 2008. Virtues and limitations of the preimplantation mouse embryo as a model<br />

system. Theriogenology 69, 10-16.<br />

Tenenhaus Dann, C., Alvarado, A. L., Hammer, R. E., and Garbers, D. L. 2006. Heritable and<br />

stable gene knockdown in rats. Proc. Natl., Acad. Sci. USA 103, 11246-11251.<br />

Van den Hout, J. M., Reuser, A. J., de Klerk, J. B., Arts, W. F., Smeitink, J. A., and Van der<br />

Ploeg, A. T. (2001). Enzyme therapy for Pompe disease with recombinant human aglucosidase<br />

from rabbit milk. J. Inherit. Metab. Dis. 24, 266–274.<br />

Van Reenen, C.G., Meuwissen, T.H.E., Hopster, H, Oldenbroek, K., Kruip, T.H., Blokhuis,<br />

H.J. 2001. Transgenesis may affect farm animal welfare: a case for systemic risk<br />

assessment. J. Anim Sci 79, 1763-1769.<br />

Wall, RJ. 1996. Transgenic livestock: progress and prospects for the future. Theriogenology<br />

45, 57-68.<br />

51


Wall, R.J., Powell, A., Paape, M.J., Kerr, D.E., Bannermann, D.D., Pursel, V.G., Wells, K.D.,<br />

Talbot, N., and Hawk, H. (2005). Genetically enhanced cows resist intramammary<br />

Staphylococcus aureus infection. Nat. Biotechnology 23, 445-451.<br />

Wall RJ, Shani M. 2008. Are animal models as good as we think? Theriogenology 69, 2-9.<br />

Wang, X., Willenbring , H., Akkari, Y., et al., 2003. Cell fusion is the principal source of<br />

bone-marrow-derived hepatocytes. Nature 422, 897-901.<br />

Watson, J.D., Tooze, J, Kurtz, DT 1983. Recombinant DNA. A short course. Scientific<br />

American Books 1983.<br />

Wells, D.N., Misica, P.M., Day, T.A., Tervit, H.R. 1997. Production of cloned lambs from an<br />

established embryonic cell line: a comparision between in vivo and in vitro-matured<br />

cytoplasts Biol. Reprod. 57, 385-393.<br />

Wianny, F., Zernicka-Goetz, M. 2000. Specific interference with gene function by doublestranded<br />

RNA in early mouse development. Nat. Cell Biol 2, 70-75.<br />

Wilmut, I. Schnieke, A. E., McWhir, J., Kind, A. J., and Campbell, K. H. 1997. Viable<br />

offspring derived from fetal and adult mammalian cells. Nature 385, 810-813.<br />

Wheeler, M. B., Bleck, G. T., and Donovan, S. M. 2001. Transgenic alteration of sow milk to<br />

improve piglet growth and health. Reproduction 58 (Suppl.), 313-324.<br />

Whitelaw, C. B., Radcliffe, P. A., Ritchie, W. A., Carlisle, A., Ellard, F. M., Pena, R. N.,<br />

Rowe, J., Clark, A. J., King, T. J., and Mitrophanous, K.A. 2004. Efficient generation of<br />

transgenic pigs using equine infectious anaemia virus (EIAV) derived vector. FEBS Lett.<br />

571, 233-236.<br />

Wolf, E, Schernthaner, W, Zakhartchenko, V, Prelle, K, Stojkovic, M, BremG 2000.<br />

Transgenic technology in farm animals-progress and perspectives. Exp Physiol 85, 615-<br />

625.<br />

Yadav, P., Kues, W.A., Herrmann, D., Carnwath, J.W., Niemann, H. 2005. Bovine ICMs<br />

express the Oct4 ortholog. Molecular Reproduction and Development 72, 182-190.<br />

Yang XW, Gong S. An overview on the generation of BAC transgenic mice for neuroscience<br />

research.Curr Protoc Neurosci. 2005 May<br />

Yamada, K., Yazawa, K., Shimizu, A., Iwanaga, T., Hisashi, Y., Nuhn, M., O’Malley, P.,<br />

Nobori, S., Vagefi, P. A., Patience, C., Fishman, J., Cooper, D. K., Hawley, R. J.,<br />

Greenstein, J., Schuurman, H. J., Awwad, M., Sykes, M., and Sachs, D. H. 2005. Marked<br />

prolongation of porcine renal xenograft survival in baboons through the use of α1,3galactosyltransferase<br />

gene-knockout donors and the cotransplantation of vascularized<br />

thymic tissue. Nature Med. 11, 32-34.<br />

Yao, Y.L., Yang, W.M., Seto, E. 2001. Regulation of transcription factor YY1 by acetylation<br />

and deacetylation. Mol Cell Biol. 21, 5979-91.<br />

52


Yarranton, GT. 1992. Inducible vectors for expression in mammalian cells. Curr Opin<br />

Biotechnol. 3, 506-11.<br />

Yeom, Y.I., Fuhrmann, C.E., Ovitt, C.E., Brehm, A., Ohbo, K., Grosse, M., Huebner ,K.,<br />

Scholer H.R., 1996. Germline regularory element of Oct4 specific for the totipotent cyle<br />

of embryonal cell. Development 122, 881-984.<br />

Ying, Q.L., Nichols, J., Evans, W.P., Smith, A.G. 2002. Changing potency by spontaneous<br />

fusion. Nature 416, 545-548.<br />

Yong-Guang, Y., Sykes, M. 2007. Xenotransplantation: current status and a perspective on<br />

the future. Nat. Reviews Immunology.<br />

Young LE, Sinclair KD, Wilmut I. 1998. Large offspring syndrome in cattle and sheep. Rev<br />

Reprod. 3, 155-63<br />

Zaidi, A., Schmoeckel, M, Bhatti, F et al., 1998. Life-supporting pig to primate renal xenotransplantation<br />

using genetically modified donors. Transplantation 65, 1584-1590.<br />

Zetterberg, A., Larsson, O., Wiman, K.G., 1995. What is the restriction point? Curr. Opin,<br />

Cell Biol. 7, 835-842.<br />

Zhong, R. 2007. Gal knockout and beyond. American J. Transplant. 7, 5-11.<br />

Zhou, D., He, Q.S., Wang, C., Zhang, J., Wong-Staal, F. 2006. RNA interference and<br />

potential applications. Curr Top Med Chem. 6, 901-11.<br />

Zuelke, R. 1998.Transgenic modification of cows milk for value-added processing. Reprod<br />

Fertil Dev. 10, 671-6.<br />

53


7.0 Danksagungen<br />

Prof. Dr. H. Niemann danke ich für die gewährte wissenschaftliche Freiheit.<br />

Dr. J. Carnwath, Prof. K. Wonigeit (MHH), Dr. D. Wirth (HZI, Bs), Dr. H. Hauser (HZI, Bs)<br />

und Dr. C. Wrenzycki (jetzt TiHo) danke ich für die intensiven Diskussionen. Dr. B.<br />

Petersen, Dr. A. Lucas-Hahn, Dr. M. Hölker (jetzt Universität Bonn) und Dr. N. Hornen (jetzt<br />

praktizierende Tierärztin) hatten entscheidenden Anteil an den Kerntransferversuchen.<br />

Frau D. Herrmann hat in ausgezeichneter Weise die zahlreichen Proben aufgear<strong>bei</strong>tet; die<br />

Zellkulturen wurden durch Frau W. Mysegades (jetzt stud. Veterinärmedizin), die Maus- und<br />

Embryonenkulturversuche durch Frau E. Lemme und Frau K. Korsawe und die<br />

Grosstierversuche durch Herrn K.-H. Hadeler technisch hervorragend unterstützt.<br />

Für die photo- und printtechnische Unterstützung wird dem Photographen gedankt.<br />

Für das <strong>zur</strong> Verfügungstellen von spezifischen Plasmiden und Zell-Linien wird Dr. M. Anger,<br />

(Universität Oxford), Dr. J. Bode (HZI, Braunschweig), Dr. M. Buchholz (MPI, Dresden), Dr.<br />

T. Cantz (MPI, Münster), Dr. K. Collins (Berkeley, Californien), Dr. E. Heinemann<br />

(Whitehead Institute, Cambridge), Dr. Lipps, (Universität Witten), Dr. U. Martin (MHH,<br />

Hannover), Dr. Ritter (Charite, Berlin), Dr. T. Saric (Universität Köln), Dr. V. Witzemann<br />

(MPI, Heidelberg) und Dr. J.K. Zhu (University of California) ganz herzlich gedankt.<br />

Für die finanzielle Förderung der DFG (Transregio 535) und des BMBF wird gedankt.<br />

54


8.0 Darstellung des eigenen Anteils an den wissenschaftlichen Publikationen<br />

Characterisierung primärer Zellen als Donorzellen für den Kerntransfer<br />

Anger, M., Kues, W.A., Klima, J., Mielenz, M., Motlik, J., Carnwath, J.W. and Niemann, H.<br />

2003 Cloning and cell cycle-specific regulation of Polo-like kinase in porcine fetal fibroblasts.<br />

Molecular Reproduction and Development 65, 245-253.<br />

a) Idee und Versuchsplanung Anger, M, Kues, W.A.<br />

b) Versuchsdurchführung Anger, M., Kues, W.A., Klima, J. Mielenz, M.<br />

c) Auswertung Anger, M., Kues, W.A.<br />

d) Verfassen des Manuskriptes Anger, M., Kues, W.A., Motlik, J., Niemann, H.<br />

Kues, W. A., Petersen, B., Mysegades, W., Carnwath, J.W. and Niemann, H. 2005a. Isolation<br />

of fetal stem cells from murine and porcine somatic tissue.<br />

Biology of Reproduction 72, 1020-1028<br />

a) Idee und Versuchsplanung Kues, W.A.<br />

b) Versuchsdurchführung Kues, W.A., Mysegardes, W.<br />

c) Auswertung Kues, W.A., Niemann, H.<br />

d) Verfassen des Manuskriptes Kues, W.A., Carnwath, J.W., Niemann, H.<br />

Yadav, P., Kues, W.A., Herrmann, D., Carnwath, J.W., Niemann, H. 2005. Bovine ICMs<br />

express the Oct4 ortholog.<br />

Molecular Reproduction and Development 72, 182-190.<br />

a) Idee und Versuchsplanung Kues, W.A., Niemann, H.<br />

b) Versuchsdurchführung Yadav, P.<br />

c) Auswertung Kues, W.A.<br />

d) Verfassen des Manuskriptes Yadav, P., Kues, W.A., Niemann, H.<br />

Dieckhoff B, Karlas A, Hofmann A, Kues WA, Petersen B, Pfeifer A, Niemann H, Kurth R,<br />

Denner J. 2007. Inhibition of porcine endogenous retroviruses (PERVs) in primary porcine<br />

cells by RNA interference using lentiviral vectors.<br />

Archives of Virology 152, 629-34.<br />

a) Idee und Versuchsplanung Denner, J<br />

b) Versuchsdurchführung Dieckhoff, B.<br />

c) Auswertung Dieckhoff, B., Kues, Denner, J.<br />

d) Verfassen des Mauskriptes Denner, J; Kues, W.A.; Denner, J., Niemann, H.<br />

55


DNA-Mikroinjektion und somatischer Kerntransfer<br />

Schätzlein S., Lucas-Hahn, A., Lemme, E., Kues, W.A., Dorsch, M., Manns, M., Niemann,<br />

H., Rudolph, K.H. 2004. Telomere length is reset during early embryogenesis.<br />

Proceedings of the National Academy of Sciences of the USA 101, 8034-8038.<br />

a) Idee und Versuchsplanung Niemann, H., Rudolph, K.H.<br />

b) Versuchsdurchführung Schätzlein, S., Lucas-Hahn, A., Lemme, E., Kues, W.A.<br />

c) Auswertung Rudolph, K.H., Schätzlein, S., Kues, W.A., Niemann, H.<br />

d) Verfassen des Manuskriptes Rudolph, K.H., Schätzlein, S., Kues, W.A., Niemann, H.<br />

Hölker, M., Petersen, B., Hassel, P., Kues, W.A., Lemme, E., Lucas-Hahn, A., Niemann, H.<br />

2005. Duration of in vitro maturation of recipient oocytes affects blastocyst development of<br />

cloned porcine embryos.<br />

Cloning and Stem Cells 7, 34-43.<br />

a) Idee und Versuchsplanung Niemann, H., Kues, W.A..<br />

b) Versuchsdurchführung Hölker, M., Petersen, B.<br />

c) Auswertung Hölker, M.., Kues, W.A., Niemann, H.<br />

d) Verfassen des Manuskriptes Hölker, M.; Kues, W.A., Niemann, H.<br />

Kues, W.A., Schwinzer, R., Wirth, D., Verhoeyen, E., Lemme, E., Herrmann, D., Barg-Kues,<br />

B., Hauser, H., Wonigeit, K., Niemann H. 2006 Epigenetic silencing and tissue independent<br />

expression of a novel tetracycline inducible system in double-transgenic pigs.<br />

FASEB J. 20, E1-E10, doi: 10.1096/fj.05-5415fje.<br />

a) Idee und Versuchsplanung Kues, W.A., Niemann, H., Hauser, H., Wonigeit, K.<br />

b) Versuchsdurchführung Kues, W.A., Schwinzer, R.<br />

c) Auswertung Kues, W.A., Schwinzer, R.<br />

d) Verfassen des Manuskriptes Kues, W.A., Niemann, H.<br />

Deppenmeier, S., Bock, O., Mengel, M., Niemann, H., Kues, W., Lemme, E., Wirth, D,<br />

Wonigeit, K., Kreipe, H. 2006. Health status of transgenic pig lines expressing human<br />

complement regulator protein CD59.<br />

Xenotransplantation 13, 345-356.<br />

a) Idee und Versuchsplanung Wonigeit, K., Niemann, H., Kreipe, H.<br />

b) Versuchsdurchführung Deppenmeier, S.<br />

c) Auswertung Deppenmeier, S., Kues, W.A., Bock, O., Wonigeit. K<br />

d) Verfassen des Manuskriptes Deppenmeier, S., Bock, O., Kues, W.A., Niemann, H.<br />

Hornen, N., Kues, W.A., Carnwath, J.W., Lucas-Huhn, A., Petersen, B., Hassel, B.,<br />

Niemann, H. 2007. Production of viable pigs from fetal somatic stem cells.<br />

Cloning and Stem Cells 9, 364-73.<br />

a) Idee und Versuchsplanung Kues, W.A., Niemann, H.<br />

b) Versuchsdurchführung Hornen, N.<br />

c) Auswertung Hornen, N., Kues, W.A.<br />

d) Verfassen des Manuskriptes Hornen, N., Niemann, H., Kues, W.A.<br />

56


Iqbal, K, Kues, W.A., Niemann, H. 2007. Parent of origin dependent gene-specific knock<br />

down in mouse embryos.<br />

Biochemical Biophysical Research Communications 358, 727-732.<br />

a) Idee und Versuchsplanung Kues, W.A.<br />

b) Versuchsdurchführung Iqbal, K.<br />

c) Auswertung Iqbal, K, Kues, W.A.<br />

d) Verfassen des Manuskriptes Kues, W.A., Niemann, H.<br />

Dieckhoff, B., Petersen, B., Kues, W.A., Kurth, R., Niemann, H., Denner, J. 2008.<br />

Knockdown of porcine endogenous retrovirus (PERV) expression by PERV-specific shRNA<br />

in transgenic pigs .<br />

Xenotransplantation 15, 36-45.<br />

a) Idee und Versuchsplanung Denner, J.<br />

b) Versuchsdurchführung Dieckhoff, B., Petersen, B., Kues, W.A.,<br />

c) Auswertung Dieckhoff, B.<br />

d) Verfassen des Manuskriptes Dieckhoff, B., Kues, W.A., Niemann, H., Denner, J.<br />

Übersichtsartikel: Transgene Nutztiere, Stammzellen und DNA-Array Analysen<br />

Niemann, H., and Kues, W.A. 2003. Application of transgenesis in livestock for agriculture<br />

and biomedicine.<br />

Animal Reproduction Science 79, 291-317.<br />

a) Idee Niemann, H., Kues, W.A.<br />

b) Umsetzung Niemann, H., Kues, W.A.<br />

c) Verfassen des Manuskriptes Niemann, H., Kues, W.A.<br />

Kues, W.A. and Niemann, H. 2004. The contribution of farm animals to human health.<br />

Trends in Biotechnology 22, 286-294.<br />

a) Idee Kues, W.A., Niemann, H.<br />

b) Umsetzung Kues, W.A., Niemann, H.<br />

c) Verfassen des Manuskriptes Kues, W.A., Niemann, H.<br />

Kues, W.A., Carnwath, J.W., Niemann, H. 2005b. From fibroblasts and stem cells:<br />

Implications for cell therapies and somatic cloning.<br />

Reproduction, Fertility and Development 17, 125-134.<br />

a) Idee Kues, W.A., Niemann, H.<br />

b) Umsetzung Kues, W.A., Niemann, H.<br />

c) Verfassen des Manuskriptes Kues, W.A., Carnwath, J.W., Niemann, H.<br />

57


Niemann H, Kues W, Carnwath JW. 2005. Transgenic farm animals: present and future.<br />

OIE Scientific and Technical Reviews (Rev. Sci. Tech. Off. Int. Epiz.) 24, 284-298.<br />

a) Idee Niemann, H.<br />

b) Umsetzung Kues, W.A., Niemann, H.<br />

c) Verfassen des Manuskriptes Niemann H, Kues W, Carnwath JW<br />

Niemann, H, Kues, W.A. 2007. Transgenic farm animals – an update.<br />

Reproduction, Fertility and Development 19, 762-770.<br />

a) Idee Kues, W.A., Niemann, H<br />

b) Umsetzung Kues, W.A., Niemann, H.<br />

c) Verfassen des Manuskriptes Kues, W.A., Niemann, H.<br />

Niemann, H., Carnwath, J.W., Kues, W.A. 2007. Application of array technology to<br />

mammalian embryos.<br />

Theriogenology 68S, S165-S177.<br />

a) Idee Niemann, H.<br />

b) Umsetzung Kues, W.A., Niemann, H.<br />

c) Verfassen des Manuskriptes Kues, W.A., Carnwath, J.W., Niemann, H.<br />

Kues, W.A., Rath, D., Niemann, H. 2008. Reproductive biotechnology goes genomics.<br />

CAB Reviews: Perspectives in Agriculture, Veterinary Science, Nutrition and Natural<br />

Resources 3, No. 036, 1-18.<br />

a) Idee Kues, W.A., Niemann, H.<br />

b) Umsetzung Kues, W.A., Niemann, H.<br />

c) Verfassen des Manuskriptes Kues, W.A., Rath, D., Niemann, H.<br />

58


9. Abkürzungsverzeichnis<br />

ACTB ß-Actin-Gen<br />

AdoMet S-Adenosylmethionin<br />

AVR akute vasculäre Abstoßung<br />

BAC bakterielles artifizielles Chromosom<br />

Bp Basenpaar<br />

BSE bovine spongiforme Enzephalitis<br />

CpG Cytosin-Guanin-Dinukleotid<br />

CMV Cytomegalovirus<br />

DNA-MT DNA-Methyltransferase<br />

EMEA European Medicine Agency<br />

ES embryonale Stammzelle<br />

Fo Founder, transgenes Tier, das <strong>zur</strong> Gründung einer Linie genutzt wird<br />

F1…n erste Generation einer transgenen Linie, …nte Generation<br />

FACS fluorescence activated cell sorting<br />

FDA Food and Drug Agency<br />

G0, G1 Zellzyklusphase G0 (Ruhephase), G1 (gap1 der Interphase)<br />

GDF8 Myostatin-Gen<br />

GFP green fluorescent protein<br />

GH Wachstumshormon<br />

GMP good manufacturing practices<br />

FSSC fetal somatic stem cell<br />

hCD59 humaner Komplementregulator CD59<br />

hDAF humaner Komplementregulator DAF, decay accelerating factor<br />

IRES interne Ribosomenbindungsstelle<br />

5mCpG 5-Methyl-CpG<br />

mRNA messenger RNA<br />

NTA bicistronische Kassette mit tTA als zweitem Cistron<br />

Oct4/OCT4 octamer binding transcription factor 4<br />

OG2 Oct4-GFP transgene Mauslinie 2<br />

PA Polyadenylierungssequenz<br />

pCMV<br />

Promoter von CMV<br />

PERV porcine endogene Retroviren<br />

PRNP Prionprotein-Gen<br />

ptTA tetracyclin-sensitiver Promoter, wird durch Bindung von tTA induziert<br />

RCA regulator of complement activation, Komplementregulator<br />

RISC RNA induced silencing complex<br />

RNAi RNA-Interferenz<br />

siRNA short interfering RNA<br />

shRNA short hairpin RNA<br />

SNP Einzelbasenpolymorphismus<br />

SPF spezifisch Pathogen-frei<br />

STE swine testis epithelial cell line<br />

SV40 Simian Virus 40<br />

tetO Tetracyclin-Operator<br />

tet-off konditionale Genexpression, tTA wird durch Tetracyclin inaktiviert<br />

tet-on konditionale Genexpression, mutierter tTA wird durch Tetracyclin aktiviert<br />

tg transgen; transgenes Tier<br />

tTA tetracycline dependent transactivator<br />

wt wildtype; nicht-transgenes Tier<br />

59


10. Anhang: Nachdruck der eigenen Publikationen<br />

In chronologischer Abfolge<br />

60


Abstract<br />

Animal Reproduction Science 79 (2003) 291–317<br />

Application of transgenesis in livestock for<br />

agriculture and biomedicine<br />

Heiner Niemann ∗ , Wilfried A. Kues<br />

Department of Biotechnology, Institut für Tierzucht Mariensee, FAL,<br />

31535 Neustadt, Germany<br />

Received 13 November 2002; received in revised form 27 January 2003; accepted 3 February 2003<br />

Microinjection of foreign DNA into pronuclei of a fertilized oocyte has predominantly been<br />

used for the generation of transgenic livestock. This technology works reliably, but is inefficient<br />

and results in random integration and variable expression patterns in the transgenic offspring.<br />

Nevertheless, remarkable achievements have been made with this technology. By targeting expression<br />

to the mammary gland, numerous heterologous recombinant human proteins have been<br />

produced in large amounts which could be purified from milk of transgenic goats, sheep, cattle<br />

and rabbit. Products such as human anti-thrombin III, -anti-trypsin and tissue plasminogen<br />

activator are currently in advanced clinical trials and are expected to be on the market within<br />

the next few years. Transgenic pigs that express human complement regulating proteins have<br />

been tested in their ability to serve as donors in human organ transplantation (i.e. xenotransplantation).<br />

In vitro and in vivo data convincingly show that the hyperacute rejection response<br />

can be overcome in a clinically acceptable manner by successful employing this strategy. It is<br />

anticipated that transgenic pigs will be available as donors for functional xenografts within a<br />

few years. Similarly, pigs may serve as donors for a variety of xenogenic cells and tissues. The<br />

recent developments in nuclear transfer and its merger with the growing genomic data allow<br />

a targeted and regulatable transgenic production. Systems for efficient homologous recombination<br />

in somatic cells are <strong>bei</strong>ng developed and the adaptation of sophisticated molecular tools,<br />

already explored in mice, for transgenic livestock production is underway. The availability of these<br />

technologies are essential to maintain “genetic security” and to ensure absence of unwanted side<br />

effects.<br />

© 2003 Elsevier B.V. All rights reserved.<br />

Keywords: Nuclear transfer; Pronuclear DNA injection; Xenotransplantation; Recombinant proteins; Gene<br />

targeting<br />

∗ Corresponding author. Tel.: +49-5034-871-148; fax: +49-5034-871-101.<br />

E-mail address: niemann@tzv.fal.de (H. Niemann).<br />

0378-4320/$ – see front matter © 2003 Elsevier B.V. All rights reserved.<br />

doi:10.1016/S0378-4320(03)00169-6


292 H. Niemann, W.A. Kues / Animal Reproduction Science 79 (2003) 291–317<br />

1. Introduction: animal breeding and biotechnology<br />

Breeding of livestock has a long standing and very successful tradition. It began with<br />

domestication by which man habituated animals to live in his proximity. Using the available<br />

methodology man has propagated useful populations of animals. The selection was predominantly<br />

done on the basis of the phenotype and specific traits. A scientifically based animal<br />

breeding has existed for approximately 50 years on the basis of the increasing knowledge<br />

in population genetics and statistics. From the early beginning, this approach incorporated<br />

biotechnological procedures for which artificial insemination (AI) is the preeminent example.<br />

In cattle AI is employed in all countries with advanced breeding programs. Recently,<br />

significant increases in AI-application are also observed in swine where currently approximately<br />

50% of the sows are artificially inseminated. This technology allows for a more<br />

efficient exploitation of the genetic potential of valuable sires and their propagation in a<br />

given population, frequently on a truly international scale. In the 1980s, embryo transfer<br />

technology has been transferred from an experimental stage to commercial application. This<br />

technology allowed for the first time a better exploitation of the genetic potential of the female<br />

in livestock breeding. As of today, >530,000 bovine embryos are transferred annually<br />

worldwide of which half are used upon freezing and thawing. In the other livestock species<br />

only very few embryos are transferred on a yearly basis (Thibier, 2001). The efforts of animal<br />

breeders and the application of artificial insemination, embryo transfer and further biotechnological<br />

procedures have resulted in the well known and remarkable increases in performances<br />

of livestock production. However, four major limitations have to be considered:<br />

1. The genetic progress reaches only 1–3% per year and is thus relatively slow.<br />

2. It is hardly or not possible to separate desired traits from non-desired traits.<br />

3. A targeted transfer of genetic information between different species has not been possible.<br />

4. Difficult adaptation to local conditions.<br />

Biotechnological procedures which have been under development during the past 20 years<br />

opened the perspectives to overcome the limitations listed above. Today, the term “biotech-<br />

Fig. 1. Generation of transgenic pigs for xenotransplantation. (A) Minigene construct for microinjection, PCMV:cytomegalovirus<br />

immediate early promoter, hCD59: human CD59 (regulator of complement) cDNA, PA: polyadenylation<br />

site, PvuI and AspI: flanking restriction enzyme sites. (B) DNA microinjection into one pronucleus of a<br />

porcine zygote. Prior to microinjection the zygote was centrifuged to separate the dark lipids from the cytoplasmatic<br />

fraction to allow optic identification of the pronuclei. (C) Transgenic founder animal identified by Southern<br />

blotting of an ear sample. (D) Organ-specific expression of hCD59 protein in an F1-offspring animal determined<br />

by Western blotting with a specific monoclonal antibody. (E) Cell surface expression of hCD59 in porcine primary<br />

cells as determined by flow cytometry. Grey shadowed curves demonstrated presence of hCD59 by usage of a monoclonal<br />

antibody, white curves indicate background values by usage of an isotype matched control antibody. (F)<br />

Functional expression of human CD59 on transgenic porcine cells as demonstrated by cytotoxicity assay. Porcine<br />

cells were incubated with heat treated human serum and a dilution series of human complement. Specific lysis of<br />

porcine cells were measured by chromium release. (G) Ex vivo perfusion of porcine kidneys from F1-animals with<br />

human blood. Nearly all transgenic porcine kidneys could be perfused for 4 h, whereas non-transgenic control<br />

kidneys failed soon after onset of perfusion due to hyperacute rejection. Mean survival times were 207.5 min for<br />

transgenic and 57.5 min for wildtype kidneys (P < 0.005) (Data from Niemann et al., 2001).


H. Niemann, W.A. Kues / Animal Reproduction Science 79 (2003) 291–317 293<br />

nology in livestock” comprises an arsenal of reproductive-biological and molecularbiological<br />

procedures. Reproductive biology includes: artificial insemination (AI), estrous<br />

synchronization, regulation of parturition, embryo transfer (ET), cryopreservation of gametes<br />

and embryos, sexing of sperm and embryos, in vitro production of embryos (IVP,<br />

including: in vitro maturation (IVM), in vitro fertilization (IVF) and in vitro culture (IVC)),<br />

embryo bisection, nuclear transfer (NT, cloning), microinjection technology (DNA, RNA,


294 H. Niemann, W.A. Kues / Animal Reproduction Science 79 (2003) 291–317<br />

Antisense). Molecular genetics comprise: genome analysis (sequencing, mapping, polymorphisms),<br />

molecular diagnostics (genetic defects (MHS, malignant hyperthermia syndrome;<br />

Dumps, deficiency of uridine monophosphate synthetase; Blad, bovine leucocyte adhesion<br />

deficiency), genetic descent, genetic diversity), functional genomics (expression patterns,<br />

interactions of genes), transgenics (additive gene transfer, knockout).<br />

Livestock biotechnology has an interdisciplinary character and comprises elements from<br />

anatomy, gynecology and obstetrics, endocrinology and physiology, andrology, ultrasoundtechnology,<br />

biochemistry, cell biology and molecular biology. The further development and<br />

refinement of bio- and gene technology with the goal to achieve commercial exploitation is<br />

considered to be an important tool to cope with future challenges in livestock production.<br />

A prominent example for biotechnological procedures with significant effects in animal<br />

production is the transgenic technology. Gene transfer is defined as the introduction of a<br />

protein coding DNA fragment into the host genome with the goal that the foreign DNA<br />

contributes to the protein synthesis of the host organism, e.g. the transgenic animal. Usually<br />

gene transfer involves gene constructs which are artificially combined DNA fragments<br />

consisting of regulatory and protein coding sequences. DNA microinjection and nuclear<br />

transfer are the two feasible technical approaches to generate transgenic livestock.<br />

Microinjection involves injection of several thousands of DNA copies into pronuclei of<br />

zygotes; zygotes are transferred into recipients and offspring is screened for integration and<br />

expression of the foreign DNA (Fig. 1A–D). Although this procedure is reliable, it is inefficient<br />

(1–4% transgenic offspring/transferred microinjected zygotes), results in random<br />

integration into the host genome and variable expression due to position effects (Pursel and<br />

Rexroad, 1993; Wall, 1996). In addition, it is time-consuming and requires substantial intellectual,<br />

financial and material resources (Seidel, 1993). Attributed to the enormous amounts<br />

of resources needed for transgenic livestock production, the costs for one expressing transgenic<br />

animal are extraordinary high. It has been calculated that one expressing transgenic<br />

mouse requires average expenses of 121 US$, whereas one expressing transgenic pig would<br />

amount to 25,000 US$, one transgenic sheep 60,000 US$ and one transgenic cow 546,000<br />

US$ (Wall et al., 1992). Transgenic production in cattle can only be practical through in<br />

vitro production of embryos. From a total of more than 36,500 microinjected zygotes ≈2300<br />

developed to blastocysts, upon transfer 28% resulted in pregnancy and 18 transgenic calves<br />

could be identified. To improve efficiency of the procedure the embryos were biopsied and<br />

analyzed by PCR for the presence of the transgene (Eyestone, 1999). The early detection of<br />

transgenesis in preimplantation embryos has been shown to be feasible, however, efficiency<br />

is limited due to an early incidence of mosaicism (Lemme et al., 1994). The propagation<br />

of the transgenic trait in a given cattle population can be accomplished through in vitro<br />

production techniques by using semen from a transgenic bull for in vitro fertilization and<br />

collecting oocytes by means of ultrasound-guided follicular aspiration from transgenic female<br />

founder animals and their subsequent use in IVF (Eyestone, 1999). Details of the<br />

microinjection technology and the potential applications of transgenic livestock have been<br />

extensively reviewed (Rexroad, 1992; Wall et al., 1992; Pursel and Rexroad, 1993; Niemann<br />

et al., 1994, 2002a; Wall, 1996; Murray, 1999).<br />

Despite the inherent limitations, microinjection has allowed commercial exploitation of<br />

transgenic technology with animals for biomedical purposes and even for few agricultural<br />

traits. Substantial progress in livestock transgenesis has been made through application of


H. Niemann, W.A. Kues / Animal Reproduction Science 79 (2003) 291–317 295<br />

somatic nuclear transfer. It is anticipated that the merger of nuclear transfer with molecular<br />

tools, such as targeted genetic modification and conditional gene expression already<br />

explored in mice, will provide another boost to livestock transgenesis (Niemann and Kues,<br />

2000).<br />

2. Current status of animal transgenesis<br />

2.1. Transgenic animals with agricultural traits<br />

An Australian group has generated transgenic pigs bearing an hMt-pGH construct that can<br />

tightly be regulated by zinc feeding. The transgenic animals show significant improvements<br />

in economically important traits such as growth rate, feed conversion and bodyfat–muscle<br />

ratio (Nottle et al., 1997, 1999). Transgenic sheep carrying a keratin-IGF-I construct show<br />

expression in the skin and the clear fleece was about 6.2% greater in transgenic versus<br />

non-transgenic animals (Damak et al., 1996a,b). In both projects no adverse effects of<br />

the transgene on health or reproduction were observed. Another interesting application<br />

could be enhanced disease resistance (Müller and Brem, 1991). A mouse model in which<br />

recombinant monoclonal antibodies, which neutralize the transmissable gastroenteritis virus<br />

(TGV), are secreted into milk, provided passive protection against gastroenteric infections to<br />

the pups (Castilla et al., 1998). The verification of this model in pigs is promising. Recently,<br />

transgenic pigs expressing a bacterial phytase gene under the transcriptional control of a<br />

salivary gland-specific promoter were shown to have improved phosphate uptake (Golovan<br />

et al., 2001). The inserted phytase gene was almost exclusively expressed in the salivary<br />

gland and enabled the pigs to digest phosphorus in phytate, which could then be metabolized<br />

by the intestine. These animals require significantly fewer inorganic phosphate supplements,<br />

release substantially reduced phosphorus levels in manure and thus reduce environmental<br />

pollution. By transgenic expression of bovine alpha-lactalbumin in the mammary gland of<br />

sows, lactation performance was improved with respect to milk composition, milk yields and<br />

piglet growth (Wheeler et al., 2001). The increased survival of piglets at weaning provides<br />

significant benefits to animal welfare and the producer.<br />

2.2. Transgenic animals in biomedicine<br />

2.2.1. Gene pharming<br />

By targeting expression to the mammary gland via the use of mammary gland-specific<br />

promoter elements, large amounts of numerous heterologous recombinant proteins have<br />

been produced. In the bovine mammary gland of transgenic cows at least two different<br />

pharmaceutical proteins have been produced, in the caprine mammary gland five proteins,<br />

in the ovine system four proteins, in the pig two and in transgenic rabbits seven proteins<br />

(Rudolph, 1999). These could be purified from milk of transgenic rabbits, sheep, goats and<br />

cattle. The biological activity of the recombinant proteins was assessed and the therapeutic<br />

effects have been characterized (Rudolph, 1999; Meade et al., 1999). Human lactoferrin has<br />

recently been reported to be produced in large amounts in the mammary gland of transgenic<br />

cows (van Berkel et al., 2002). Products such as anti-thrombin III (ATIII), -anti-trypsin


296 H. Niemann, W.A. Kues / Animal Reproduction Science 79 (2003) 291–317<br />

Table 1<br />

Status of selected recombinant pharmaceutical proteins produced in the mammary gland of transgenic livestock<br />

with regard to clinical testing/registration<br />

Donor species Product Clinical trial Treatment of Expected on<br />

market<br />

Company<br />

Sheep -AT Phase II/III -AT-deficiency;<br />

cystic fibrosis<br />

2006 PPL/GTC<br />

Goat TPA Phase II/III Coronary clots 2006 GTC<br />

Goat AT III Phase III Heparin-resistance 2004 GTC<br />

Rabbit -Glucosidasea Phase II/III Pompe’s disease 2004 Pharming<br />

a Orphan drug registration.<br />

(-AT) or tissue plasminogen activator (tPA) are currently in advanced clinical trials and<br />

are expected to be on the market within the next few years (Ziomek, 1998; Rudolph, 1999)<br />

(Table 1). On the basis of average expression levels, daily milk volumes and purification<br />

efficiency, 5400 cows would be needed to produce the 100,000 kg human serum albumin<br />

(HSA) that are required per year worldwide, 4500 sheep for the production of 5000 kg -AT,<br />

100 goats for 100 kg of monoclonal antibodies, 75 goats for the 75 kg ATIII and two pigs<br />

to produce 2 kg human clotting factor IX (Rudolph, 1999).<br />

Although a variety of proteins has been produced in the mammary gland of transgenic<br />

animals, not every protein can obviously be expressed at the desired high amounts. Erythropoietin<br />

(EPO) could not be expressed in the mammary gland of transgenic cattle (Hyttinen<br />

et al., 1994). We have shown that human clotting factor VIII cDNA constructs can be expressed<br />

in the mammary gland of transgenic mice, rabbit and sheep. However, the recovery<br />

rates of hFVIII protein were low and dependent on the donor, storage temperature and dilution<br />

of milk samples. hFVIII was rapidly sequestered in ovine milk (Halter et al., 1993;<br />

Espanion et al., 1997; Niemann et al., 1999; Hiripi et al., 2003). These latter results show<br />

that the technology needs further improvements to achieve high level expression of extraordinary<br />

large and complex regulated genes, such as hFVIII, although higher levels of hFVIII<br />

were reported in transgenic swine (Paleyanda et al., 1997).<br />

2.2.2. Xenotransplantation<br />

2.2.2.1. Transplantation of solid organs. Approximately 250,000 people are currently<br />

only living because of transplantation of an appropriate human organ (e.g. allotransplantation).<br />

In most cases no alternative therapeutic treatment was available and the recipients<br />

would have died without the organ transplantation. However, the enormous progress in organ<br />

transplantation technology which today is the basis for a normal life of thousands of patients<br />

has led to an acute shortage of appropriate organs, while the willingness to donate organs<br />

has remained unchanged or is even slightly reduced. Estimations in the USA have revealed<br />

that approximately 45,000 people, younger than 65 years need a heart transplant whereas<br />

only 2000 human hearts are transplanted annually (Michler, 1996). In the USA, more than<br />

74,000 people are awaiting organ transplants and a new person is added to the waiting list<br />

every 14 min. Only 21,000 patients received a transplant in the year 2000 (Petit-Zeman,<br />

2001). In Germany approximately 2400 kidneys, 730 livers, 540 hearts and 180 pancreas


H. Niemann, W.A. Kues / Animal Reproduction Science 79 (2003) 291–317 297<br />

are transplanted annually. However, the demand is twice as high as these figures. This has<br />

led to the sad and ethically challenging situation that several thousand patients die every<br />

year who could have survived if appropriate organs would have been available.<br />

To close this growing gap between demand and availability of appropriate organs, xenotransplantation<br />

(=the transplantation of organs between discordant species e.g. from animals<br />

to human) is considered as the solution of choice (Bach, 1998; Platt and Lin, 1998). The<br />

pig seems to be the optimal donor animal because<br />

• the organs have a similar size as human organs,<br />

• porcine anatomy and physiology are not too different from those in humans,<br />

• pigs have short reproduction cycles and large litters,<br />

• pigs grow rapidly,<br />

• maintenance is possible at high hygienic standards at relatively low costs,<br />

• pigs are a domesticated species.<br />

The process of evaluating transgenic pigs as potential donors for xenotransplants involves<br />

a variety of complex steps and is extremely time-, labor- and resource-intensive.<br />

Essential prerequisites for a successful xenotransplantation are:<br />

1. Prevention of transmission of zoonoses from the donor animal to the human recipient.<br />

This aspect gained particular significance since a few years ago it was shown that porcine<br />

endogenous retroviruses (PERV) can be produced by porcine cell lines and can even infect<br />

human cell lines (Patience et al., 1997). However, until today no infection has been<br />

found in patients that had received various forms of living porcine tissues (e.g. islet cells,<br />

insulin, skin, extracorporal liver) for up to 12 years (Paradis et al., 1999). Recent intensive<br />

research revealed that PERV do not present a noticeable risk for recipients of xenotransplantation<br />

provided all necessary precautions are made (Patience et al., 1998; Dinsmore<br />

et al., 2000; Switzer et al., 2001; Martin et al., 2002). In addition, a strain of miniature<br />

pigs has been identified which does not produce infective PERU (Oldmixon et al., 2002).<br />

2. Compatibility of the donor organs in anatomy and physiology with the human organ<br />

system, e.g. lifespan differences, growth rate, expected body weight.<br />

3. Overcoming of the immunological rejection of the transplanted organ. The immunological<br />

hurdles are as follows (White, 1996a):<br />

(a) Hyperacute rejection response (HAR) occurs within seconds or minutes. In the case<br />

of a discordant organ, e.g. from pig to human, naturally occurring antibodies react<br />

with antigenic structures on the surface of the porcine organ and induce HAR<br />

by activating the complement cascade which is achieved via the antigen–antibody<br />

complex. Ultimately, this results in the formation of the membrane attack complex<br />

(MAC). However, the complement cascade can be shut down at various points by<br />

expression of regulatory genes which prevent the formation of MAC. Well known<br />

regulators of the complement cascade are CD55 (=decay accelerating factor, DAF),<br />

CD46 (=membrane cofactor protein, MCP) or CD59. MAC disrupts the endothelial<br />

cell layer of the blood vessels which leads to lysis, thrombosis, loss of vascular<br />

integrity and ultimately to rejection of the transplanted organ.<br />

(b) Acute vascular rejection (AVR) occurs within days. Induced xenoreactive antibodies<br />

are thought to be responsible for AVR. The endothelial cells of the graft’s


298 H. Niemann, W.A. Kues / Animal Reproduction Science 79 (2003) 291–317<br />

microvasculature lose their anti-thrombic properties, attract leucocytes, monocytes<br />

and platelets leading to anemia and organ failure.<br />

(c) Cellular rejection occurs within weeks after transplantation. In this process the blood<br />

vessels of the transplanted organ are damaged by T-cells which invade the intercellular<br />

spaces and destroy the organ. This rejection is observed after allotransplantation<br />

and normally is suppressed by administration of immunosuppressive drugs. These<br />

have to be taken by the recipients for the rest of their lives.<br />

(d) Chronic rejection is a complex immunological process resulting in the rejection of<br />

the transplanted organ after several years. This process is slow and progressive and<br />

its etiology is largely unknown. The only remaining therapeutic option is another<br />

transplantation.<br />

When employing a discordant donor species such as the pig, overcoming the HAR is the<br />

preeminent goal. This cannot be achieved by administering high doses of an appropriate<br />

immunosuppressive drug as these do not affect the complement regulated rejection process.<br />

The most promising strategy to overcome the HAR is the synthesis of human complement<br />

regulatory proteins in transgenic pigs (Cozzi and White, 1995; White, 1996b; Bach, 1998;<br />

Platt and Lin, 1998). Following transplantation, the porcine organ would produce the complement<br />

regulatory protein and can thus prevent the complement attack of the recipient.<br />

Pigs transgenic for DAF have been generated and their hearts have been transplanted either<br />

heterotopically, e.g. in addition to the recipient’s own organ or orthotopically (=life supportive)<br />

into non-human primates. Upon heterotopic transplantation, the average survival<br />

of the recipients reached a maximum of 40–90 days, whereas the non-transgenic control<br />

organs were destroyed within a few minutes. The primates had to be treated with high doses<br />

of immunosuppressive drugs to maintain survival of the xenotransplant. Following moderate<br />

doses of immunosuppression survival rates of 20–25 days could be obtained (Bach<br />

et al., 1997; Cozzi et al., 2000). Employing the genomic clone of hCD46 (MCP), transgenic<br />

pigs showed a similar expression pattern for the transgene as found for the endogenous<br />

gene of the patients. Survival of a hCD46 porcine heart upon transplantation to baboons<br />

exceeded 23 days (Diamond et al., 2001). Similarly, transgenic expression of hCD59 was<br />

compatible with an extended survival of porcine hearts following transfer into primates<br />

(Fodor et al., 1994). Transplantation of hDAF-transgenic porcine kidneys was compatible<br />

with an extended survival of the recipients. The physiological function of the kidneys was<br />

maintained for up to 3–4 weeks (Zaidi et al., 1998) (Table 2). These data show that HAR<br />

can be overcome in a clinically acceptable manner by successful employing this strategy<br />

(Bach, 1998).<br />

Prior to primate experiments, extremely labor and time-consuming research is required<br />

to identify those transgenic animals with the most promising expression pattern of the transgene.<br />

Four research groups with strong links to the pharmaceutical industry have reported<br />

the generation of transgenic pigs with expression of human complement regulators. The<br />

screening of transgenic lines with suitable expression profiles is still inefficient. Previously,<br />

only one out of 30 lines showed an suitable expression pattern for xenotransplantation<br />

(Cozzi et al., 2000). A more efficient selection of transgenic pigs with an optimized expression<br />

pattern in two out of five tested lines was provided by employing the CMV promoter<br />

(Niemann et al., 2001). They provide a useful tool for the study of basic immunological


H. Niemann, W.A. Kues / Animal Reproduction Science 79 (2003) 291–317 299<br />

Table 2<br />

Success rates of RCA-transgenic porcine organs upon transplantation to primate recipients<br />

RCA Organ/kind of<br />

transplant<br />

Recipient Immunosuppression Survival<br />

hDAF Heart/heterotopic Cynomologus +++ a ∼60 days<br />

Heterotopic Cynomologus ++ b ∼90 days<br />

Orthotopic Cynomologus +++ ∼10 days<br />

Heterotopic Cynomologus + c ∼21 days<br />

Kidney/orthotopic Cynomologus ++ ∼13 days<br />

(maximum 35 days)<br />

hCD59 Heart, heterotopic Baboon ++ ∼30 h<br />

hCD46 Heart, heterotopic Baboon ++ ∼23 days<br />

a Heavy immunosuppression.<br />

b Moderate immunosuppression.<br />

c Weak immunosuppression.<br />

questions in xenotransplantation. Transgenic pigs that show high expression of hCD59 predominantly<br />

in the heart, kidney and pancreas but also other target organs, were identified<br />

and transgenic lines established (Fig. 1D and E). Transgenic endothelial cells and fibroblasts<br />

were protected against complement mediated lysis (Fig. 1F). Perfusion studies using<br />

isolated porcine kidneys employing human blood revealed a significant protective effect<br />

against HAR (Fig. 1G). Orthotopic transplantation of a CMV-hCD59 transgenic porcine<br />

kidney into cynomologous monkey was compatible with extended survival of >20 days.<br />

Another promising strategy towards successful xenotransplantation is the knockout of<br />

the antigenic structures on the surface of the porcine organ. These structures are known<br />

as 1,3--gal-epitopes and are produced from the gene for the 1,3--galactosyltransferase.<br />

Recently, the generation of piglets in which one allele of the -galactosyltransferase locus<br />

had been knocked out, was reported (Lai et al., 2002; Dai et al., 2002). These animals<br />

are now in breeding programs to obtain homozygous knockout animals. The usefulness of<br />

organs from these pigs for xenotransplantation is currently <strong>bei</strong>ng tested. The birth of piglets<br />

with disruption of both allelic loci has recently been publicly announced.<br />

The strength of the cellular response to xenografts can be so great that it is unlikely<br />

to be fully controlled by immunosuppressive treatment and transgene expression. Further<br />

improvements of the success in xenotransplantation might arise from the possibility of<br />

inducing a permanent tolerance across xenogenic barriers (Greenstein and Sachs, 1997;<br />

Auchincloss and Sachs, 1998). A promising strategy for long-term graft acceptance seems to<br />

induce a permanent chimerism via intraportal injection of embryonic stem cell like structures<br />

(Fändrich et al., 2002). Although xenotransplantation poses numerous further challenges to<br />

research, it is expected that transgenic pigs will be available as organ donors within the next<br />

5–10 years (Jones, 1996). Guidelines for the clinical application of porcine xenotransplants<br />

are currently <strong>bei</strong>ng developed in several countries or are already available (USA).<br />

2.2.2.2. Use of xenogenic cells and tissue. Another promising area of application for transgenic<br />

animals will be the supply of xenogenic cells and tissue. Several intractable diseases,<br />

disorders and injuries are associated with irreversible cell death and/or aberrant cellular func-


300 H. Niemann, W.A. Kues / Animal Reproduction Science 79 (2003) 291–317<br />

tion. Despite numerous attempts, primary human cells cannot yet be expanded well enough<br />

in culture. In the future, human embryonic stem cells may serve as a source for specific differentiated<br />

cell types that can be used in cell therapy. Xenogenic cells, in particular from the pig,<br />

hold great promise with regard to a successful cell therapy for human patients (Edge et al.,<br />

1998). These cells provide several significant advantages over other approaches, such as implantation<br />

at the optimal therapeutic location (i.e. immunoprivileged sites such as the brain),<br />

possibility for manipulation prior to transplantation to enhance cell function, banking and<br />

cryopreservation, combination with different cell types in the same graft (Edge et al., 1998).<br />

There are already numerous examples for successful application of xenogenic cell therapy.<br />

Porcine islet cells have been transplanted to diabetic patients and were shown to be at<br />

least partially functional over a limited period of time (Groth et al., 1994). Porcine fetal neural<br />

cells were transplanted into the brain of patients suffering from Parkinson’s disease and<br />

Huntington’s disease (Deacon et al., 1997; Fink et al., 2000). In a single autopsied patient<br />

the graft survived for more than 7 months and the transplanted cells formed dopaminergic<br />

neurons and glial cells. Pig neurons extended axons from the graft site into the host brain<br />

(Deacon et al., 1997). Further examples for the potential use of porcine neural cells are<br />

stroke and focal epilepsy (Björklund, 1991). Human, fetal neuronal cells have also been<br />

employed as transplants into Parkinson’s and Huntington’s disease patients. The advantages<br />

of porcine neural cells over their human counterparts are the abundant availability and<br />

the option to introduce fail safe mechanisms via suicide genes. Olfactory ensheathing cells<br />

(OECs) or Schwann cells derived from hCD59 transgenic pigs promoted axonal regeneration<br />

in rat spinal cord lesion (Imaizumi et al., 2000). Thus, cells from genetically modified<br />

pigs may serve as therapeutic measure to restore electrophysiologically functional axons<br />

across the site of a spinal cord transsection. Xenogenic porcine cells may also be useful as<br />

novel therapy for liver diseases. Upon transplantation of porcine hepatocytes to Watanabe<br />

heritable hyperlipidemic (WHHL) rabbits (a model for familial hypercholesterolemia), the<br />

xenogenic cells migrated out of the vessels and integrated into the hepatic parenchyma.<br />

The integrated porcine hepatocytes provided functional LDL receptors and thus reduced<br />

cholesterol levels by 30–60% for at least 100 days (Gunsalus et al., 1997).<br />

A clone of bovine adrenocortical cells restored adrenal function upon transplantation to<br />

adrenalectomized SCID mice. This finding shows that functional endocrine tissue can be<br />

derived from a single somatic cell (Thomas et al., 1997). Bovine neuronal cells were collected<br />

from transgenic fetuses, transplanted into the brain of rats and resulted in significant<br />

improvements of symptoms of Parkinson’s disease (Zawada et al., 1998). Furthermore,<br />

xenotransplantation of retinal pigment epithelial cells holds promise to treat retinal diseases<br />

such as macular degeneration which is associated with photoreceptor losses. Porcine<br />

or bovine fetal cardiomyocytes or myoblasts may provide a therapeutic approach for the<br />

treatment of ischemic heart disease. Similarly, xenogenic porcine cells may be valuable for<br />

the repair of skin or cartilage damage (Edge et al., 1998). In light of the emergence of more<br />

efficient protocols for genetic modification of donor pigs and new powerful immunosuppressive<br />

drugs, one can expect xenogenic cell therapy to evolve as an important therapeutic<br />

option for the treatment of human diseases.<br />

The above description demonstrates that although the challenges to generate a transgenic<br />

animal with an optimized tailored expression pattern are enormous, within less than 20 years<br />

transgenic livestock have emerged that provide valuable contributions to human health.


H. Niemann, W.A. Kues / Animal Reproduction Science 79 (2003) 291–317 301<br />

With increasing knowledge about the genetic basis of agricultural traits and improvements<br />

in the technology to generate transgenic animals, numerous further useful commercial<br />

applications will be developed.<br />

3. Improvements of transgenic technology<br />

3.1. Inducible gene expression<br />

A major advancement in transgenic technology would be tight control of transgene expression.<br />

Control elements that are known to regulate the activity of transgenes are the metallothionein<br />

promoter (see Nottle et al., 1999), heat shock promoter, or steroid responsive<br />

elements. However, these have been used with limited success attributed to low induction<br />

levels and physiological effects of the inducer elements (Yarranton, 1992). Significant improvements<br />

of the temporal control of gene expression could be achieved by employing the<br />

antibiotic tetracycline system (Gossen et al., 1995). This involves a transcriptional transactivator<br />

which has been created by fusion of the VP16 activation domain with a mutant Tet<br />

repressor from Escherichia coli. This transactivator requires the presence of tetracycline or<br />

an analogue for DNA binding and transcriptional activation (Fig. 2). It has been shown that<br />

Fig. 2. Schematic drawing of Tet-on system. In the original system two independent transgene constructs were<br />

integrated in the genome. The transactivator gene is continuously transcribed and the inactive form of the transactivator<br />

is translated. By exogenous addition (feeding) of a tetracycline analogue the transactivator becomes activated,<br />

subsequently binds to the Tet-promoter and induces expression of the transgene. Removal of the compound leads<br />

to transcriptional silencing of the transgene.


302 H. Niemann, W.A. Kues / Animal Reproduction Science 79 (2003) 291–317<br />

Fig. 3. Conditional transgene expression in porcine muscle. (A) Transgenic pigs carrying a Tet-off expression<br />

cassette show massive expression in skeletal muscle fibers. Muscle biopsies of a wildtype and a transgenic animal<br />

were stained with a specific antibody against the transgene. Arrows indicate a positive muscle fiber. (B) Conditional<br />

ablation of transgene expression. Expression levels of the transgene were determined in muscle biopsies before<br />

and after feeding with doxycycline. In total, four animals were treated with doxycyline and showed identical<br />

downregulation of expression.<br />

the presence of a tetracycline analogue led to a burst of expression in cell lines and even<br />

transgenic mice (Tet-on) (Gossen et al., 1995). This system can also be modified in a way<br />

that the presence of tetracycline suppresses expression of the target gene (Tet-off). In this<br />

system, tetracycline binds to the transactivator and shuts down expression of the transgene<br />

(Furth et al., 1994). The original technology requires two constructs, which make it unfeasible<br />

for application in livestock. However, it has been shown that fusion of both control<br />

elements into one construct allows efficient and tight control of gene expression in vivo<br />

(Schultze et al., 1996). In recent own work, the transgene expression in pigs, carrying a single<br />

Tet-off gene construct, could be regulated by doxycycline feeding (Fig. 3, unpublished<br />

data). Further studies will show the usefulness of this system for the conditional expression<br />

of transgenes suitable in xenotransplantation.<br />

3.2. Artificial chromosomes: YACs, BACs and MACs<br />

Artificial chromosomes are DNA molecules of predictable structure which are assembled<br />

in vitro from defined constituents that are similar to natural chromosomes. The first<br />

artificial chromosomes have been constructed in yeast (Saccharomyces cerevisiae)(Fig. 4).<br />

They include centromeres, telomeres, and origins of replication as essential components<br />

(Brown et al., 2000). These yeast artificial chromosomes (YACs) can be introduced into<br />

cell lines (Strauss et al., 1993). They carry much larger amounts of DNA than usually can<br />

be employed in microinjection. Microinjection of a 450 kb genomic YAC harboring the


H. Niemann, W.A. Kues / Animal Reproduction Science 79 (2003) 291–317 303<br />

Fig. 4. Schematic comparison of yeast and mammalian artificial chromosomes used for the generation of transgenic<br />

animals. Injection of yeast or bacterial artificial chromosomes (YAC, BAC) into mammalian zygotes leads to<br />

a random integration, since their structural chromosomal elements are not functional in mammalian cells. In<br />

contrast, injection of mammalian artificial chromosomes results in the inheritance of these vectors as independent<br />

chromosomal elements.<br />

murine tyrosinase gene resulted in transgenic mice which showed position independent and<br />

copy number dependent expression of the transgene. Albinism was rescued in transgenic<br />

mice and rabbits (Schedl et al., 1992, 1993; Brem et al., 1996). A 210 kb YAC construct has<br />

been microinjected into rat pronuclei and -lactoglobulin and human growth factor were<br />

expressed in the mammary gland of transgenic rats (Fujiwara et al., 1997, 1999). Artificial<br />

chromosomes can also be constructed in bacteria (BACs), which can be genetically modified<br />

easier and even allow homologous recombination. Transgenic mice were generated via<br />

pronuclear injection of BACs and germline transmission and proper expression of the transgene<br />

was achieved (Yang et al., 1997). However, up to now, transgenic livestock has not been


304 H. Niemann, W.A. Kues / Animal Reproduction Science 79 (2003) 291–317<br />

reported upon transfer of a YAC construct. This may be attributed to the inherent problems of<br />

this technology, such as difficulties to isolate YAC DNA with sufficient purity and the inherent<br />

instability with a tendency for deleting regions from the insert (Monaco and Larin, 1994).<br />

Mammalian artificial chromosomes (MAC) have been engineered by employing endogenous<br />

chromosomal elements from YACs or extra chromosomal elements from viruses or<br />

BACs and P1 artificial chromosomes (PACs) (Vos, 1997). MACs with a size of 1–5 Mb were<br />

formed by a de novo mechanism and segregated like normal chromosomes upon introduction<br />

into cell lines (Ikeno et al., 1998). A human artificial chromosome (HAC) containing<br />

the entire sequences of the human immunoglobulin heavy and light chain loci has been<br />

introduced into bovine fibroblasts, which were then used in nuclear transfer. Transchromosomal<br />

offspring were obtained that expressed human immunoglobulin in their blood.<br />

This system could be a significant step forward in the production of therapeutic polyclonal<br />

antibodies (Kuroiwa et al., 2002).<br />

Satellite-DNA based artificial chromosomes (SATAC) are neochromosomes that are<br />

formed by de novo amplification of pericentric heterochromatin yielding chromosomes<br />

from 10 to 360 megabases. These can serve as chromosomal vectors for exogenous DNA<br />

(Perez et al., 2000). Transgenic mice have been generated by microinjection of SATACs<br />

into pronuclei of zygotes. The additional chromosome showed germline transmission over<br />

three generations (Co et al., 2000). Microinjection of SATACs was also compatible with<br />

the development of bovine embryos. Transgenic embryos could be identified by staining<br />

for the presence of a reporter gene and FISH detection of the extra chromosome (Co et al.,<br />

2000). Moreover, SATACs could be isolated and purified by flow cytometry due to their<br />

high AT content.<br />

However, all efforts to assemble MACs as functional vectors for foreign DNA have met<br />

with limited success. Mini-chromosomes are considered to overcome the problems with<br />

artificial chromosomes. In contrast to artificial chromosomes, mini-chromosomes possess a<br />

defined structure, can be engineered to harbor large fragments of DNA and can go through<br />

the vertebrate germline (Brown et al., 2000). Mini-chromosomes could play an important<br />

role in the development of transchromosomal animals as models of human genetic disorders<br />

such as trisomy 21 (Down’s syndrome) (Hernandez et al., 1999). Recently, mice have<br />

been engineered that contain certain fragments of human chromosomes with each of the<br />

immunoglobulin heavy chain and light chain gene. In these animals, the endogenous immunoglobulin<br />

genes were deleted and the mice produced humanized antibodies (Tomizuka<br />

et al., 2000). Further progress in this area will ultimately result in a variety of different<br />

vector systems which can be used for large scale transgenesis in experimental animals,<br />

agricultural livestock and crop plants.<br />

4. Nuclear transfer and targeted genetic modification<br />

4.1. Nuclear transfer technology: results and limitations<br />

As microinjection has a number of significant shortcomings mentioned above, research<br />

has focused on alternate methodologies for improving the generation of transgenic livestock.<br />

These include sperm mediated DNA transfer (Gandolfi, 1998; Squires, 1999; Chang et al.,


H. Niemann, W.A. Kues / Animal Reproduction Science 79 (2003) 291–317 305<br />

2002; Lavitrano et al., 2002), the intracytoplasmic injection (ICSI) of sperm heads carrying<br />

foreign DNA (Perry et al., 1999, 2001), the use of retroviral vectors either by injection<br />

or infection of oocytes or embryos (Haskell and Bowen, 1995; Chan et al., 1998; Cabot<br />

et al., 2001) or the use of genetically modified donor cells from livestock in nuclear transfer<br />

(Schnieke et al., 1997; Cibelli et al., 1998a; Baguisi et al., 1999; Park et al., 2001). Further improvements<br />

may be derived from the adaptation of technologies that allow precise modifications<br />

of the murine genome. These include targeted chromosomal integration by site-specific<br />

DNA recombinases, such as Cre or FLP or homologous recombination that would enable<br />

generation of transgenic animals with a gain (knockin) or a loss of function (knockout)<br />

(Capecchi, 1989; Kilby et al., 1993). In light of the recent advances, somatic nuclear transfer<br />

holds the greatest promise for significant improvements in the generation of transgenic livestock.<br />

A major prerequisite is the availability of suitable primary cells or cell lines compatible<br />

with techniques for precise genetic modifications either for gain or loss of function. Another<br />

prerequisite is a significantly improved knowledge of gene sequences and organization of the<br />

livestock genome, which currently is lagging much behind that of mouse and human. In the<br />

latter, the putative 3 billion base pairs have been sequenced in the year 2001. Surprisingly, the<br />

human genome only contained approximately 30–35,000 genes (Baltimore, 2001). However,<br />

RNA editing and alternative splicing significantly augments the number of proteins<br />

synthesized from a gene. Whereas only 250 out of the 6000 genes in Saccharomyces cerevisiae<br />

contain introns, the vast majority of the estimated human genes are thought to contain<br />

these structures (Graveley, 2001; Modrek and Lee, 2001). In contrast, in livestock species<br />

only a minority of genes have been mapped and sequenced until now (Table 3). However, the<br />

technology developed during deciphering the human genome will improve and accelerate<br />

sequencing of genomes from other species (O’Brien et al., 1999). Even the currently limited<br />

genetic information in livestock species allows the application of cDNA array technologies<br />

and or high density DNA chips to obtain gene expression profiles of nearly any tissue of<br />

interest. Improvements of RNA isolation and unbiased amplification of tiny amounts of<br />

mRNA (picogram) enable to analyze even single embryos (Brambrink et al., 2002).<br />

Reports on the generation of transgenic livestock (Schnieke et al., 1997; Cibelli et al.,<br />

1998a; Park et al., 2001) via somatic nuclear transfer had raised great expectations about<br />

this elegant approach to improve the generation of transgenic livestock. Fetal fibroblasts<br />

were transfected in vitro, screened for transgene integration and then transferred into enucleated<br />

oocytes. After fusion of both components and activation of the reconstituted nuclear<br />

transfer complexes, blastocysts were transferred to synchronized recipients and gave rise<br />

to transgenic offspring. Compared with the microinjection procedure in which screening<br />

for transgenesis and optimal expression of the transgenes takes place at the level of the<br />

offspring, cloning by nuclear transfer can accelerate the time-consuming transgenic production<br />

by rapid screening in vitro and 100% transgenic offspring. The first commercial<br />

data of the use of nuclear transfer to generate transgenic cattle show the feasibility of this<br />

approach (Forsberg et al., 2001).<br />

Offspring from nuclear transfer have been born in all major livestock species cattle,<br />

sheep, goat and swine (Wilmut et al., 1997; Cibelli et al., 1998b; Baguisi et al., 1999;<br />

Polejaeva et al., 2000; Betthauser et al., 2000). A variety of different cell types of embryonic,<br />

fetal and somatic origin has been successfully employed as donors in nuclear transfer.<br />

However, the overall efficiency of nuclear transfer is low (Colman, 2000). Factors affecting


306 H. Niemann, W.A. Kues / Animal Reproduction Science 79 (2003) 291–317<br />

Table 3<br />

Whole genome sequencing<br />

Species No. annotated genes Genome size<br />

(×10 6 bp)<br />

Sequencing<br />

completed<br />

Milestones in genome sequencing and size of selected genomes<br />

Haemophilus influenca (bacteria) ∼1700 1.7 1995<br />

Saccaromyces cerevisae (fungi) ∼6000 12.0 1996<br />

Caenorhabditis elegans (nematodes) ∼18000 97.0 1998<br />

Drosophila melanogaster (insects) ∼14000 137.0 2000<br />

Arabidopsis thalia (plants) ∼25000 125.0 2000<br />

Homo sapiensa (mammals) ∼35000 3300.0 2001<br />

Diversity of genome sequencing approaches (10/2002)<br />

Archaea Complete genomes of 16 species<br />

Bacteria Complete genomes of 81 species<br />

Eukaryota complete genomes of 9 species.<br />

In addition the genomes of Homo<br />

sapiens, Mus musculus, Rattus<br />

norvegicus, Danio rerio and<br />

some crop plants are mapped,<br />

however gene annotation and/or<br />

sequencing of untranscribed<br />

regions is still in progress<br />

Species Chromosome<br />

number (haploid)<br />

No. mapped<br />

genes<br />

Genome size<br />

(×1000)<br />

Human b 23 >30000 3.300<br />

Mouse c 20 30000 2.450<br />

Cattle c 30 ∼1000 ∼3.500<br />

Sheep c 27 ∼500 ∼3.100<br />

Pig c 19 ∼600 ∼2.400<br />

Horse c 32 200 ∼4.000<br />

a Gene annotation still in progress.<br />

b February 2001 sequencing completed.<br />

c Information for 2002.<br />

the success of nuclear transfer are poorly defined and the average percentage of live offspring<br />

does not exceed 1–3% of the transferred reconstituted embryos (Wakayama et al., 1998;<br />

Wilmut et al., 2002). A better understanding of the underlying fundamental molecular<br />

and cellular processes, such as cell cycle compatibilities between recipient cytoplasm and<br />

donor nucleus (Campbell et al., 1996), cell cycle synchronization of the donor cells (Boquest<br />

et al., 1999; Kues et al., 2000), reprogramming and the relevance of differentiation versus<br />

totipotency is urgently needed. Upon serum deprivation or treatment with chemical cell<br />

cycle inhibitors, the majority of porcine donor cells was synchronized at the presumptive<br />

optimal cell cycle stage at G0/G1 without compromising their viability (Kues et al., 2000,<br />

2002; Anger et al., 2003). This contributes substantially to standardize the nuclear transfer<br />

procedures. In addition, methods have to be established that allow reliable determination


H. Niemann, W.A. Kues / Animal Reproduction Science 79 (2003) 291–317 307<br />

of the capacity of a given nuclear transfer embryo to develop into a normal offspring. The<br />

large offspring syndrome (LOS) including an increased peri- and postnatal mortality, is<br />

found in offspring, predominantly from ruminants, derived from nuclear transfer embryos<br />

(Wilmut et al., 1997; Young et al., 1998; Kato et al., 1998). These unwanted side effects<br />

need to be overcome prior to an eventual commercial exploitation. Aberrations of the well<br />

orchestrated pattern of gene expression are thought to be involved in the high incidence of<br />

LOS. A primary mechanism may be alterations in the methylation of genes, including those<br />

that are subject to imprinting (Young et al., 1998; Niemann and Wrenzycki, 2000; Niemann<br />

et al., 2002b).<br />

4.2. Embryonic and adult cell lines<br />

The fundamental tools for loss-of-function transgenics in the mouse are the availability<br />

of embryonic stem cells (ES cells), homologous recombination and the high probability<br />

with which ES cells give rise to germline contribution after injection into host blastocysts.<br />

This provides a powerful approach to introduce specific genetic changes into the murine<br />

genome (Evans and Kaufman, 1981; Martin, 1981). The essential characteristics of ES cells<br />

include derivation from the preimplantation embryo (inner cell mass cells), undifferentiated<br />

and indefinitive proliferation in vitro and the developmental potential to differentiate into<br />

all cell types under appropriate conditions.<br />

Besides ES cells, embryonic germ (EG) and embryonic carcinoma (EC) cells have been<br />

established in the mouse model. EG cells are isolated from cultured primordial germ cells<br />

(PGC) (Matsui et al., 1992; Resnick et al., 1992) and share several characteristics with<br />

ES cells, including morphology, pluripotency, and the capacity for germline transmission.<br />

Similar to ES cells, EG cells express alkaline phosphatase and Oct-4 and can be aggregated<br />

to form embryoid bodies.<br />

The ultimate criterion for true totipotent stem cell lines is the contribution to the germline<br />

in chimeras or starting a new development upon nuclear transfer. ES or EG-like cell lines<br />

have been isolated from sheep, goat, pig and cattle (Wheeler, 1994; Anderson, 1999). Porcine<br />

and bovine cell lines were capable to contribute to chimera formation upon injection into<br />

appropriate host blastocysts (Wheeler, 1994; Shim et al., 1997; Piedrahita et al., 1998; Cibelli<br />

et al., 1998a). However, no germline transmission has been reported so far. True totipotent<br />

stem cell lines might require specific culture conditions, growth factor supplements and<br />

probably a specific genetic background, as only few mouse strains are suitable for ES cell<br />

line isolation (Hogan et al., 1994).<br />

Recent experiments in mice have revealed that adult cells from the hematopoetic lineage<br />

possess a greater plasticity than previously assumed (Jiang et al., 2002). Besides<br />

long-term proliferation in vitro, these mesenchymal derived cells express embryonic markers,<br />

such as Oct-4, rex and telomerase and can be differentiated into several cell types<br />

in vitro and in vivo. Upon injection into blastocysts, chimeric mice could be generated,<br />

which showed a high percentage of chimerism in nearly all organs. Similar cells were<br />

isolated from the hematopoetic system of rats and even humans, suggesting that mesenchymal<br />

stem cells could also be a useful source of research and application in livestock<br />

species.


308 H. Niemann, W.A. Kues / Animal Reproduction Science 79 (2003) 291–317<br />

4.3. Homologous recombination and nucleus donor cell strains<br />

Homologous recombination in murine ES cells is the most straightforward approach to<br />

eliminate gene function and is therefore the preferred method to establish a null genotype.<br />

Several strategies for gene targeting in murine ES cells have been developed (Kühn<br />

et al., 1995; Mayford et al., 1997). More than 1000 knockout strains have been created<br />

via gene targeting in embryonic stem cells (mouse knockout and mutation database<br />

http://www.biomednet.com). The potential for a knockout technology in livestock production<br />

is highlighted by the discovery that several beef cattle breeds, like Belgian Blue<br />

and Piedmontese, are accidentally homozygous for the mutated myostatin gene, which<br />

is functionally inactive and could be referred to as a natural knockout (McPherron and<br />

Lee, 1997; Kambadur et al., 1997; Grobet et al., 1997). The similarity in phenotypes<br />

of myostatin mutated cattle and myostatin null mice (McPherron et al., 1997) is striking<br />

and suggests that myostatin could be a useful target for genetic modification in farm<br />

animals.<br />

Nuclear transfer techniques promise to circumvent the need for true totipotent cells for<br />

the generation of loss-of-function transgenic livestock. The future challenge for the production<br />

of transgenic livestock is the isolation and handling of primary cell cultures, either<br />

from somatic or embryonic origin (Schnieke et al., 1997; Cibelli et al., 1998b; Kues<br />

et al., 1998). These could be used for sophisticated genetic modifications, clonal selection<br />

and subsequently for nuclear transfer. Gene targeting in somatic cells of livestock<br />

has important applications in combination with nuclear transfer (Fig. 5). Recent progress<br />

in the generation of gene targeted livestock clearly shows the significant potential of this<br />

field. An efficient and reproducible targeting protocol in fetal fibroblasts by which the<br />

transgenic construct -lactoglobulin--anti-trypsin was placed at the ovine 1 precollagen<br />

locus was described and the production of live sheep upon nuclear transfer was reported<br />

(McCreath et al., 2000). The recombinant -AT was highly expressed in the mammary<br />

gland. However, the proportion of embryonic and fetal losses was significantly increased<br />

in pregnancies with targeted NT derived sheep and porcine embryos. The first piglets, carrying<br />

a knockout for one allele of the -galactosyltransferase gene did not show gross<br />

abnormalities (Lai et al., 2002; Dai et al., 2002). By performing two rounds of targeting<br />

in cell culture, both alleles can be knocked out and time-consuming breeding will<br />

be avoided in future (Fig. 5). Exploiting the targeting technology would allow to produce<br />

large amounts of human serum albumin (HSA) when inserted at the BSA locus in<br />

transgenic cattle. A knockout of the PrP gene locus could render cattle non-susceptible<br />

to BSE. Disease models in large animals would be another promising application models<br />

for targeted transgenesis. The pig is obviously a good model for human eye disease<br />

(Petters et al., 1997; Theuring et al., 1997) and sheep could be a good model for cystic<br />

fibrosis.<br />

4.4. Conditional mutagenesis and region-specific knockouts<br />

The above findings demonstrate the power of gene targeting, but they also point to a need<br />

for additional genetic techniques when precise genetic modifications are desired. Gene<br />

knockouts per se have no spatial or temporal restriction. As the targeted gene product


H. Niemann, W.A. Kues / Animal Reproduction Science 79 (2003) 291–317 309<br />

Fig. 5. Schematic drawing of gene targeting in livestock. Gene targeting of somatic primary cells by homologous<br />

recombination employing a promoterless targeting vector. Optionally, cell clones with the desired targeting event<br />

could be screened for loss of heterozygosity and subsequently be employed in nuclear transfer.<br />

is absent for the entire life of the animal in all cells, it is difficult to assign a phenotype<br />

to a specific knockout as a mutant organism may compensate for the loss of a gene<br />

product or the knockout may have complex, secondary effects, depending on the genetic<br />

background (Gerlai, 1996), or may even result in embryonic lethality. A powerful<br />

tool for the design of genetic switches and for accelerating the creation of genetically<br />

modified animals, is the Cre site-specific DNA recombinase of bacteriophage P1 (Sauer,<br />

1998). A single 38 kDa Cre protein is required to catalyze recombination between two<br />

loxP recognition sites, which are 34 bp DNA sequences. Recombination can occur between<br />

directly repeated loxP sites on the same molecule to excise the intervening DNA<br />

sequence, irrespective of whether the recognition sites are located on a plasmid or an<br />

mammalian chromosome (Sauer and Henderson, 1988). The combination of tissue-specific<br />

promoter elements with Cre DNA recombinase, enables a gene to be knocked out of a<br />

restricted cell type or tissue (Orban et al., 1992; Gu et al., 1994). In principle, recombinase<br />

mediated recombination should allow gene replacement, e.g. the exchange of the<br />

reading frames of milk proteins with the sequences of genes encoding pharmaceutical<br />

proteins.


310 H. Niemann, W.A. Kues / Animal Reproduction Science 79 (2003) 291–317<br />

5. Perspectives and outlook<br />

The merger of the recent advancements in the reproductive technologies with the tools<br />

of molecular biology opens the horizon for a completely new era in animal production.<br />

However, major prerequisites will be the continuous refinement of reproductive biotechnologies<br />

and a rapid increase of livestock genome mapping. This would permit exploitation<br />

the great potential of bio- and gene technology with regard to a diversified animal production.<br />

One attractive example could be dairy production (Bawden et al., 1994). Apart from<br />

conventional dairy products, it could be possible to produce fat-reduced or even fat-free<br />

milk via knockout of enzymes involved in lipid metabolism, to increase curd and cheese<br />

production by enhancing expression of the casein gene family in the mammary gland, to<br />

increase efficiency of the production for creamers and liquor by increasing the proportion<br />

of -casein in milk, to create “hypoallergic” milk by knockout of the -lactoglobulin gene,<br />

to generate lactose-free milk via knockout of -lactalbumin which is the key molecule in<br />

milk sugar synthesis, to produce small infant milk in which human lactoferrin is abundantly<br />

available or to produce milk with a highly improved hygienic standard via an increased level<br />

of lysozyme or other anti-microbiological substances in the udder. Lactose free-milk could<br />

render dairy products ready for consumption by a large proportion of the world’s population<br />

who do not possess the active enzyme lactase in their intestine. In addition, calculations<br />

have revealed that reduction of the fat level in milk from the current level of 3.8–2% would<br />

permit an increase in the proportion of roughage in the feed from 55 to more than 80% and<br />

concomitantly to a significant reduction in the concentrated feed (Yom and Bremel, 1993).<br />

This would reduce the price and cost of animal nutrition. Such targeted production is also<br />

imaginable for other areas of animal production. However, one has to take into account that<br />

commercial application of this technology for agricultural traits will likely not be possible<br />

before the next decade. The biomedical area will see the first commercial application before<br />

the year 2010 (Table 4).<br />

On a broader scale, bio- and gene technology as mentioned above will be important tools<br />

to cope with the challenges in animal production in the years ahead. In light of the worldwide<br />

problems in the management of a proper environment and resources, biotechnology<br />

can contribute to quantitative and qualitative increases in food production, cost reduction,<br />

Table 4<br />

Projections for field application of transgenic livestock a<br />

Year<br />

Biomedical traits<br />

Recombinant pharmaceutical proteins 2005<br />

Xenotransplantation of solid organs<br />

Agricultural traits<br />

>2008<br />

Dairy products >2010<br />

Meat and meat products >2008<br />

Wool products >2005<br />

a Application will likely be different between USA and Europe because of the controversial debate on genetic<br />

engineering in Europe.


H. Niemann, W.A. Kues / Animal Reproduction Science 79 (2003) 291–317 311<br />

environmental protection, maintenance of genetic resources and improvements in animal<br />

welfare as well as the above mentioned diversified production. However, it should be kept<br />

in mind that putative biomedical applications like xenotransplantation or usage of transgenic<br />

farm animals for food production will require strict standards on “genetic security”<br />

and reliable and sensitive methods for the molecular characterization of the “products”.<br />

A major contribution towards this goal will come from DNA chips or arrays establishing<br />

“fingerprint” profiles at the transcriptional and/or protein level and allowing in depth insight<br />

into the proper function of a transgenic organism. Besides the technical problems,<br />

public acceptance of recombinant products, organs or food has to be taken into account.<br />

Additionally, welfare of the transgenic animals must be secured and pain or suffering due<br />

to the genetic modification could not be tolerated (Fox, 1988; Hughes et al., 1996).<br />

Acknowledgements<br />

The authors express their sincere gratitude to Christa Knöchelmann for the careful preparation<br />

of this manuscript.<br />

References<br />

Anderson, G.B., 1999. Embryonic stem cells in agricultural species. In: Murray, J.D., Anderson, G.B., Oberbauer,<br />

A.M., McGloughlin, M.M. (Eds.), Transgenic Animals in Agriculture. CABI Publishing, New York, USA,<br />

pp. 57–66.<br />

Anger, M., Kues, W.A., Klima, J., Mielenz, M., Kubelka, M., Motlik, J., Esner, M., Dvorak, P., Carnwath, J.W.,<br />

Niemann, H., 2003. Cell cycle based expression of Plk1 in synchronized porcine fetal fibroblasts, Mol. Reprod.<br />

Dev. 65, 245–253.<br />

Auchincloss Jr., H., Sachs, D.H., 1998. Xenogenic transplantation. Annu. Rev. Immunol. 16, 433–470.<br />

Bach, F.H., 1998. Xenotransplantation: problems and prospects. Ann. Rev. Mol. 49, 301–310.<br />

Bach, F.H., Ferran, C., Soares, M., Wrighton, C.J., Anrather, J., Winkler, H., Robson, S.C., Hancock, W.W., 1997.<br />

Modification of vascular responses in xenotransplantation: inflammation and apoptosis. Nat. Med. 3, 944–948.<br />

Baguisi, A., Behboodi, E., Melican, D., Pollock, J.S., Destrempes, M.M., Cammuso, C., Williams, J.L., Nims,<br />

S.D., Porter, C.A., Midura, P., Palacios, M.J., Ayres, S.L., Denniston, R.S., Hayes, M.L., Ziomek, C.A., Meade,<br />

H.M., Godke, R.A., Gavin, W.G., Overström, E.W., Echelard, Y., 1999. Production of goats by somatic cell<br />

nuclear transfer. Nat. Biotechnol. 17, 456–461.<br />

Baltimore, D., 2001. Our genome unveiled. Nature 409, 814–816.<br />

Bawden, W., Passey, R.J., MacKinlay, A.G., 1994. The genes encoding the major milk specific proteins and their<br />

use in transgenic studies and protein engineering. Biotechnol. Genet. Eng. Rev. 12, 89–137.<br />

Betthauser, J., Forsberg, E., Augenstein, M., Childs, L., Eilertsen, K., Enos, J., Forsythe, T., Golueke, P., Koppang,<br />

R., Lesmeister, T., Mallon, K., Mell, G., Misica, P., Pace, M., Pfister-Genskow, M., Strelchenko, N., Voelker,<br />

G., Watt, S., Thompson, S., Bishop, M., 2000. Production of cloned pigs from in vitro systems. Nat. Biotechnol.<br />

18, 1055–1059.<br />

Björklund, A., 1991. Neural transplantation—an experimental tool with clinical possibilities. Trends Neurosci.<br />

14, 319–322.<br />

Boquest, A.C., Day, B.N., Prather, R.S., 1999. Flow cytometric cell cycle analysis of cultured porcine fetal<br />

fibroblasts cells. Biol. Reprod. 60, 1013–1019.<br />

Brambrink, T., Wabnitz, P., Halter, R., Klocke, R., Carnwath, J., Kues, W.A., Wrenzycki, C., Paul, D., Niemann,<br />

H., 2002. Application of cDNA arrays to monitor mRNA profiles in single preimplantation mouse embryos.<br />

Biotechniques 33, 376–378.


312 H. Niemann, W.A. Kues / Animal Reproduction Science 79 (2003) 291–317<br />

Brem, G., Besenfelder, U., Aigner, B., Müller, M., Liebl, I., Schütz, G., Montoliu, L., 1996. YAC transgenesis in<br />

farm animals: rescue of albinism in rabbits. Mol. Reprod. Dev. 44, 56–62.<br />

Brown, W.R.A., Mee, P.J., Shen, M.H., 2000. Artificial chromosomes: ideal vectors? TIBTECH 18, 218–223.<br />

Cabot, R.A., Kuhholzer, B., Chan, A.W., Lai, L., Park, K.W., Chong, K.Y., Schatten, G., Murphy, C.N., Abeydeera,<br />

L.R., Day, B.N., Prather, R.S., 2001. Transgenic pigs produced using in vitro matured oocytes infected with a<br />

retroviral vector. Anim. Biotechnol. 12, 205–214.<br />

Campbell, K.H., McWhir, J., Ritchie, W.A., Wilmut, I., 1996. Sheep cloned by nuclear transfer from a cultured<br />

cell line. Nature 380, 64–66.<br />

Capecchi, M.R., 1989. The new mouse genetics: altering the genome by gene targeting. TIG 5, 70–76.<br />

Castilla, J., Pintado, B., Sola, I., Sánchez-Morgado, J.M., Enjuanes, L., 1998. Engineering passive immunity in<br />

transgenic mice secreting virus-neutralizing antibodies in milk. Nat. Biotechnol. 16, 349–354.<br />

Chan, A.W.S., Homan, E.J., Ballou, L.U., Burns, J.C., Bremel, R.D., 1998. Transgenic cattle produced by<br />

reverse-transcribed gene transfer in oocytes. Proc. Natl. Acad. Sci. USA 95, 14028–14033.<br />

Chang, K., Qian, J., Jiang, M., Liu, Y.H., Wu, M.C., Chen, C.D., Lai, C.K., Lo, H.L., Hsiao, C.T., Brown, L., Bolen<br />

Jr., J., Huang, H.I., Ho, P.Y., Shih, P.Y., Yao, C.W., Lin, W.J., Chen, C.H., Wu, F.Y., Lin, Y.J., Xu, J., Wang, K.,<br />

2002. Effective generation of transgenic pigs and mice by linker based sperm-mediated gene transfer. BMC<br />

Biotechnol. 2, 5.<br />

Cibelli, J.B., Stice, S.L., Golueke, P.L., Kane, J.J., Jerry, J., Blackwell, C., Ponce de Leon, F.A., Robl, J.M., 1998a.<br />

Transgenic bovine chimeric offspring produced from somatic cell-derived stem like cells. Nat. Biotechnol. 16,<br />

642–646.<br />

Cibelli, J.B., Stice, S.L., Golueke, P.L., Kane, J.J., Jerry, J., Blackwell, C., Ponce de Leon, F.A., Robl, J.M., 1998b.<br />

Cloned calves produced from nonquiescent fetal fibroblasts. Science 280, 1256–1258.<br />

Co, D.O., Borowski, A.H., Leung, J.D., Kaa van der, J., Hengst, S., Platenburg, G., Pieper, F.R., Perez, C.F., Jirik,<br />

F.R., Drayer, J.I., 2000. Generation of transgenic mice and germline transmission of a mammalian artificial<br />

chromosome introduced into embryos by pronuclear microinjection. Chrom. Res. 8, 183–191.<br />

Colman, A., 2000. Somatic cell nuclear transfer in mammals: progress and applications. Cloning 1, 185–200.<br />

Cozzi, E., White, D.J.G., 1995. The generation of transgenic pigs as potential organ donors for humans. Nat. Med.<br />

1, 964–966.<br />

Cozzi, E., Masroar, S., Soni, B., Vial, C., White, D.J., 2000. Progress in xenotransplantation. Clin. Neprol. 53,<br />

13–18.<br />

Dai, Y., Vaught, T.D., Boone, J., Chen, S.H., Phelps, C.J., Ball, S., Monahan, J.A., Jobst, P.M., McCreath, K.J.,<br />

Lamborn, A.E., Cowell-Lucero, J.L., Wells, K.D., Colman, A., Polejaeva, I.A., Ayares, D.L., 2002. Targeted<br />

disruption of the alpha 1,3-galactosyltransferase gene in cloned pigs. Nat. Biotechnol. 20, 251–255.<br />

Damak, S., Jay, N.P., Barrell, G.K., Bullock, D.W., 1996a. Targeting gene expression to the wool follicle in<br />

transgenic sheep. Biotechnology 14, 181–184.<br />

Damak, S., Su, H., Jay, N.P., Bullock, D.W., 1996b. Improved wool production in transgenic sheep expressing<br />

insulin-like growth factor 1. Biotechnology 14, 185–188.<br />

Deacon, T., Schumacher, J., Dinsmore, J., Thomas, C., Palmer, P., Kott, S., Edge, A., Penney, D., Kassissieh, S.,<br />

Dempsey, P., Isacson, O., 1997. Histological evidence of fetal pig neutral cell survival after transplantation into<br />

a patient with Parkinson’s disease. Nat. Med. 3, 350–353.<br />

Diamond, L.E., Quinn, C.M., Martin, M.J., Lawson, J., Platt, J.L., Logan, J.S., 2001. A human CD46 transgenic<br />

pig model system for the study of discordant xenotransplantation. Transplantation 71, 132–142.<br />

Dinsmore, L.E., Manhardt, C., Raineri, R., Jacoby, D.B., Moore, A., 2000. No evidence for infection of human<br />

cells with porcine endogenous retrovirus (PERV) after exposure to porcine fetal neuronal cells. Transplantation<br />

71, 132–142.<br />

Edge, A.S.B., Gosse, M.E., Dinsmore, J., 1998. Xenogenic cell therapy: current progress and future developments<br />

in porcine cell transplantation. Cell Transplant. 7, 525–539.<br />

Espanion, G., Halter, R., Wrenzycki, C., Carnwath, J.W., Herrmann, D., Lemme, E., Niemann, H., Paul, D.,<br />

1997. mRNA expression of human blood coagulation factor VIII (FVIII) gene constructs in transgenic mice.<br />

Transgenics 2, 175–182.<br />

Evans, M.J., Kaufman, M.H., 1981. Establishment in culture of pluripotential cells from mouse embryos. Nature<br />

292, 154–156.<br />

Eyestone, W.H., 1999. Production and breeding of transgenic cattle using in vitro embryo production technology.<br />

Theriogenology 51, 509–517.


H. Niemann, W.A. Kues / Animal Reproduction Science 79 (2003) 291–317 313<br />

Fändrich, F., Lin, X., Chai, G., Schulze, M., Ganten, D., Bader, M., Holle, J., Huang, D.S., Parwaresch, R.,<br />

Zavazava, N., Binas, B., 2002. Preimplantation-stage stem cells induce long-term allogeneic graft acceptance<br />

without supplementary host conditioning. Nat. Med. 8, 171–178.<br />

Fink, J.S., Schumacher, J.M., Ellias, S.L., Palmer, E.P., Saint-Hilaire, M., Shannon, K., Penn, R., Starr, P.,<br />

VanHoorne, C., Kott, H.S., Dempsey, P.K., Fischman, A.J., Raineri, R., Manhart, D., Dinsmore, J., Isacson, O.,<br />

2000. Porcine xenografts in Parkinson’s disease and Huntington’s disease patients: preliminary results. Cell<br />

Transplant. 2, 273–278.<br />

Fodor, W.L., Williams, B.L., Matis, L.A., Madri, J.A., Rollins, S.A., Knight, J.W., Velander, W., Squinto, S.P.,<br />

1994. Expression of a functional human complement inhibitor in a transgenic pig as a model for the prevention<br />

of xenogenic hyperacute organ rejection. Proc. Natl. Acad. Sci. USA 91, 11153–11157.<br />

Forsberg, E.J., Betthauser, J., Strelchenko, N., Golueke, P., Childs, L., Jurgella, R., Koppang, G., Mell, G., Pace,<br />

M.M., Bishop, M., 2001. Cloning non-transgenic and transgenic cattle. Theriogenology 55, 269 (abstract).<br />

Fox, M.W., 1988. Genetic engineering biotechnology: animal welfare and environmental concerns. Appl. Anim.<br />

Behav. Sci. 20, 83–94.<br />

Fujiwara, Y., Miwa, M., Takahashi, R., Hirabayashi, M., Suzuki, T., Ueda, M., 1997. Position-independent and<br />

high-level expression of human alpha-lactalbumin in the milk of transgenic rats carrying a 210-kb YAC DNA.<br />

Development 124, 3621–3632.<br />

Fujiwara, Y., Miwa, M., Takahashi, R., Hirabayashi, M., Suzuki, T., Ueda, M., 1999. High-level expressing YAC<br />

vector for transgenic animal bioreactors. Mol. Reprod. Dev. 52, 414–420.<br />

Furth, P.A., Onge, L., Böger, H., Gruss, P., Gossen, M., Kistner, A., Bujard, H., Hennighausen, L., 1994. Temporal<br />

control of gene expression in transgenic mice by a tetracycline-responsive promoter. Proc. Natl. Acad. Sci.<br />

USA 91, 9302–9306.<br />

Gandolfi, F., 1998. Spermatozoa, DNA binding and transgenic animals. Transgenic Res. 7, 147–155.<br />

Gerlai, R., 1996. Gene-targeting studies of mammalian behavior: is it the mutation or the background genotype.<br />

Trends Neurosci. 19, 177–181.<br />

Golovan, S.P., Meidinger, R.G., Ajakaiye, A., Cottrill, M., Wiederkehr, M.Z., Barney, D.J., Plante, C., Pollard,<br />

J.W., Fan, M.Z., Hayes, M.A., Laursen, J., Hjorth, J.P., Hacker, R.R., Phillips, J.P., Forsberg, C.W., 2001. Pigs<br />

expressing salivary phytase produce low-phosphorus manure. Nat. Biotechnol. 19, 741–745.<br />

Gossen, M., Freundlieb, S., Bender, G., Müller, G., Hillen, W., Bujard, H., 1995. Transcriptional activation by<br />

tetracyclines in mammalian cells. Science 268, 1766–1769.<br />

Graveley, B.R., 2001. Alternative splicing: increasing diversity in the proteomic world. Trends Genet. 17, 100–107.<br />

Greenstein, J., Sachs, D.H., 1997. The use of tolerance for transplantation across xenogenic barriers. Nat.<br />

Biotechnol. 15, 235–238.<br />

Grobet, L., Martin, L.J., Poncelet, D., Pirottin, D., Brouwers, B., Riquet, J., Schoeberlein, A., Dunner, S., Menissier,<br />

F., Massabanda, J., Fries, R., Hanset, R., Georges, M., 1997. A deletion in the bovine myostatin gene causes<br />

the double-muscled phenotype in cattle. Nat. Genet. 17, 71–74.<br />

Groth, C.G., Korsgren, O., Tibell, A., Tollemar, J., Möller, E., Bolinder, J., Östman, J., Reinholt, F.P., Hellerström,<br />

C., Andersson, A., 1994. Transplantation of porcine fetal pancreas to diabetic patients. Lancet 344, 1402–1404.<br />

Gu, H., Marth, J.D., Orban, P.C., Mossmann, H., Rajewsky, K., 1994. Deletion of a DNA polymerase gene<br />

segment in T cells using cell type-specific gene targeting. Science 265, 103–106.<br />

Gunsalus, J.R., Brady, D.A., Coulter, S.M., Gray, B.M., Edge, A.S.B., 1997. Reduction of serum cholesterol in<br />

Watanabe rabbits by xenogenic hepatocellular transplantation. Nat. Med. 3, 49–53.<br />

Halter, R., Carnwath, J., Espanion, G., Herrmann, D., Lemme, E., Niemann, H., Paul, D., 1993. Strategies to<br />

express factor VIII gene constructs in the ovine mammary gland. Theriogenology 39, 137–149.<br />

Haskell, R.E., Bowen, R.A., 1995. Efficient production of transgenic cattle by retroviral infection of early embryos.<br />

Mol. Reprod. Dev. 40, 386–390.<br />

Hernandez, D., Mee, P.J., Martin, J.E., Tybulewicz, V.L., Fisher, E.M., 1999. Transchromosomal mouse embryonic<br />

stem cell lines and chimeric mice that contain freely segregating segments of human chromosome 21. Hum.<br />

Mol. Genet. 8, 923–933.<br />

Hiripi, L., Makovics, F., Baranyi, M., Paul, D., Carnwath, J.W., Bösze, Z., Niemann, H., 2003. Expression of active<br />

human blood clotting factor VIII in the mammary gland of transgenic rabbits. DNA Cell Biol., 22, 41–45.<br />

Hogan, B., Constantini, F., Lacy, E., 1994. Manipulating the Mouse Embryo: A Laboratory Manual. Cold Spring<br />

Harbour Laboratory, Cold Spring Harbour Press, Plainview, NY.


314 H. Niemann, W.A. Kues / Animal Reproduction Science 79 (2003) 291–317<br />

Hughes, B.O., Hughes, G.S., Waddington, D., Appleby, M.-C., 1996. Behavioural comparison of transgenic and<br />

control sheep: movement order, behaviour on pasture and in covered pens, behaviour on pasture and in covered<br />

pens. J. Anim. Sci. 63, 91–101.<br />

Hyttinen, J.-M., Peura, T., Tolvanen, M., Aalto, J., Alhonen, L., Sinervirta, R., Halmekytö, M., Myöhänen, S.,<br />

Jänne, J., 1994. Generation of transgenic dairy cattle from transgene-analyzed and sexed embryos produced in<br />

vitro. Biotechnology 12, 606–608.<br />

Ikeno, M., Grimes, B., Okazaki, T., Nakano, M., Saitoh, K., Hoshino, H., McGill, I., Cooke, H., Masumoto, H.,<br />

1998. Construction of YAC-based mammalian artificial chromosomes. Nat. Biotechnol. 16, 431–439.<br />

Imaizumi, T., Lankford, K.L., Burton, W.V., Fodor, W., Kocsis, J.D., 2000. Xenotransplantation of transgenic pig<br />

olfactory ensheathing cells promotes axonal regeneration in rat spinal cord. Nat. Biotechnol. 18, 949–953.<br />

Jiang, Y., Jahagirdar, B.N., Reinhardt, R.L., Schwartz, R.E., Keene, C.D., Ortiz-Gonzalez, X.R., Reyes, M., Lenvik,<br />

T., Lund, T., Blackstad, M., Du, J., Aldrich, S., Lisberg, A., Low, W.C., Largaespada, D.A., Verfaillie, C.M.,<br />

2002. Pluripotency of mesenchymal stem cells derived from adult marrow. Nature 418, 41–49.<br />

Jones, I., 1996. 2010—a pig odyssey. Nat. Biotechnol. 14, 698–699.<br />

Kambadur, R., Sharma, M., Smith, T.P., Bass, J.J., 1997. Mutations in myostatin (GDF8) in double-muscled<br />

Belgian Blue and Piedmontese cattle. Genome Res. 7, 910–916.<br />

Kato, Y., Tani, T., Sotomaru, Y., Kurokawa, K., Kato, J., Doguchi, H., Yasue, H., Tsunoda, Y., 1998. Eight calves<br />

cloned from somatic cells of a single adult. Science 282, 2095–2098.<br />

Kilby, N.J., Snaith, M.R., Murray, J.A.H., 1993. Site-specific recombinases: tools for genome engineering. TIG<br />

9, 413–421.<br />

Kues, W.A., Carnwath, J.W., Paul, D., Niemann, H., 1998. A DNA-vector-based selection strategy for totipotent<br />

porcine cell lines. Theriogenology 49, 223 (abstract).<br />

Kues, W.A., Anger, M., Carnwath, J.W., Paul, D., Motlik, J., Niemann, H., 2000. Cell cycle synchronization of<br />

porcine fetal fibroblasts: effects of serum deprivation and reversible cell cycle inhibitors. Biol. Reprod. 62,<br />

412–419.<br />

Kues, W.A., Carnwath, J.W., Paul, D., Niemann, H., 2002. Cell cycle synchronization of porcine fetal fibroblasts<br />

by serum deprivation initiates a nonconventional form of apoptosis. Cloning Stem Cells 4, 231–243.<br />

Kühn, R., Schwenk, F., Aguet, M., Rajewsky, K., 1995. Inducible gene targeting in mice. Science 269, 1427–1429.<br />

Kuroiwa, Y., Kasinathan, P., Choi, Y.J., Naeem, R., Tomizuka, K., Sullivan, E.J., Knott, J.G., Duteau, A.,<br />

Goldsby, R., Osborne, B.A., Ishida, I., Robl, J., 2002. Cloned transchromosomic calves producing human<br />

immunoglobulin. Nat. Biotechnol. 20, 889–894.<br />

Lai, L., Kolber-Simonds, D., Park, K.W., Cheong, H.T., Greenstein, J.L., Im, G.S., Samuel, M., Bonk,<br />

A., Rieke, A., Day, B.N., Murphy, C.N., Carter, D.B., Hawley, R.J., Prather, R.S., 2002. Production of<br />

alpha-1,3-galactosyltransferase knockout pigs by nuclear transfer cloning. Science 295, 1089–1092.<br />

Lavitrano, M., Bacci, M.L., Forni, M., Lazzereschi, D., Di Stefano, C., Fioretti, D., Giancotti, P., Marfe, G., Pucci,<br />

L., Renzi, L., Wang, H., Stoppacciaro, A., Stassi, G., Sargiacomo, M., Sinibaldi, P., Turchi, V., Giovannoni,<br />

R., Della Casa, G., Seren, E., Rossi, G., 2002. Efficient production by sperm-mediated gene transfer of human<br />

decay accelerating factor (hDAF) transgenic pigs for xenotransplantation. Proc. Natl. Acad. Sci. USA 99,<br />

14230–14235.<br />

Lemme, E., Eckert, J., Carnwath, J.W., Niemann, H., 1994. Expression of 6WTK-LacZ gene construct in in<br />

vitro produced bovine embryos following microinjection into pronuclei or cytoplasm. Theriogenology 41, 236<br />

(abstract).<br />

Martin, G.R., 1981. Isolation of a pluripotent cell line from early mouse embryos cultured in medium conditioned<br />

by teratocarcinoma stem cells. Proc. Natl. Acad. Sci. USA 78, 7634–7638.<br />

Martin, U., Tacke, S.J., Simon, A., Schröder, C., Wiebe, K., Lapin, B., Haverich, A., Denner, J., Steinhoff, G.,<br />

2002. Absence of PERV specific humoral immune response in baboons after transplantation of porcine cells<br />

of organs. Transplant. Int. 15, 361–368.<br />

Matsui, Y., Zsebo, K., Hogan, B.L., 1992. Derivation of pluripotential embryonic stem cells from murine primordial<br />

germ cells in culture. Cell 70, 841–847.<br />

Mayford, M., Mansuy, I.M., Muller, R.U., Kandel, E.R., 1997. Memory and behavior: a second generation of<br />

genetically modified mice. Curr. Biol. 7, 580–589.<br />

McCreath, K., Howcroft, J., Campbell, K.H.S., Colman, A., Schnieke, A., Kind, A., 2000. Production of<br />

gene-targeted sheep by nuclear transfer from cultured somatic cells. Nature 405, 1066–1069.


H. Niemann, W.A. Kues / Animal Reproduction Science 79 (2003) 291–317 315<br />

McPherron, A.C., Lee, S.J., 1997. Double muscling in cattle due to mutations in the myostatin gene. Proc. Natl.<br />

Acad. Sci. USA 94, 12457–12461.<br />

McPherron, A.C., Lawler, A.M., Lee, S.J., 1997. Regulation of skeletal muscle mass in mice by a new TGF-beta<br />

superfamily member. Nature 38, 83–90.<br />

Meade, H.M., Echelard, Y., Ziomek, C.A., Young, M.W., Harvey, M., Cole, E.S., Groet, S., Smith, T.E., Curling,<br />

J.M., 1999. Expression of recombinant proteins in the milk of transgenic animals. In: Fernandez, J.M., Hoeffler,<br />

J.P. (Eds.), Gene Expression Systems. Academic Press, San Diego, USA, pp. 399–427.<br />

Michler, R.E., 1996. Xenotransplantation: risks and prospects. Xeno 4, 21–26.<br />

Modrek, B., Lee, C., 2001. A genomic view of alternative splicing. Nat. Genet. 30, 13–19.<br />

Monaco, A.P., Larin, Z., 1994. YACs, BACs, PACs and MACs: artificial chromosomes as research tools. TIBTECH<br />

12, 280–286.<br />

Müller, M., Brem, G., 1991. Disease resistance in farm animals. Experientia 47, 923–934.<br />

Murray, J.D., 1999. Genetic modification of animals in the next century. Theriogenology 51, 149–159.<br />

Niemann, H., Kues, A.W., 2000. Transgenic livestock: premises and promises. Anim. Reprod. Sci. 60–61, 277–293.<br />

Niemann, H., Wrenzycki, C., 2000. Alterations of expression of developmentally important genes in<br />

preimplantation bovine embryos by in vitro culture conditions: implications for subsequent development.<br />

Theriogenology 54, 21–24.<br />

Niemann, H., Halter, R., Paul, D., 1994. Gene transfer in cattle and sheep: a summary perspective. In: Proceedings of<br />

the Fifth World Congress on Genetics Applied to Livestock Production, vol. 21, Guelph, Canada, pp. 339–346.<br />

Niemann, H., Halter, R., Carnwath, J.W., Herrmann, D., Lemme, E., Paul, D., 1999. Expression of human blood<br />

clotting factor VIII in the mammary gland of transgenic sheep. Transgenic Res. 8, 237–247.<br />

Niemann, H., Verhoeyen, E., Wonigeit, K., Lorenz, R., Hecker, J., Schwinzer, R., Hauser, H.J., Kues, W.A., Halter,<br />

R., Lemme, E., Herrmann, D., Winkler, M., Wirth, D., Paul, D., 2001. CMV early promoter induced expression<br />

of hCD59 in porcine organs provides protection against hyperacute rejection. Transplantation 72, 1898–1906.<br />

Niemann, H., Döpke, H.H., Hadeler, K.G., 2002a. Production of transgenic ruminants by DNA microinjection.<br />

In: Pinkert, C. (Ed.), Transgenic Animal Technology: A Laboratory Handbook, 2nd ed. Academic Press, San<br />

Diego, USA, pp. 337–357.<br />

Niemann, H., Wrenzycki, C., Lucas-Hahn, A., Brambrink, T., Kues, W.A., Carnwath, J.W., 2002b. Gene expression<br />

patterns in bovine in vitro produced and nuclear transfer derived embryos and their implication for early<br />

development. Cloning Stem Cells 4, 29–37.<br />

Nottle, M.B., Nagashima, H., Verma, P.J., Du, Z.T., Grupen, C.G., Ashman, R.J., Macllfatrick, S., 1997.<br />

Developments in transgenic techniques in pigs. J. Reprod. Fertil. Suppl. 52, 237–244.<br />

Nottle, M.B., Nagashima, H., Verma, P.J., Du, Z.T., Grupen, C.G., McIlfatrick, S.M., Ashman, R.J., Harding, M.P.,<br />

Giannakis, C., Wigley, P.L., Lyons, I.G., Harrison, D.T., Luxford, B.G., Campbell, R.G., Crawford, R.J., Robins,<br />

A.J., 1999. Production and analysis of transgenic pigs containing a metallothionein porcine growth hormone<br />

gene construct. In: Murray, J.D., Anderson, G.B., Oberbauer, A.M., McGloughlin, M.M. (Eds.), Transgenic<br />

Animals in Agriculture. CABI Publishing, New York, USA, pp. 145–156.<br />

O’Brien, S.J., Menotti-Raymond, M., Murphy, W.J., Nash, W.G., Wienberg, J., Stanyon, R., Copeland, N.G.,<br />

Jenkins, N.A., Womack, J.E., Graves, J.A.M., 1999. The promise of comparative genomics in mammals.<br />

Science 286, 458–481.<br />

Oldmixon, B.A., Wood, J.C., Ericcson, T.A., Wilson, C.A., White-Scharf, M.E., Andersson, G., Greenstein, J.L.,<br />

Schuurman, H.J., Patience, C., 2002. Porcine endogenous retrovirus transmission characteristics of an inbred<br />

herd of miniature swine. J. Virol. 76, 3045–3048.<br />

Orban, P.C., Chui, D., Marth, J.D., 1992. Tissue- and site-specific DNA recombination in transgenic mice. Proc.<br />

Natl. Acad. Sci. USA 89, 6861–6865.<br />

Paleyanda, R.K., Velander, W.H., Lee, T.K., Scandella, D.H., Gwazdauskas, F.C., Knight, J.W., Hoyer, L.W.,<br />

Drohan, W.H., Lubon, H., 1997. Transgenic pigs produce functional human factor VIII in milk. Nat. Biotechnol.<br />

15, 971–975.<br />

Paradis, K., Langford, G., Long, Z., Heneine, W., Sandstrom P., Switzer, W.M., Chapman, L.E., Lockey, C., Onions,<br />

D., The XEN 111 Study Group, Otto, E., 1999. Search for cross-species transmission of porcine endogenous<br />

retrovirus in patients treated with living pig tissue. Science 285, 1236–1241.<br />

Park, K.W., Cheong, H.T., Lai, L., Im, G.S., Kuhholzer, B., Bonk, A., Samuel, M., Rieke, A., Day, B.N., Murphy,<br />

C.N., Carter, D.B., Prather, R.S., 2001. Production of nuclear transfer-derived swine that express the enhanced<br />

green fluorescent protein. Anim. Biotechnol. 12, 173–181.


316 H. Niemann, W.A. Kues / Animal Reproduction Science 79 (2003) 291–317<br />

Patience, C., Takeuchi, Y., Weiss, R.A., 1997. Infection of human cells by an endogenous retrovirus of pigs. Nat.<br />

Med. 3, 282–286.<br />

Patience, C., Patton, G.S., Takeuchi, Y., Weiss, R.A., McClure, M.O., Rydberg, L., 1998. No evidence of pig<br />

DNA of retroviral infection in patients with short-term extracorporal connection of pig kidneys. Lancet 352,<br />

699–701.<br />

Perez, C., Jong de, G., Drayer, J., 2000. Satellite DNA-based artificial chromosomes—chromosomal vectors.<br />

TIBTECH 18, 402–403.<br />

Perry, A.C.F., Wakayama, T., Kishikawa, H., Kasai, T., Okabe, M., Toyoda, Y., Yanagimachi, R., 1999. Mammalian<br />

transgenesis by intracytoplasmic sperm injection. Science 284, 1180–1183.<br />

Perry, A.C., Rothman, A., de las Heras, J.I., Feinstein, P., Mombaerts, P., Cooke, H.J., Wakayama, T., 2001.<br />

Efficient metaphase II transgenesis with different transgene archetypes. Nat. Biotechnol. 19, 1071–1073.<br />

Petit-Zeman, S., 2001. Regenerative medicine. Nat. Biotechnol. 19, 201–206.<br />

Petters, R.M., Alexander, C.A., Wells, K.D., Collins, E.B., Sommer, J.R., Blanton, M.R., Rojas, G.H.Y., Flowers,<br />

W.L., Banin, E., Cideciyan, A.V., Jacobson, S.G., Wong, F., 1997. Genetically engineered large animal model for<br />

studying cone photoreceptor survival and degeneration in retinitis pigmentosa. Nat. Biotechnol. 15, 965–970.<br />

Piedrahita, J.A., Moore, K., Oetama, B., Lee, C.K., Scales, N., Ramsoondar, J., Bazer, F.W., Ott, T., 1998.<br />

Generation of transgenic porcine chimeras using primordial germ cell-derived colonies. Biol. Reprod. 58,<br />

1321–1329.<br />

Platt, J.L., Lin, S.S., 1998. The future promises of xenotransplantation. In: Fishman, J., Sachs, D., Shaikh, R.<br />

(Eds.), Xenotransplantation—Scientific Frontiers and Public Policy. Ann. New York Acad. Sci. 862, 5–18.<br />

Polejaeva, I.A., Chen, S.H., Vaught, T.D., Page, R.L., Mullins, J., Ball, S., Dal, Y., Boano, J., Walker, S., Ayares,<br />

D., Colman, A., Campbell, K.H.S., 2000. Cloned pigs produced by nuclear transfer from adult somatic cells.<br />

Nature 407, 505–509.<br />

Pursel, V.G., Rexroad Jr., C.E., 1993. Status of research with transgenic farm animals. J. Anim. Sci. 71, 10–19.<br />

Resnick, J.L., Bixler, L.S., Cheng, L., Donovan, P.J., 1992. Long-term proliferation of mouse primordial germ<br />

cells in culture. Nature 359, 550–551.<br />

Rexroad Jr., C.E., 1992. Transgenic technology in animal agriculture. Anim. Biotechnol. 3, 1–13.<br />

Rudolph, N.S., 1999. Biopharmaceutical production in transgenic livestock. TIBTECH 17, 367–374.<br />

Sauer, B., 1998. Inducible gene targeting in mice using the Cre/lox system. Methods 14, 381–392.<br />

Sauer, B., Henderson, N., 1988. Site-specific DNA recombination in mammalian cells by the Cre recombinase of<br />

bacteriophage P1. Proc. Natl. Acad. Sci. USA 85, 5166–5170.<br />

Schedl, A., Beermann, F., Thies, E., Montoliu, L., Kelsey, G., Schütz, G., 1992. Transgenic mice generated by<br />

pronuclear injection of a yeast artificial chromosome. Nucleic Acids Res. 20, 3073–3077.<br />

Schedl, A., Montoliu, L., Kelsey, G., Schütz, G., 1993. A yeast artificial chromosome covering the tyrosinase gene<br />

confers copy number-dependent expression in transgenic mice. Nature 362, 258–261.<br />

Schnieke, A.E., Kind, A.J., Ritchie, W.A., Mycock, K., Scott, A.R., Ritchie, M., Wilmut, I., Colman, A., Campbell,<br />

K.H.S., 1997. Human factor IX transgenic sheep produced by transfer of nuclei from transfected fetal fibroblasts.<br />

Science 278, 2130–2133.<br />

Schultze, N., Burki, Y., Lang, Y., Certa, U., Bluethmann, H., 1996. Efficient control of gene expression by single<br />

step integration of the tetracycline system in transgenic mice. Nat. Biotechnol. 14, 499–503.<br />

Seidel Jr., G.E., 1993. Resource requirements for transgenic livestock research. J. Anim. Sci. 71, 26–33.<br />

Shim, H., Gutierrez-Adan, A., Chen, L.R., BonDurant, R.H., Behboodi, E., Anderson, G.B., 1997. Isolation of<br />

pluripotent stem cells from cultured porcine primordial germ cells. Biol. Reprod. 57, 1089–1095.<br />

Squires, E.J., 1999. Status of sperm-mediated delivery methods for gene transfer. In: Murray, J.D., Anderson,<br />

G.B., Oberbauer, A.M., McGloughlin, M.M. (Eds.), Transgenic Animals in Agriculture. CABI Publishing,<br />

New York, USA, pp. 87–96.<br />

Strauss, W.M., Dausman, J., Beard, C., Johnson, C., Lawrence, J.B., Jaenisch, R., 1993. Germ line transmission<br />

of a yeast artificial chromosome spanning the murine 1 collagen locus. Science 259, 1904–1907.<br />

Switzer, W.M., Michler, R.E., Shanmugan, V., Matthews, A., Hussain, A.I., Wright, A., Sandstrom, P., Chapman,<br />

L., Weber, C., Safley, S., Denny, R.R., Navarro, A., Evans, V., Norein, A.J., Kwiatkowski, P., Heneine, W., 2001.<br />

Lack of cross-species transmission of porcine endogenous retrovirus infection to nonhuman primate recipients<br />

of porcine cells, tissues and organs. Transplantation 71, 959–965.<br />

Theuring, F., Thunecke, M., Kosciessa, U., Tuern, J.D., 1997. Transgenic animals as models of neurodegenerative<br />

disease in humans. TIBTECH 15, 320–325.


H. Niemann, W.A. Kues / Animal Reproduction Science 79 (2003) 291–317 317<br />

Thibier, M., 2001. The animal embryo transfer industry in figures. IETS-Newsletter 19, 16–22.<br />

Thomas, M., Northrup, R., Hornsby, P.J., 1997. Adrenocortical tissue formed by transplantation of normal clones<br />

of bovine adrenocortical cells in scid mice replaces the essential functions of the animals’ adrenal glands. Nat.<br />

Med. 3, 978–983.<br />

Tomizuka, K., Shinohara, T., Yoshida, H., Uekima, H., Ohguma, A., Tanaka, S., Sato, K., Oshimura, M., Ishida,<br />

I., 2000. Double trans-chromosomic mice: maintenance of two individual human chromosome fragments<br />

containing Ig heavy and kappa loci and expression of fully human antibodies. Proc. Natl. Acad. Sci. USA 97,<br />

722–727.<br />

van Berkel, P.H., Welling, M.M., Geerts, M., van Veen, H.A., Ravensbergen, B., Salaheddine, M., Pauwels, E.K.,<br />

Pieper, F., Nuijens, J.H., Nibbering, P.H., 2002. Large scale production of recombinant human lactoferrin in<br />

the milk of transgenic cows. Nat. Biotechnol. 20, 484–487.<br />

Vos, J.-M., 1997. The simplicity of complex MACs. Nat. Biotechnol. 15, 1257–1259.<br />

Wakayama, T., Perry, A.C.F., Zuccotti, M., Johnson, K.R., Yanagimachi, R., 1998. Full-term development of mice<br />

from enucleated oocytes injected with cumulus cell nuclei. Nature 394, 369–374.<br />

Wall, R.J., 1996. Transgenic livestock: progress and prospects for the future. Theriogenology 45, 57–68.<br />

Wall, R.J., Hawk, H.W., Nel, N., 1992. Making transgenic livestock: genetic engineering on a large scale. J. Cell.<br />

Biochem. 49, 113–120.<br />

Wheeler, M.B., 1994. Development and validation of swine embryonic stem cells: a review. Reprod. Fertil. Dev.<br />

6, 563–568.<br />

Wheeler, M.B., Bleck, G.T., Donovan, S.M., 2001. Transgenic alteration of sow milk to improve piglet growth<br />

and health. Reprod. Suppl. 58, 313–324.<br />

White, D., 1996a. Alteration of complement activity: a strategy for xenotransplantation. TIB 14, 3–5.<br />

White, D., 1996b. hDAF transgenic pig organs: are they concordant for human transplantation. Xeno 4, 50–54.<br />

Wilmut, I., Schnieke, A.E., McWhir, J., Kind, A.J., Campbell, K.H., 1997. Viable offspring derived from fetal and<br />

adult mammalian cells. Nature 385, 803–810.<br />

Wilmut, I., Beaujean, N., De Sousa, P.A., Dinnyes, A., King, T.J., Paterson, L.A., Wells, D.N., Young, L.E., 2002.<br />

Somatic cell nuclear transfer. Nature 419, 583–586.<br />

Yang, Y.W., Model, P., Heintz, N., 1997. Homologous recombination based modification in Escherichia coli and<br />

germline transmission in transgenic mice of a bacterial artificial chromosome. Nat. Biotechnol. 15, 859–865.<br />

Yarranton, G.T., 1992. Inducible vectors for expression in mammalian cells. Curr. Opin. Biotechnol. 3, 506–511.<br />

Yom, H.-C., Bremel, R.D., 1993. Genetic engineering of milk composition: modification of milk components in<br />

lactating transgenic animals. Am. J. Clin. Nutr. 58 (Suppl.), 299–306.<br />

Young, L.E., Sinclair, K.D., Wilmut, I., 1998. Large offspring syndrome in cattle and sheep. Rev. Reprod. 3,<br />

155–163.<br />

Zaidi, A., Schmoeckel, M., Bhatti, F., Waterworth, P., Tolan, M., Cozzi, E., Chavez, G., Langford, G., Thiru, S.,<br />

Wallwork, J., White, D., Friend, P., 1998. Life-supporting pig to primated renal xeno-transplantation using<br />

genetically modified donors. Transplantation 65, 1584–1590.<br />

Zawada, W.M., Cibelli, J.B., Chei, P.K., Clarkson, E.D., Golueke, P.J., Witta, S.E., Beil, K.P., Kane, J., Abel<br />

Ponce de Leon, F., Jerry, D.J., Robl, J.M., Freed, C.R., Stice, S.L., 1998. Somatic cell cloned transgenic bovine<br />

neurons for transplantation in Parkinson rats. Nat. Med. 4, 569–574.<br />

Ziomek, C.A., 1998. Commercialization of proteins produced in the mammary gland. Theriogenology 49, 139–<br />

144.


MOLECULAR REPRODUCTION AND DEVELOPMENT 65:245–253 (2003)<br />

Cell Cycle Dependent Expression of Plk1 in<br />

Synchronized Porcine Fetal Fibroblasts<br />

MARTIN ANGER, 1 WILFRIED A. KUES, 2 JIRI KLIMA, 1 MANFRED MIELENZ, 2 MICHAL KUBELKA, 1<br />

JAN MOTLIK, 1 MILAN ESNER, 3 PETR DVORAK, 3 JOSEPH W. CARNWATH, 2 AND HEINER NIEMANN 2 *<br />

1 Institute of Animal Physiology and Genetics, Libechov, Czech Republic<br />

2 Department of Biotechnology, Institute for Animal Science, Mariensee, Neustadt, Germany<br />

3 Laboratory of Molecular Embryology, Mendel University, Brno, Czech Republic<br />

ABSTRACT Enzymes of the Polo-like kinase<br />

(Plk) family are active in the pathways controlling<br />

mitosis in several species. We have cloned cDNA fragments<br />

of the porcine homologues of Plk1, Plk2, and<br />

Plk3 employing fetal fibroblasts as source. All three<br />

partial cDNAs showed high sequence homology with<br />

their mouse and human counterparts and contained the<br />

Polo box, a domain characteristic for all Polo kinases.<br />

The expression levels of Plk1 mRNA at various points<br />

of the cell cycle in synchronized porcine fetal fibroblasts<br />

were analyzed by both RT-PCR and the ribonuclease<br />

protection assay. Plk1 mRNA was barely<br />

detectable in G0 and G1, increased during S phase and<br />

peaked after the G2/M transition. A monoclonal antibody<br />

was generated against an in vitro expressed porcine<br />

Plk1-protein fragment and used to detect changes<br />

in Plk1 expression at the protein level. Plk1 protein<br />

was first detected by immunoblotting at the beginning<br />

of S phase and was highest after the G2/M transition.<br />

In summary, the Plk1 expression pattern in the pig is<br />

similar to that reported for other species. The absence<br />

of Plk1 mRNA and protein appears to be a good marker<br />

for G0/G1 and thus for the selection of donor cells for<br />

nuclear transfer based somatic cloning. Mol. Reprod.<br />

Dev. 65: 245–253, 2003. ß 2003 Wiley-Liss, Inc.<br />

Key Words: Plk1; serum deprivation; cell cycle;<br />

butyrolactonuI; aphidicolin<br />

INTRODUCTION<br />

Polo kinase was first identified in yeast and Drosophila<br />

(Llamazares et al., 1991). The homologous molecule,<br />

Polo-like kinase (Plk), was subsequently isolated<br />

from higher vertebrates including, mouse (Clay et al.,<br />

1993), human (Lake and Jelinek, 1993), and Xenopus<br />

(Kumagai and Dunphy, 1996). All members of the Polo<br />

kinase family contain two copies of a unique C-terminal<br />

sequence, the ‘‘Polo box,’’ thought to be responsible for<br />

the intracellular localization of the enzyme (Clay et al.,<br />

1997; Lee et al., 1998). Two additional kinases belonging<br />

to the Polo family but unique to vertebrates are called<br />

Plk2 and Plk3 (Kauselmann et al., 1999; Holtrich et al.,<br />

2000; Hudson et al., 2001; Nigg, 2001). Since there are no<br />

homologues of Plk2 and Plk3 in lower eukaryotes,<br />

ß 2003 WILEY-LISS, INC.<br />

detailed knowledge about the role of the Polo kinases<br />

relates to Plk1.<br />

Plk1 is abundant in proliferating cells and tissues<br />

including fetal tissues and certain types of human<br />

cancer (Yuan et al., 1997; Wolf et al., 1997, 2000). In<br />

cultured mouse cells, Plk1 mRNA is first detectable at<br />

G1/S transition, with a peak at G2/M transition (Lake<br />

and Jelinek, 1993; Uchiumi et al., 1997). Several studies<br />

have demonstrated that Plk1 is critically involved in cell<br />

cycle related activities. The kinase activity of Plk1 is<br />

highest at G2/M and when it is phosphorylated during M<br />

phase and timing of this activation is coincident with<br />

activation of MPF (Golsteyn et al., 1995; Hamanaka<br />

et al., 1995; Mundt et al., 1997). Indeed, it has been suggested<br />

that activated Plk1 is involved in activation of<br />

MPF and thus on mitotic entry (Roshak et al., 2000).<br />

Plk1 activity regulates maturation of the spindle; is<br />

involved in checkpoint monitoring of the integrity of<br />

replicated DNA (Smits et al., 2000), and regulates<br />

substrate specificity of the anaphase-promoting complex<br />

(APC) (reviewed by Glover et al., 1996, 1998;<br />

Golsteyn et al., 1996; Nigg, 1998, 2001). In late mitosis,<br />

Plk1 appears to be involved in the control of cyclin B<br />

proteolysis, thus initiating processes resulting in the<br />

exit from mitosis (Abrieu et al., 1998; Descombes and<br />

Nigg, 1998; Ferris et al., 1998; Kotani et al., 1998; Qian<br />

et al., 1998; 1999; Karaiskou et al., 1999).<br />

In the present study, we have cloned the porcine Plks<br />

1, 2, and 3 and compared the mRNA expression levels of<br />

these genes in synchronized murine and porcine fetal<br />

fibroblasts throughout the cell cycle. In addition, a<br />

monoclonal antibody against Plk1 was generated and<br />

characterized. Expression of Plks in porcine cells had<br />

not been reported previously. The data indicate that<br />

Plk1 mRNA shows the most dramatic changes during<br />

Grant sponsor: FIRCA; Grant number: R03-TW-05530-01; Grant<br />

sponsor: German Ministry for Education and Research (BMBF);<br />

Grant sponsor: Ministry of the Education Youth and Sport of the<br />

Czech Republic; Grant number: LN00A065.<br />

*Correspondence to: Prof., Dr. Heiner Niemann, Department of<br />

Biotechnology, Institute for Animal Science, Mariensee, D-31535<br />

Neustadt, Germany. E-mail: niemann@tzv.fal.de<br />

Received 19 August 2002; Accepted 13 January 2003<br />

Published online in Wiley InterScience (www.interscience.wiley.com).<br />

DOI 10.1002/mrd.10289


246 M. ANGER ET AL.<br />

cell cycle in porcine fetal fibroblasts. We further tested<br />

the hypothesis that absence of Plk1 mRNA and protein<br />

in nonproliferating cells and tissues (Kues et al., 2000),<br />

i.e., cells in the G0 and G1 phase of the cell cycle,<br />

provides a suitable marker to identify populations of<br />

growth-arrested cells. To do this, expression levels of<br />

Plk1 mRNA and protein were measured after fetal<br />

calf serum (FCS) stimulation of serum deprived cells<br />

for defined time intervals as well as in cells blocked<br />

by either aphidicolin or aphidocolin and butyrolactone I<br />

treatment. Determining Plk1 expression to define the<br />

cell cycle stage is of great interest because these cells are<br />

frequently employed as nucleus donor cells in somatic<br />

cloning experiments (Onishi et al., 2000; Polejaeva et al.,<br />

2000) and it had been suggested that it is beneficial to<br />

block donor cells in G0 by serum deprivation (Wilmut<br />

and Campbell, 1998; Lee and Piedrahita, 2002).<br />

MATERIALS AND METHODS<br />

Cell Culture, Cell-Cycle Synchronization,<br />

and FACS Analysis<br />

STO and HeLa cells were purchased from ATTCC<br />

(American Type Culture Collection, Manassas VA),<br />

bovine skin fibroblasts were a kind gift from Dr. Jiri<br />

Kanka, mouse fetal fibroblasts were prepared according<br />

to standard protocols (Hogan et al., 1994). Cells were<br />

thawed and grown to 80% confluency in Dulbecco’s<br />

Modified Eagle Medium (DMEM) plus 10% FCS (Sigma<br />

Chemical Co., St. Louis, MO). Mouse fetal fibroblasts<br />

were synchronized in G0/G1 by reduction of serum<br />

content to 0.2% for 3 days followed by stimulation with<br />

10% FCS 1, 11, 16, and 21 hr before harvesting. Porcine<br />

primary fibroblasts were prepared as recently described<br />

(Kues et al., 2000). All experiments were performed from<br />

the same batch of fetal fibroblasts harvested after the<br />

2nd passage, frozen and stored in liquid nitrogen until<br />

use. For each series of experiments, fibroblasts were<br />

thawed and grown to 80% confluency in DMEM plus<br />

10% FCS (Sigma Chemical Co.). To obtain cell-cycle<br />

expression profiles for Plk1, Plk2, and Plk3 mRNAs,<br />

cells were synchronized by deprivation of FCS (0.2%) in<br />

culture media for 3 days followed by stimulation with<br />

10% FCS for 1, 4, 18, 21 hr. Nocodazole (Sigma Chemical<br />

Co.) was used to obtain M phase synchronized cells.<br />

Therefore, cells stimulated for 16 hr with 10% FCS were<br />

synchronized by 16.5 mM Nocodazole for 8 hr. The following<br />

synchronization protocols were used to measure<br />

Plk1 expression in porcine fetal fibroblasts:<br />

1. G0/G1 cells—obtained by reduction of FCS in the<br />

culture media to 0.2% for 48 hr;<br />

2. G1/S cells—obtained by stimulation of G0/G1 cells by<br />

addition of fresh medium containing 10% FCS and<br />

blocking at the beginning of S phase with 6 mM<br />

aphidicolin (Sigma Chemical Co.), a DNA polymerase<br />

inhibitor, for 14 hr;<br />

3. G2/M cells—produced by washing G1/S cells with<br />

PBS (to remove aphidicolin), followed by stimulation<br />

with medium containing 10% FCS for 3 hr and finally<br />

adding medium containing 10% FCS and 118 mM<br />

butyrolactone I (Affinity, Exeter, UK), a cyclin dependent<br />

kinase inhibitor, for 5 hr.<br />

4. Partially synchronized cells—obtained by stimulating<br />

G0/G1 cells with 10% FCS for 14 or 23 hr,<br />

respectively.<br />

DNA content was measured by FACS to confirm the<br />

cell cycle stage and the efficiency of synchronization as<br />

previously described (Kues et al., 2000).<br />

Cloning of cDNA of Plks<br />

Total RNA was prepared from fibroblasts with the<br />

RNeasy Mini kit (Qiagen, Hilden, Germany) and 1 mgof<br />

the total RNA was used as template for a reverse<br />

transcription (RT) reaction. The RT reaction mix (40 ml<br />

total volume) contained: 1 PCR buffer II (Perkin<br />

Elmer, Billerica, MA); 5 mM MgCl 2 (Gibco-BRL,<br />

Gaithersburg, MD); 1 mM dNTPs (USB Corporation,<br />

Cleveland, OH); RNase inhibitor 20 U (Perkin Elmer);<br />

MuLV RT 2.5 ml (Perkin Elmer); 2.5 mM oligo (dT)12–18<br />

(Gibco-BRL). RT was carried out for 1 hr at 428C with<br />

final denaturing for 5 min at 758C. An aliquot of 5 ml of<br />

the RT reaction was used as a template for a 100 ml<br />

PCR reaction. The primers (Table 1) were based on<br />

regions of consensus between the human and mouse<br />

Plk’s and modified by adding restriction sites for BamHI<br />

and XhoI. The PCR reaction mixture (100 ml total<br />

volume) contained: 1 Pfu polymerase reaction buffer<br />

(Stratagene, La Jolla, CA); 0.2 mM dNTPs (USB<br />

Corporation), cloned Pfu DNA polymerase (Stratagene)<br />

2.5 U; primers 0.5 mM each (MWG Biotech AG,<br />

Ebersberg, Germany). PCR temperature cycle: 948C<br />

for 45 sec, 608C for 45 sec, 728C for 3 min, repeated for 29<br />

cycles, final extension 728C for 10 min. PCR products<br />

were gel purified and sub-cloned into the pGEM 1 -4Z<br />

vector (Promega GmbH, Mannheim, Germany) by using<br />

TA cloning strategy pGEM-T Easy kit (Promega GmbH)<br />

and sequenced.<br />

Semiquantitative RT-PCR<br />

Procedures for RNA isolation and the RT reaction<br />

were as described above for cloning except that priming<br />

of the RT reaction was performed with random hexamers.<br />

Relative changes in the abundance of Plk transcripts<br />

were determined by a multiplex RT-PCR assay,<br />

using either b-actin or GAPDH mRNA as internal<br />

standards. Primers used for the quantification are listed<br />

in Table 1. The PCR reaction mixture (20 ml) contained:<br />

5 ml of the RT reaction as a template; 1 PCR buffer<br />

(Gibco-BRL); 0.2 mM each dNTP (USB Corporation);<br />

1.5 mM MgCl2 (Gibco-BRL); PCR primers 0.5 mM each<br />

(MWG Biotech AG); Taq DNA polymerase 0.5 U (Gibco-<br />

BRL). PCR temperature profile: initial denaturation at<br />

948C for 1 min; then cycles of 948C for 15 sec; 608C for<br />

30 sec; 728C for 1 min; and final elongation 728C for<br />

5 min. Because of the abundance of GAPDH transcript,<br />

the PCR mixture initially contained only Plk primers<br />

and the GAPDH primers were added after the 5th PCRcycle<br />

during a programmed hold at 728C. The total<br />

number of PCR-cycles was 27 for Plk1 and 22 for


GAPDH. The number of cycles was optimized separately<br />

for each gene during preliminary experiments to obtain<br />

linear amplification of the template. Primers used<br />

separately gave the same amount of product as in the<br />

multiplex reaction. The control reactions for each PCR<br />

included one tube without the cDNA template and<br />

one tube with RNA that had not been exposed to RT.<br />

The PCR products were separated on 1.5% agarose<br />

gels (Gibco-BRL); stained with ethidium bromide; and<br />

recorded using a Photometrics CCD camera. Densitometric<br />

evaluation was performed using Image Master<br />

1D Elite V3 (Amersham Pharmacia, Piscataway,<br />

New Jersey).<br />

mRNA Quantification by the<br />

Ribonuclease Protection Assay<br />

Linearized plasmid pGEM-4Z containing Plk1 or<br />

GAPDH inserts was used as a template for sense and<br />

antisense RNA probe synthesis with a MAXIscript TM<br />

T3/T7 Kit (Ambion, Austin, TX). SP6 polymerase was<br />

purchased from New England Biolabs (New England<br />

Biolabs GmbH, Frankfurt am, Main, Germany). Probes<br />

were labeled with [ 32 P]UTP (800 Ci/mmol, 20 mCi/ml)<br />

(Amersham Pharmacia Biotech, Freiburg, Germany).<br />

The length of protected fragments was 224 bp for<br />

GAPDH and 458 bp for Plk1. After in vitro transcription,<br />

full-length transcripts were purified from a polyacrylamide<br />

gel. For each reaction, 1 mg of total RNA was used<br />

for solution hybridization (RPA II kit, Ambion) with the<br />

molar access of the probe determined for each probe<br />

separately in preliminary experiments. Solution hybridization<br />

was carried out at 428C overnight, subsequently<br />

RNA/RNA hybrids were treated by RNase T1 and<br />

protected fragments were separated on a 5% polyacrylamide<br />

gel. The labeled fragments were detected by<br />

autoradiography on a Kodak Biomax ML film (Sigma<br />

Chemical Co.). Developed films were recorded with a<br />

CCD camera and densitometrically analyzed with<br />

Image Master 1D Elite V3 (Amersham Pharmacia).<br />

TABLE 1. Sequences of Used Primers<br />

Primers Product size (bp)<br />

A (for cloning)<br />

Plk1 sense GGG ATC CGG CGA AAG AGA TCC CGG AGG TCC TAG T<br />

Plk1 antisense CCT CGA GTT ACG GCT GGC CAG CTC CTT GCA GCA<br />

Plk2 sense GGG ATC CAT GTG GAA CCC CAA ATT ATC TTT<br />

Plk2 antisense GCT CGA GTC AGC ACC GTC ACT TGA CTG ATG<br />

Plk3 sense GGG ATC CGC ATC AAG CAG GTT CAT TAC G<br />

Plk3 antisense GCT CGA GTG GGA AGC GAG GTA AGT ACA<br />

B (for quantification)<br />

Plk1 sense CTC CTG GAG CTG CAC AAG AGG AGG AA<br />

Plk1 antisense TCT GTC TGA AGC ATC TTC TGG ATG AG 455<br />

Plk2 sense CCG GAC AGA CTC TCT TCC AGC TGT T<br />

Plk2 antisense GGT ACC CAA AGC CAT ATT TGT TGG A 538<br />

Plk3 sense TGG AGA TGG GTT TGA AGA AGG T<br />

Plk3 antisense TGG GGT TGT AGT GCA CCG TC 274<br />

GAPDH sense GTT CCA GTA TGA TTC CAC CCA CGG CAA GTT<br />

GAPDH antisense TGC CAG CCC CAG CAT CAA AGG TAG AAG AGT 763<br />

b-Actin sense GTC GAC AAC GGC TCC GGC A<br />

b-Actin antisense GTC AGG TCC CGG CCA GCC A 529<br />

Plk: Polo-like kinase.<br />

IDENTIFICATION OF POLO-LIKE KINASES IN PORCINE FIBROBLASTS 247<br />

Production of Recombinant Plk1, Antibody<br />

Generation, and Western Blotting<br />

A 344 bp fragment of porcine Plk1 cDNA corresponding<br />

to bases 730–1,074 of GeneBank sequence<br />

AF339021 was subcloned into the pET30b (Novagen,<br />

Madison, WI) bacterial expression vector. The recombinant<br />

plasmid was introduced into competent E. coli<br />

(BL21[DE3] LysS, Novagen). A single colony was selected<br />

and inoculated into LB media containing kanamycin<br />

and chloramphenicol (Sigma Chemical Co.). The bacteria<br />

cells were cultured at 378C for 3 hr, then IPTG<br />

(Sigma Chemical Co.) was added to a final concentration<br />

of 0.3 mM. The bacteria were then cultured for 5 hr at<br />

258C and finally, were harvested by centrifugation. The<br />

recombinant protein was isolated by His Bind TM and<br />

S Tag TM purification kits (Novagen).<br />

For the production of monoclonal antibodies a standard<br />

protocol was used (Harlow and Lane, 1988).<br />

Briefly, female Balb/c mice were repeatedly immunized<br />

with 30 mg purified soluble recombinant protein. Spleen<br />

cells were fused with SP 2/0 mouse myeloma cells in 50%<br />

PEG (Sigma Chemical Co.) and the resultant hybridomas<br />

were selected and screened by ELISA and Western<br />

blotting using the porcine Plk1 fusion protein and<br />

porcine fetal fibroblast lysate. The supernatant from<br />

clone PL-12 was used in this study.<br />

Proteins were separated on 10% SDS–PAGE gels<br />

and blotted onto PVDF membranes (Amersham) following<br />

the protocol of Harlow and Lane (1988). Membranes<br />

were blocked with 5% FCS dissolved in 20 mM Tris-<br />

HCl, pH 7.6, 0.1% Tween-20, 137 mM sodium chloride<br />

(TTBS) (all from Sigma Chemical Co.). Plk1 was detected<br />

with 1:50 diluted hybridoma cell culture supernatant.<br />

Goat anti-mouse antibodies coupled to horseradish<br />

peroxidase (HRP) (Sigma Chemical Co.) were used as<br />

secondary antibody. HRP activity was visualized by<br />

chemiluminescence using the ECL Plus Western blotting<br />

system (Amersham Pharmacia Biotech) and detected by<br />

exposure to Kodak BioMax Light-1 film.


248 M. ANGER ET AL.<br />

Statistical Analysis<br />

The t-test was used to evaluate the observed differences<br />

between groups. P values lower than 0.05 were<br />

considered to be significant. All statistical analysis was<br />

performed using KyPlot software.<br />

RESULTS<br />

Cloning of Porcine Plks 1, 2, and 3<br />

Three partial Plk sequences, representing the porcine<br />

homologous cDNAs of Plk1, Plk2, and Plk3, were isolated<br />

from fetal fibroblasts by a RT-PCR approach.<br />

Using BLAST (Tatusova and Madden, 1999), the<br />

amino acid sequence of the cloned 1,579 bp fragment of<br />

porcine Plk1 (GeneBank Acc. No. AF339021) was found<br />

to be highly conserved to the published Plk1 sequences<br />

of other eukaryotes (Fig. 1). The deduced amino acid<br />

sequence of this fragment is 96% identical with human<br />

Plk1 (AAA36659), 95% with mouse Plk1 (AAA16071),<br />

82% with Xenopus Plx1 (AAC60017), and 51% with<br />

Drosophila melanogaster Polo (NP_524179). BLAST<br />

also identified two characteristic Plk domains employing<br />

a search of the Conserved Domain Database<br />

(Altschul et al., 1997). The first one is an N-terminal<br />

serine/threonine protein kinase catalytic domain consisting<br />

of the 252 amino acids located between positions<br />

8 and 260 of the amino acid sequence essential for kinase<br />

activity. The amino acid homology in this domain in<br />

comparison with the corresponding serine/threonine<br />

protein kinase catalytic domains of human, mouse,<br />

Xenopus, and Drosophila Plk1 is 99, 98, 86, and 65%,<br />

respectively, suggesting that it is slightly more conserved<br />

than the rest of the sequence. The second<br />

conserved domain is the C terminal Polo box distinguishing<br />

the Polo kinases from other serine/threonine<br />

protein kinases. Two Polo box sequences were located.<br />

The first occurred between amino acids 372–435 and<br />

the second was found between amino acids 470–523 of<br />

porcine Plk1. Figure 1 shows the alignment of Plk1<br />

kinases from porcine, human, mouse, Xenopus, and<br />

Drosophila species; the serine/threonine kinase domain<br />

and the Polo boxes are underlined.<br />

The cloned fragment of porcine Plk2 is 952 bp long<br />

(GeneBank Acc. No. AF348424) and the deduced amino<br />

acid sequence had 316 amino acid residues. Among<br />

vertebrates, the porcine Plk2 amino acid sequence had<br />

99% identity with the homologous human sequence<br />

(XP_041712), 97% with the mouse sequence (P53351),<br />

and 78% with the Xenopus sequence (AAL30175). No<br />

homologues were identified in nonvertebrate organisms.<br />

The porcine Plk2 fragment contained a portion of<br />

the serine/threonine kinase domain and a Polo box<br />

identified by a search against the Conserved Domain<br />

Database (Altschul et al., 1997).<br />

The cloned fragment of Plk3 (GeneBank Acc.<br />

No. AF348425) was 1,007 bp long and its deduced<br />

amino acid sequence possessed 335 amino acids. The<br />

porcine Plk3 fragment had 92, 86, 53% identical amino<br />

acids with corresponding human (XP_046526), mouse<br />

(AAC52191), and Xenopus (AAL30177) sequences. In<br />

the region corresponding to the fragment sequenced<br />

in this study, we found some regions of discontinuity<br />

with those vertebrate homologues, which may affect its<br />

function. For example, the murine sequence of Plk3<br />

contained gaps of 17 amino acids when compared to<br />

the porcine and human sequences. Like the other porcine<br />

Plks, the porcine Plk3 contained a serine/threonine<br />

kinase domain and the Polo box.<br />

Cell Cycle Stage Specific Expression<br />

of Plk1, Plk2, and Plk3<br />

Initially, we analyzed the expression levels of Plk1,<br />

Plk2, and Plk3 in murine and porcine fetal fibroblasts<br />

(Figure 2A,B) during the cell cycle. Serum deprived cell<br />

cultures were mitotically stimulated by addition of<br />

10% FCS and the relative abundance of Plk transcripts<br />

was determined. For each experiment a FACS analysis<br />

was performed (Table 2). Plk2 and Plk3 mRNA levels<br />

peaked rapidly within 1 hr of serum stimulation, whereas<br />

the highest levels of Plk1 mRNA were found after<br />

18–21 hr of stimulation. In porcine cell cultures, Plk1<br />

mRNA showed the highest changes in expression levels<br />

throughout the cell cycle. We further analyzed the<br />

changes in the expression levels of Plk1 with our optimized<br />

synchronization protocols to compare cells synchronized<br />

by serum stimulation with cells synchronized<br />

by specific inhibitors of cell cycle progression. The mRNA<br />

levels of Plk1 were determined by RT-PCR analysis at<br />

five points in the cell cycle (Fig. 3A). Samples were<br />

obtained by serum deprivation of porcine fetal fibroblasts<br />

(0.2% FCS) for 2 days to block the cells in G0 and then<br />

releasing the block by the addition of 10% FCS. Cells<br />

were harvested at G0/G1 (without serum stimulation)<br />

and at 14 and 23 hr following serum stimulation. The<br />

remaining two time points were obtained by releasing<br />

the fibroblasts from G0 and blocking them at G1/S or G2/<br />

M transition using aphidicolin and butyrolactone I (Kues<br />

et al., 2000). The efficiencies of the cell cycle synchronisations<br />

were measured by FACS analysis (Table 3). The<br />

lowest levels of Plk1 mRNA were observed in G0 fibroblasts<br />

(Fig. 3A). After serum stimulation, a significant<br />

increase in Plk1 mRNA was evident at 14, and the<br />

highest amount was detected at 23 hr (Fig. 3A). Expression<br />

levels of Plk1 mRNA in cells reentering the cell<br />

cycle following serum stimulation were compared with<br />

fibroblasts at G1/S and at G2/M transition using<br />

RT-PCR. These results were confirmed by RPA (Fig. 4).<br />

The amount of Plk1 mRNA was significantly higher<br />

(P 0.01) in cells entering S phase (stimulated with 10%<br />

FCS for 14 hr) than in cells blocked at G1/S transition by<br />

growth in 10% FCS followed by aphidicolin treatment<br />

for 14 hr. Significantly higher Plk1 mRNA was found in<br />

cells stimulated with FCS for 23 hr compared with cells<br />

blocked at the beginning of S phase by aphidicolin or<br />

stimulated for only 14 hr by FCS (P 0.05)<br />

Expression of Plk1 Protein in<br />

Course of Cell Cycle<br />

A monoclonal antibody against porcine Plk1 was generated,<br />

to determine, whether the cell cycle dependent


IDENTIFICATION OF POLO-LIKE KINASES IN PORCINE FIBROBLASTS 249<br />

Fig. 1. Alignment of Polo-like kinase 1 (Plk1) sequences. The<br />

deduced amino acid sequence of porcine Plk1 (AAK28550) aligned<br />

with the homologous sequences of man (AAA36659), mouse<br />

(AAA16071), Xenopus (AAC60017), and Drosophila (NP_524179).<br />

transcript levels of Plk1 reflect different protein levels.<br />

A recombinant Plk1 fragment (bases 730–1,074), that<br />

showed the greatest divergence compared to porcine<br />

Plk2 and Plk3 was used to immunize mice. The<br />

The conserved residues are in black columns. The sequences of<br />

the serine/threonine kinase catalytic domain (residues 8–260 of the<br />

porcine sequence) and the Polo boxes (residues 372–435 and 470–523<br />

of the porcine sequence) are underlined.<br />

monoclonal antibody PL-12 was found to be specific for<br />

Plk1 and identified a single band of apparently 68 kDa<br />

on Western blots from porcine fetal fibroblasts (Fig. 3B).<br />

PL-12 antibody did not show reactivity with Plk2 or Plk3


250 M. ANGER ET AL.<br />

Fig. 2. Expression profiles of Plks in murine and porcine fetal<br />

fibroblasts. A: Summarized Plk1, Plk2, and Plk3 mRNA expression<br />

levels from three independent experiments with mouse fetal fibroblasts.<br />

The black bars represent Plk1, gray bars represent Plk2 and<br />

white bars represent Plk3. The error bar represents the SD value for<br />

each experiment. B: Plk1, Plk2, and Plk3 mRNA expression levels from<br />

a single experiment with porcine fetal fibroblasts. Semiquantitative<br />

RT-PCR analysis was repeated three times for this experiment with<br />

similar results.<br />

(data not shown). Western blot analysis using the<br />

antibody PL-12 showed that the cell cycle stage dependent<br />

translation of the protein (Fig. 3B) followed the<br />

same pattern as RT-PCR had revealed for transcription,<br />

i.e., Plk1 was first detected at G1/S transition, increased<br />

in cells entering S phase (after 14 hr of FCS stimulation),<br />

and reached its highest level after G2/M transition when<br />

cells were already in mitosis (after 23 hr of FCS stimulation).<br />

We also tested the reactivity of the antibody with<br />

cell lysates from different species and found that it<br />

crossreacts with Plk1 from mouse STO fibroblasts,<br />

human HeLa cells and bovine skin fibroblasts (Fig. 3C).<br />

DISCUSSION<br />

Plks appear to be crucially involved in cell cycle<br />

regulation in several eukaryotic species (Nigg, 1998). In<br />

the present study, we cloned and sequenced cDNA<br />

fragments of porcine Plk1, Plk2, and Plk3. While our RT-<br />

PCR cloning strategy did not allow to clone full-length<br />

cDNA, 1,579 bp of sequence were obtained for porcine<br />

Plk1; 952 bp for porcine Plk2; and 1,007 bp for porcine<br />

Plk3. All three showed a high degree of sequence<br />

homology with Plks from other species with the closest<br />

similarity to their corresponding human Plks. We identified<br />

two domains shared by all three cloned porcine<br />

Plks, a catalytic domain typical for serine/threonine<br />

kinases, and the Polo box domain, which is the hallmark<br />

of all Polo kinases (Clay et al., 1997; Lee et al., 1998). The<br />

highest conservation between pig, human, mouse, frog,<br />

and fruit fly Plks was found in the catalytic and Polo box<br />

domains.<br />

Initial experiments on the cell cycle dependent expression<br />

of Plk genes in mouse and porcine fetal<br />

fibroblasts showed that the regulation of transcription<br />

differs dramatically between Plk genes. Plk2 and Plk3<br />

mRNA levels peaked rapidly within 1 hr of serum<br />

stimulation of serum deprived cells, whereas the highest<br />

Plk1 levels were found after 18–21 hr. In porcine fetal<br />

fibroblasts, the greatest changes during the course of<br />

TABLE 2. Effects of Serum Deprivation and Serum Stimulation on<br />

the Percentages ( SD) of Cells in the Different Cell Cycle Stages<br />

G0/G1 phase cells S phase cells G2/M phase cells<br />

A. Mouse fetal fibroblasts a<br />

3 days starved 86.3 6.1 4.7 2.3 9.0 3.7<br />

1 hr stimulated 88.2 6.5 5.3 2.9 6.5 3.8<br />

11 hr stimulated 76.5 5.4 12.9 2.0 10.2 6.0<br />

16 hr stimulated 53.2 7.1 35.4 1.5 11.4 2.3<br />

18 hr stimulated 45.4 5.1 31.2 1.8 23.4 1.7<br />

21 hr stimulated 44.9 4.7 21.0 3.6 34.1 1.6<br />

B. Porcine fetal fibroblasts b<br />

3 days starved 83.62 5.5 10.88<br />

1 hr stimulated 85.63 4.88 9.49<br />

4 hr stimulated 81.69 6.82 11.49<br />

18 hr stimulated 67.28 21.88 10.85<br />

21 hr stimulated 38.11 50.43 11.46<br />

Nocodazole 31.60 31.59 31.60<br />

a Summarized FACS data of three independent experiments.<br />

b FACS data of a single experiment.


IDENTIFICATION OF POLO-LIKE KINASES IN PORCINE FIBROBLASTS 251<br />

Fig. 3. Plk1 expression in synchronized porcine fibroblasts. A:<br />

mRNA expression in serum deprived (G0/G1), fetal calf serum (FCS)<br />

stimulated (14 and 23 hr) and cells synchronized by aphidicolin (G1/S)<br />

or butyrolactone I (G2/M). The gel picture shows representative results<br />

of the semiquantitative multiplex RT-PCR. B: Plk1 protein expression<br />

cell cycle were found for Plk1. To characterize Plk1<br />

expression in more details, optimized cell cycle synchronization<br />

protocols were employed. Under normal culture<br />

conditions, 66% of the cells had a DNA content characteristic<br />

of G0/G1 and a complete cell cycle lasted<br />

approximately 22 hr. Serum starvation increased the<br />

proportion of cells in G0/G1 to 80%. The aphidicolin<br />

block protocol produced a population in which 78% of<br />

cells were blocked at G1/S transition, <strong>bei</strong>ng unable<br />

to initiate DNA synthesis. FCS stimulation for 14 hr<br />

after serum starvation resulted in a population of cells<br />

in which 49% had DNA content characteristic of G1<br />

and 38% had entered S phase. The protocol with<br />

aphidicolin followed by butyrolactone I produced a<br />

population in which 41% of the cells were blocked at<br />

the G2/M transition and 38% were in G1. FCS stimulation<br />

of starved cells for 23 hr produced a population in<br />

which 81% of the cells were either in late M phase or<br />

early G1. We previously reported that porcine fetal<br />

fibroblasts released from blocking with aphidicolin<br />

and/or butyrolactone I show no residual effect on<br />

in synchronized porcine fetal fibroblasts. The Plk-1 protein was<br />

detected by Western blotting with monoclonal antibody PL-12. Cells<br />

were treated as was indicated above at RT-PCR analysis. C: Crossreactivity<br />

of PL-12 antibody with Plk1 from mouse STO cells, human<br />

HeLa cells and bovine skin fibroblasts (BF).<br />

proliferation rates or on Plk1 mRNA expression (Kues<br />

et al., 2000).<br />

The human Plk1 promoter contains domains similar<br />

to those found in the promoters of other cell cycle<br />

regulated proteins, including cyclin A and cyclin B<br />

which are known to be activated at G1/S transition<br />

(Uchiumi et al., 1997). This is consistent with the<br />

observation that cells blocked by aphidicolin at the<br />

G1/S transition had considerably more Plk1 mRNA than<br />

the starved cells in G0/G1. Our data for porcine fetal<br />

fibroblasts are also consistent with those for cultured<br />

mouse and human cells (Lake and Jelinek, 1993;<br />

Golsteyn et al., 1994) and with results (Uchiumi et al.,<br />

1997) in which various Plk1 promoter—luciferase transgene<br />

constructs were employed in HeLa cells synchronized<br />

by a double thymidine block. Cells initiated Plk1<br />

transcription at the G1/S transition. These data from<br />

human cells coincide with the expression profile of Plk1<br />

in porcine fetal fibroblasts. Interestingly, cells released<br />

from the aphidicolin block were able to complete S phase<br />

in a significantly shorter time period than starved cells<br />

TABLE 3. Cell Cycle Synchronization of Porcine Fetal Fibroblasts ( SD)<br />

Treatment of cells G0/G1 phase cells S phase cells G2/M phase cells<br />

Unsynchronized 24 hr 66.0 6.0 a<br />

G0/G1 80.3 3.5 b<br />

G1/S 77.7 5.1 a,b<br />

14 hr FCS 49.0 8.7 c,d<br />

G2/M 38.3 2.1 d<br />

23 hr FCS 46.3 4.2 c<br />

15.3 3.5 a<br />

18.3 3.1 a<br />

6.0 3.0 b<br />

13.7 5.5 a<br />

6.0 1.0 b<br />

16.3 4.2 a<br />

38.0 9.8 13.0 1.7 a<br />

20.3 1.2 a,c<br />

41.3 3.2 b<br />

19.3 3.2 a,c<br />

34.3 4.9 b<br />

Summarized FACS data of three independent experiments. The variability between groups<br />

was tested and the data within columns, which are significantly different are marked by<br />

different superscripts. FCS, fetal calf serum.


252 M. ANGER ET AL.<br />

Fig. 4. Comparative quantification of Plk1 transcripts employing<br />

RT-PCR and RPA. Semiquantitative multiplex RT-PCR analysis was<br />

repeated three times for each of three experiments and results are<br />

summarized in gray bars. The black bars represent the three RPA<br />

measurements of RNA obtained from single experiment. The RT-PCR<br />

expression levels in all groups differ significantly (P 0.05) from each<br />

other with the exception of groups 14 hr FCS and G2/M cells. In RPA<br />

analysis, all groups were significantly different (P 0.05), except<br />

groups 23 hr FCS and G2/M cells.<br />

stimulated by FCS (unpublished observation). This<br />

suggests that processes unrelated to DNA replication<br />

itself are rate limiting in this phase of the cell cycle<br />

and that they are able to proceed in the presence of<br />

aphidicolin that inhibits the primer elongation step of<br />

DNA replication.<br />

Apparently Plk1 expression is a marker for dividing<br />

cells. The protein was undetectable and the mRNA was<br />

only barely detectable in serum deprived cells arrested<br />

in G0/G1. The apparent background of S and G2/M<br />

phase cells in serum deprived cultures seems to reflect<br />

the fact that early passages of primary cells were used. It<br />

is well described that initially primary cultures consist<br />

of subpopulations that differ substantially in their<br />

potential to divide (Rubin, 1998). The consequence is<br />

that even at early passage numbers a certain percentage<br />

of cells stops dividing in a stochastic manner and eventually<br />

undergoes senescence. However, when the proportion<br />

of cells in G0/G1 decrease during mitotic<br />

stimulation, the amount of both Plk1 mRNA and protein<br />

increased. Highest levels of Plk1 mRNA and protein<br />

were found in populations with the highest proportion of<br />

cells in mitosis; the two FCS stimulated populations and<br />

the population blocked at the G2/M transition. Plk1<br />

expression can be used to characterize the level of synchronization<br />

in cells frequently used for a wide spectrum<br />

of in vitro techniques, in particular somatic nucleus<br />

transfer.<br />

In the present experiments, the levels of Plk1 mRNA<br />

determined by RT-PCR which is based on enzymatic<br />

amplification of the target molecule were compared to<br />

data obtained by RPA which is based on hybridization.<br />

The results obtained by these two methods were nearly<br />

identical. However, RPA requires more starting material<br />

and more time than RT-PCR. Immunodetection of<br />

Plk1 protein affords a third alternative for cell cycle<br />

analysis. It also requires more material and time than<br />

RT-PCR but obviously G0 cells contain no Plk1 protein<br />

while early G1 cells contain residual Plk1 protein from<br />

the previous cell cycle. Thus protein analysis is the most<br />

accurate method for the identification of cells at G0<br />

phase. For a rapid evaluation of synchronization efficiency<br />

or the proliferation status of cultured cells,<br />

especially when starting material is limited, RT-PCR<br />

measurement of Plk1 expression is the method of choice.<br />

In conclusion, results of the study reported here, indicate<br />

that expression of Plk1 is a suitable marker for<br />

determining cell cycle stage in porcine fetal fibroblasts<br />

and may thus be a useful tool in the future improvement<br />

of porcine somatic nuclear transfer.<br />

ACKNOWLEDGMENTS<br />

The authors thank Doris Herrmann and Lenka<br />

Travnickova for excellent technical assistance in the<br />

laboratory. We are grateful to Dr. Josef Fulka for discussion<br />

and critical reading of manuscript. We also thank<br />

Dr. Jiri Kanka for bovine skin fibroblasts.<br />

REFERENCES<br />

Abrieu A, Brassac T, Galas S, Fisher D, Labbe JC, Doree M. 1998. The<br />

Polo-like kinase Plx1 is a component of the MPF amplification loop at<br />

the G2/M-phase transition of the cell cycle in Xenopus eggs. J Cell Sci<br />

111:1751–1757.<br />

Altschul SF, Stephen F, Madden TL, Thomas L, Schäffer AA, Alejandro<br />

A, Jinghui Zhang, Zheng Zhang, Webb Miller, Lipman David J. 1997.<br />

Gapped BLAST and PSI-BLAST: A new generation of protein<br />

database search programs. Nucleic Acids Res 25:3389–3402.<br />

Clay FJ, McEwen SJ, Bertoncello I, Wilks AF, Dunn AR. 1993.<br />

Identification and cloning of a protein kinase-encoding mouse gene,<br />

Plk, related to the polo gene of Drosophila. Proc Natl Acad Sci USA<br />

9011:4882–4886.<br />

Clay FJ, Ernst MR, Trueman JW, Flegg R, Dunn AR. 1997. The mouse<br />

Plk gene: Structural characterization, chromosomal localization, and<br />

identification of a processed Plk pseudogene. Gene 198:329–339.<br />

Descombes P, Nigg EA. 1998. The polo-like kinase Plx1 is required for<br />

M phase exit and destruction of mitotic regulators in Xenopus egg<br />

extracts. EMBO J 175:1328–1335.<br />

Ferris DK, Maloid SC, Li CC. 1998. Ubiquitination and proteasome<br />

mediated degradation of polo-like kinase. Biochem Biophys Res<br />

Commun 18:340–344.<br />

Glover DM, Ohkura H, Tavares A. 1996. Polo kinase: The choreographer<br />

of the mitotic stage? J Cell Biol 135:1681–1684.<br />

Glover DM, Hagan IM, Tavares AA. 1998. Polo-like kinases: A team<br />

that plays throughout mitosis. Genes Dev 1224:3777–3787.<br />

Golsteyn RM, Schultz SJ, Bartek J, Ziemiecki A, Ried T, Nigg EA. 1994.<br />

Cell cycle analysis and chromosomal localization of human Plk1, a<br />

putative homologue of the mitotic kinases Drosophila polo and<br />

Saccharomyces cerevisiae Cdc5. J Cell Sci 107:1509–1517.<br />

Golsteyn RM, Mundt KE, Fry AM, Nigg EA. 1995. Cell cycle regulation<br />

of the activity and subcellular localization of PLK 1, a human protein<br />

kinase implicated in mitotic spindle function. J Cell Biol 129:1617–<br />

1628.<br />

Golsteyn RM, Lane HA, Mundt KE, Arnaud L, Nigg EA. 1996. The<br />

family of polo-like kinases. Prog Cell Cycle Res 2:107–114.<br />

Hamanaka R, Smith MR, O’Connor PM, Maloid S, Mihalic K, Spivak<br />

JL, Longo DL, Ferris DK. 1995. Polo-like kinase is a cell cycleregulated<br />

kinase activated during mitosis. J Biol Chem 8:21086–<br />

21091.<br />

Harlow E, Lane D. 1988. Antibodies. A laboratory mannual. Woodburg,<br />

NY. Cold Spring Harbor Laboratory.<br />

Hogan B, Constantini F, Lacy E. 1994. Manipulating the mouse<br />

embryo: A laboratory manual. Plainview, NY: Cold Spring Habour<br />

Press.


IDENTIFICATION OF POLO-LIKE KINASES IN PORCINE FIBROBLASTS 253<br />

Holtrich U, Wolf G, Yuan J, Bereiter-Hahn J, Karn T, Weiler M,<br />

Kauselmann G, Rehli M, Andreesen R, Kaufmann M, Kuhl D,<br />

Strebhardt K. 2000. Adhesion induced expression of the serine/<br />

threonine kinase Fnk in human macrophages. Oncogene 5:4832–<br />

4839.<br />

Hudson JW, Kozarova A, Cheung P, Macmillan JC, Swallow CJ, Cross<br />

JC, Dennis JW. 2001. Late mitotic failure in mice lacking Sak, a pololike<br />

kinase. Curr Biol 11:441–446.<br />

Karaiskou A, Jessus C, Brassac T, Ozon R. 1999. Phosphatase 2A and<br />

polo kinase, two antagonistic regulators of cdc25 activation and MPF<br />

auto-amplification. J Cell Sci 112:3747–3756.<br />

Kauselmann G, Weiler M, Wulff P, Jessberger S, Konietzko U, Scafidi<br />

J, Staubli U, Bereiter-Hahn J, Strebhardt K, Kuhl D. 1999. The pololike<br />

protein kinases Fnk and Snk associate with a Ca(2þ)- and<br />

integrin-binding protein and are regulated dynamically with<br />

synaptic plasticity. EMBO J 18:5528–5539.<br />

Kotani S, Tugendreich S, Fujii M, Jorgensen PM, Watanabe N, Hoog C,<br />

Hieter P, Todokoro K. 1998. PKA and MPF-activated polo-like kinase<br />

regulate anaphase-promoting complex activity and mitosis progression.<br />

Mol Cell 13:371–380.<br />

Kues WA, Anger M, Carnwath JW, Paul D, Motlik J, Niemann H. 2000.<br />

Cell cycle synchronization of porcine fetal fibroblasts: Effects of<br />

serum deprivation and reversible cell cycle inhibitors. Biol Reprod<br />

62:412–419.<br />

Kumagai A, Dunphy WG. 1996. Purification and molecular cloning of<br />

Plx1, a Cdc25-regulatory kinase from Xenopus egg extracts. Science<br />

6:1377–1380.<br />

Lake RJ, Jelinek WR. 1993. Cell cycle- and terminal differentiationassociated<br />

regulation of the mouse mRNA encoding a conserved<br />

mitotic protein kinase. Mol Cell Biol 13:7793–7801.<br />

Lee CK, Piedrahita JA. 2002. Inhibition of apoptosis in serum starved<br />

porcine embryonic fibroblasts. Mol Reprod Dev 62:106–112.<br />

Lee KS, Grenfell TZ, Yarm FR, Erikson RL. 1998. Mutation of<br />

the polo-box disrupts localization and mitotic functions of the<br />

mammalian polo kinase Plk. Proc Natl Acad Sci USA 9516:9301–<br />

9306.<br />

Llamazares S, Moreira A, Tavares A, Girdham C, Spruce BA, Gonzalez<br />

C, Karess RE, Glover DM, Sunkel CE. 1991. Polo encodes a protein<br />

kinase homolog required for mitosis in Drosophila. Genes Dev 12A:<br />

2153–2165.<br />

Mundt KE, Golsteyn RM, Lane HA, Nigg EA. 1997. On the regulation<br />

and function of human polo-like kinase 1 PLK1: Effects of overexpression<br />

on cell cycle progression. Biochem Biophys Res Commun<br />

2392:377–385.<br />

Nigg EA. 1998. Polo-like kinases: Positive regulators of cell division<br />

from start to finish. Curr Opin Cell Biol 106:776–783.<br />

Nigg EA. 2001. Mitotic kinases as regulators of cell division and its<br />

checkpoints. Nat Rev Mol Cell Biol 2:21–32.<br />

Onishi A, Iwamoto M, Akita T, Mikawa S, Takeda K, Awata T, Hanada<br />

H, Perry AC. 2000. Pig cloning by microinjection of fetal fibroblast<br />

nuclei. Science 18:1188–1190.<br />

Polejaeva IA, Chen SH, Vaught TD, Page RL, Mullins J, Ball S, Dai Y,<br />

Boone J, Walker S, Ayares DL, Colman A, Campbell KH. 2000.<br />

Cloned pigs produced by nuclear transfer from adult somatic cells.<br />

Nature 407:86–90.<br />

Qian YW, Erikson E, Li C, Maller JL. 1998. Activated polo-like kinase<br />

Plx1 is required at multiple points during mitosis in Xenopus laevis.<br />

Mol Cell Biol 187:4262–4271.<br />

Qian YW, Erikson E, Maller JL. 1999. Mitotic effects of a constitutively<br />

active mutant of the Xenopus polo-like kinase Plx1. Mol Cell Biol<br />

19:8625–8632.<br />

Roshak AK, Capper EA, Imburgia C, Fornwald J, Scott G, Marshall LA.<br />

2000. The human polo-like kinase, PLK, regulates cdc2/cyclin B<br />

through phosphorylation and activation of the cdc25C phosphatase.<br />

Cell Signal 12:405–411.<br />

Rubin H. 1998. Cell aging in vivo and in vitro. Mech Ageing Dev<br />

100:209–210.<br />

Smits VA, Klompmaker R, Arnaud L, Rijksen G, Nigg EA, Medema RH.<br />

2000. Polo-like kinase-1 is a target of the DNA damage checkpoint.<br />

Nat Cell Biol 29:672–676.<br />

Tatusova TA, Madden TL. 1999. Blast 2 sequences, a new tool for<br />

comparing protein and nucleotide sequences. FEMS Microbiol Lett<br />

15:247–250.<br />

Uchiumi T, Longo DL, Ferris DK. 1997. Cell cycle regulation of the<br />

human polo-like kinase PLK promoter. J Biol Chem 4:9166–9174.<br />

Wilmut I, Campbell KH. 1998. Quiescence in nuclear transfer. Science<br />

11:1611.<br />

Wolf G, Elez R, Doermer A, Holtrich U, Ackermann H, Stutte HJ,<br />

Altmannsberger HM, Rubsamen-Waigmann H, Strebhardt K. 1997.<br />

Prognostic significance of polo-like kinase PLK expression in nonsmall<br />

cell lung cancer. Oncogene 14:543–549.<br />

Wolf G, Hildenbrand R, Schwar C, Grobholz R, Kaufmann M, Stutte<br />

HJ, Strebhardt K, Bleyl U. 2000. Polo-like kinase: A novel marker of<br />

proliferation: Correlation with estrogen-receptor expression in<br />

human breast cancer. Pathol Res Pract 196:753–759.<br />

Yuan J, Horlin A, Hock B, Stutte HJ, Rubsamen-Waigmann H,<br />

Strebhardt K. 1997. Polo-like kinase, a novel marker for cellular<br />

proliferation. Am J Pathol 1165–1172.


Telomere length is reset during early<br />

mammalian embryogenesis<br />

Sonja Schaetzlein*, Andrea Lucas-Hahn † , Erika Lemme † , Wilfried A. Kues † , Martina Dorsch ‡ , Michael P. Manns*,<br />

Heiner Niemann †§ , and K. Lenhard Rudolph* §<br />

*Department of Gastroenterology, Hepatology, and Endocrinology, and ‡ Institute for Animal Science, Hannover Medical School, 30625 Hannover, Germany;<br />

and † Institute for Animal Science, Mariensee, Federal Agricultural Research Center, Höltystrasse 10, D-31535 Neustadt, Germany<br />

Communicated by Neal L. First, University of Wisconsin, Madison, WI, April 5, 2004 (received for review January 15, 2004)<br />

The enzyme telomerase is active in germ cells and early embryonic<br />

development and is crucial for the maintenance of telomere length.<br />

Whereas the different length of telomeres in germ cells and<br />

somatic cells is well documented, information on telomere length<br />

regulation during embryogenesis is lacking. In this study, we<br />

demonstrate a telomere elongation program at the transition from<br />

morula to blastocyst in mice and cattle that establishes a specific<br />

telomere length set point during embryogenesis. We show that<br />

this process restores telomeres in cloned embryos derived from<br />

fibroblasts, regardless of the telomere length of donor nuclei, and<br />

that telomere elongation at this stage of embryogenesis is telomerase-dependent<br />

because it is abrogated in telomerase-deficient<br />

mice. These data demonstrate that early mammalian embryos have<br />

a telomerase-dependent genetic program that elongates telomeres<br />

to a defined length, possibly required to ensure sufficient<br />

telomere reserves for species integrity.<br />

The enzyme telomerase is active in germ cells and during early<br />

embryogenesis (1–5) and is crucial for the maintenance of<br />

telomere length and germ cell viability in successive generations<br />

of a species (6, 7). Telomere shortening limits the regenerative<br />

capacity of cells and is correlated with the onset of cancer, aging,<br />

and chronic diseases (8–12). The segregation of telomere length<br />

from one generation to the next seems to be essential for normal<br />

development and well <strong>bei</strong>ng in mammals. In line with this<br />

hypothesis, telomere shortening in telomerase knockout<br />

(mTERC/ ) mice is correlated with a variety of phenotypes<br />

including impaired organ regeneration, chromosomal instability,<br />

and cancer (9, 13–16).<br />

Telomere length segregation is ensured by the active elongation<br />

and maintenance program in germ cells. Mature germ cells<br />

possess significantly longer telomeres compared to somatic cells<br />

(17), possibly because of telomere elongation during germ cell<br />

maturation (18). In contrast to telomere elongation in germ cells,<br />

little is known about telomere length regulation during embryogenesis.<br />

Establishment of telomere length during embryogenesis<br />

could determine telomere reserves in newborns. In newborn<br />

mice derived from crossbreeding of late generation mTERC/ mice and mTERC/ mice, telomere function is rescued correlating<br />

with the elongation of critically short telomeres (19).<br />

Similarly, mice derived from intercrosses of mouse strains with<br />

different telomere lengths show a telomerase-dependent elongation<br />

of those telomeres derived from the strain with shorter<br />

telomeres (20, 21). These observations can be explained only if<br />

the telomeres are elongated during embryogenesis. In line with<br />

this hypothesis, studies on cloned mice and cattle have revealed<br />

telomere length restoration of cloned animals even when senescent<br />

fibroblasts had been used as the donor cells (22–25). In<br />

contrast, cloned sheep derived from epithelial cells had shorter<br />

telomeres than control animals (26). Telomere length rescue in<br />

cloned animals seems to depend on the used donor cell type (27).<br />

Alternatively, rather than telomere restoration during embryogenesis,<br />

cloned animals may be derived from the selective<br />

propagation of cells with enhanced telomere reserves during the<br />

cloning process andor early embryonic development, which<br />

would be consistent with the low efficiency of somatic cloning.<br />

If telomere restoration in fact occurs during preimplantation<br />

development, it remains to be investigated whether this restoration<br />

is related to the cloning process itself or represents a<br />

general mechanism during embryogenesis to ensure the presence<br />

of adequate telomere reserves for species integrity. To test<br />

whether telomere length is regulated during embryogenesis, we<br />

have analyzed telomere length at different stages of bovine and<br />

mouse early development employing embryos derived from<br />

somatic nuclear transfer, in vitro production (in vitro maturation,<br />

fertilization, and culture), and conventional breeding. Results<br />

show that telomere length is determined at morula to blastocyst<br />

transition by a telomerase-dependent mechanism. Telomere<br />

elongation is restricted to this stage of development and restores<br />

telomeres of fibroblast-derived cloned embryos to normal<br />

length.<br />

Methods<br />

Quantitative Fluorescence in Situ Hybridization (Q-FISH) on Interphase<br />

Nuclei. Q-FISH was performed on interphase nuclei as described<br />

on nuclei of adult and fetal fibroblasts (11, 28). Nuclei were<br />

isolated by cell lysis with 0.075 M KCl and fixed in glacial<br />

methanolacetic acid (3:1). Bovine adult and fetal fibroblasts<br />

and a human transformed kidney cell line (Phoenix cells, a gift<br />

from Gary Nolan, Stanford University, Stanford, CA) were<br />

analyzed in parallel with Q-FISH and Southern blotting to<br />

validate Q-FISH data. Results correlated closely between the<br />

two methods. To ensure linearity of the Q-FISH method,<br />

telomere length of liver cells from different generations of<br />

telomerase-deficient mice were analyzed by Q-FISH showing the<br />

expected decrease in telomere length from one generation to the<br />

other (data not shown).<br />

Telomere Restriction Fragment (TRF) Length Analysis. TRF were<br />

determined in skin biopsies and blood samples (WBC) of cloned<br />

and conventionally produced cattle as well as bovine in vitro<br />

matured oocytes (see below) and bovine semen as controls. TRF<br />

length analysis was done as described in ref. 28. The mean TRF<br />

length was calculated by measuring the signal intensity in 10 squares<br />

covering the entire TRF smear. All calculations were performed<br />

with PCBAS and EXCEL (Microsoft) computer programs.<br />

Telomeric Repeat Amplification Protocol. To measure telomerase<br />

activity in early embryos, a telomeric repeat amplification protocol<br />

assay was performed with a TRAPeze telomerase detection<br />

kit (Intergen, Purchase, NY) according to recommendations<br />

of the manufacturer.<br />

Abbreviations: Q-FISH, quantitative fluorescence in situ hybridization; TRF, telomere restriction<br />

fragment.<br />

§ To whom correspondence should be addressed. E-mail: niemann@tzv.fal.de or rudolph.<br />

lenhard@mh-hannover.de.<br />

© 2004 by The National Academy of Sciences of the USA<br />

8034–8038 PNAS May 25, 2004 vol. 101 no. 21 www.pnas.orgcgidoi10.1073pnas.0402400101


In Vitro Production of Bovine Embryos. Bovine embryos were<br />

produced as described in refs. 29 and 30. Briefly, viable cumulus–<br />

oocyte complexes derived from abattoir ovaries were matured in<br />

vitro in groups of 15–20 in 100 l of TCM-199 supplemented with<br />

10 IU of pregnant mare serum gonadotropin and 5 IU of human<br />

chorionic gonadotropin (Suigonan, Intervet, Tönisvorst, Germany)<br />

and 0.1% BSA (Sigma, A7030) under silicone oil in a<br />

humidified atmosphere composed of 5% CO2 in air at 39°C for<br />

24 h. Matured cumulus–oocyte complexes were fertilized in vitro<br />

employing 1 10 6 sperm per ml of frozenthawed semen from<br />

one bull with proven fertility in in vitro fertilization (29) during<br />

a 19-h coincubation under the same temperature and gas conditions<br />

as described for in vitro maturation. Presumptive zygotes<br />

were cultured up to the morula stage (day 6 after insemination)<br />

or blastocyst stage (day 8 after insemination) in synthetic oviduct<br />

fluid medium supplemented with BSA, in a mixture of 5% O2,<br />

90% N2, and 5% CO2 (Air Products, Hattingen, Germany) in<br />

modular incubator chambers (ICN).<br />

In Vivo Production of Bovine Embryos. In vivo grown bovine embryos<br />

were collected from superovulated donor animals inseminated with<br />

semen from the same bull as used for in vitro fertilization. Morulae<br />

and blastocysts were nonsurgically recovered from the genital tract<br />

of donors employing established protocols at days 6 and 8 after<br />

artificial insemination, respectively (31).<br />

Generation of Nuclear Transfer-Derived Embryos. Bovine fetal fibroblasts<br />

were obtained from a 61-day female bovine fetus after<br />

evisceration and decapitation. Adult fibroblasts were established<br />

from ear skin biopsies of an adult female animal collected from a<br />

local abattoir. The fibroblasts used for nuclear transfer in these<br />

experiments were from passages 2–4 and were induced to enter a<br />

period of quiescence (presumptive G0) by serum starvation for 2–3<br />

days (0.5% FCS). Donor cells and cloned animals showed an<br />

identical microsatellite pattern (not shown). For enucleation and<br />

nuclear transfer, cytochalasin B (7.5 gml) was added to TCM-air.<br />

For fusion, 0.285 M mannitol containing 0.1 mM MgSO4 and 0.05%<br />

BSA was used. Embryos were reconstructed, and cloned offspring<br />

were produced as described in ref. 32. Briefly, in vitro matured<br />

oocytes were enucleated by aspirating the first polar body and the<br />

metaphase II plate. The donor cells were pelleted and resuspended<br />

in TCM-air and remained in this medium until insertion. A single<br />

cell was sucked into a 30-m (outer diameter) pipette and was then<br />

carefully transferred into the perivitelline space of the recipient<br />

oocyte. Cell fusion was induced with 1–2 DCpulsesof0.7kVcm<br />

for 30 s each generated by a Kruess electrofusion machine (CFA<br />

400, Hamburg, Germany). At 27 h after onset of maturation, the<br />

reconstructed embryos were chemically activated by incubation in<br />

5 M ionomycin (Sigma) in TCM-199 for 5 min followed by a 3–4-h<br />

incubation in 2 mM 6-dimethylaminopurine (Sigma) in TCM-199 at<br />

37°C. After three washes, the embryos were cultured for 6 (morulae)<br />

or 8 (blastocysts) days as described above. Morulae and<br />

blastocysts were used in Q-FISH for determination of telomere<br />

length.<br />

Production of Mouse Embryos. Wild-type and telomerase knockout<br />

(mTERC / , generation 2) females were killed on days 3<br />

(eight-cell and morula stage) or 4 (blastocyst stage) of pregnancy.<br />

Day 0.5 is the morning after insemination. The embryos<br />

were collected from the oviducts or the uterus as described (33).<br />

Statistical Programs. Student’s t test and GRAPHPAD INSTAT software<br />

were used to calculate statistical significances and standard<br />

deviations.<br />

Lucas-Hahn, A., Lemme, E., Hadeler, K. G., Sander, H. G. & Niemann, H. (2002) Theriogenology<br />

57, 433 (abstr.).<br />

Fig. 1. Telomere length analysis of bovine germ cells and primary fibroblasts<br />

used for cloning. (A) Distribution of the mean telomere length as determined<br />

by Q-FISH in different nuclei of adult fibroblasts (total mean, 10.84 5.73 kb;<br />

n 118), (B) fetal fibroblasts (total mean, 14.61 4.58 kb; n 40), and (C)<br />

oocytes (total mean, 16.95 2.51 kb; n 13). (D and E) The length of TRFs was<br />

determined by Southern blotting on DNA of bovine fetal fibroblasts, adult<br />

fibroblasts, and semen. (D) Histogram on the mean value of TRF length as<br />

determined in triplicate. Telomere length was significantly shorter in both<br />

fibroblast cell lines (adult fibroblasts, 12.03 0.46 kb; fetal fibroblasts, 13.37 <br />

0.46 kb) used for cloning compared to bovine semen (15.83 0.41 kb) used for<br />

in vivo or in vitro fertilization. (E) Representative radiograph of a Southern<br />

blot showing the TRF smear of adult bovine fibroblasts and semen. Dotted line<br />

in A–C indicates mean value.<br />

All experiments were conducted according to the German<br />

animal welfare law and comply with international regulations.<br />

Results<br />

Telomere Length in Bovine Preimplantation Embryos at the Morula<br />

and Blastocyst Stage. We analyzed telomere length regulation<br />

during early embryogenesis in bovine embryos and animals derived<br />

from cloning or conventional production to elucidate whether the<br />

embryo has a defined telomere-elongation program. Q-FISH (Fig.<br />

1 A–C) and Southern blot (Fig. 1E) analysis revealed significantly<br />

shorter telomeres in both adult and fetal fibroblasts that had been<br />

used for somatic nuclear transfer compared to bovine oocytes and<br />

semen serving as controls (Fig. 1 C and E). The telomere length of<br />

the adult fibroblasts was shorter than that of fetal fibroblasts (Fig.<br />

1 A, B, and D). Telomere length was analyzed in morula stages,<br />

which are known to be telomerase negative or to possess minimum<br />

telomerase activity (3–5). Q-FISH analysis revealed significantly<br />

shorter mean telomeres in cloned morulae compared to their<br />

counterparts derived from in vivo or in vitro fertilization (Fig. 2).<br />

The finding of similar telomere lengths in morulae derived from in<br />

vivo and in vitro fertilization ruled out the possibility that the in vitro<br />

culture per se had a significant effect on telomere length (Fig. 2 A,<br />

Schaetzlein et al. PNAS May 25, 2004 vol. 101 no. 21 8035<br />

DEVELOPMENTAL<br />

BIOLOGY


Fig. 2. Telomere length of cloned and fertilized bovine morulae by Q-FISH.<br />

(A–D). Distribution of mean telomere length over all analyzed nuclei of<br />

morulae (day 6 of embryonic development) derived from in vivo fertilization<br />

(n 234) (A), in vitro fertilization (n 118) (B), cloning from adult bovine<br />

fibroblasts (n 203) (C), and cloning from fetal bovine fibroblasts (n 134)<br />

(D). (E) Mean value of the mean telomere lengths determined for individual<br />

morulae derived by cloning from fetal fibroblasts (8.71 1.81 kb, n 8) or<br />

adult fibroblasts (9.47 2.07 kb, n 6) from in vivo fertilization (12.42 1.37<br />

kb, n 7) or in vitro fertilization (14.36 2.67 kb, n 4). The difference in<br />

telomere length between morulae derived from in vivo and in vitro fertilization<br />

was not significant. Dotted line in A–D indicates mean value.<br />

B, and E). Telomeres were shorter in all groups of morulae<br />

compared to donor cells or germ cells, indicating that there was no<br />

significant selection in favor of cells with longer telomere reserves<br />

during the cloning process.<br />

Next, telomere lengths were analyzed in blastocysts derived from<br />

in vitro fertilization or somatic cloning. In all groups, blastocysts had<br />

significantly elongated telomeres relative to those at the morula<br />

stage (Fig. 3), indicating that telomeres are elongated at this<br />

particular stage of bovine embryogenesis. Similarly, a sequential<br />

analysis of telomere length in embryos derived from a single in vitro<br />

fertilization experiment showed significant telomere elongation in<br />

Fig. 3. Telomere length of cloned and fertilized bovine blastocysts by<br />

Q-FISH. (A) Representative photographs of telomere spots in nuclei of blastocysts<br />

(day 8 of development) derived from cloning of adult and fetal<br />

fibroblasts and in vitro fertilization. Magnification bar, 20 m. (B–D) Distribution<br />

of mean telomere length over all analyzed nuclei of blastocysts derived<br />

from in vitro fertilization (n 129) (B), cloning from fetal bovine fibroblasts<br />

(n 192) (C), and cloning from adult bovine fibroblasts (n 126) (D). (E) Mean<br />

value of the mean telomere lengths determined for individual blastocysts<br />

derived from in vitro fertilization (19.53 4.6 kb, n 6) and from cloning<br />

using fetal fibroblasts (17.3 2.66 kb, n 6) and adult fibroblasts (21.67 3.92<br />

kb, n 5). Note that telomeres were elongated in all groups compared to the<br />

morula stage and that in contrast to the differences in telomere length at the<br />

morula stage, there was no significant difference between the different<br />

groups of blastocysts. (F) A sequential analysis of telomere length in embryos<br />

derived from a single in vitro fertilization experiment showed telomere<br />

elongation at the morula–blastocyst transition, with an average of 15.8 3.1<br />

kb for morula stages (n 6) and 28.7 7.4 kb for blastocyst stages (n 6).<br />

Dotted line in B–D indicates mean value.<br />

embryos of an identical genetic background at the morula–<br />

blastocyst transition (Fig. 3F). In accordance with previous studies,<br />

we did not detect telomerase activity at the morula stage. In<br />

contrast, a strong up-regulation of telomerase was observed at the<br />

blastocyst stage (3–5), indicating that telomere elongation correlated<br />

with the timing of telomerase reactivation during embryogenesis.<br />

Telomere lengths of bovine blastocysts were similar regardless<br />

of whether they were derived from somatic cloning or in<br />

vitro fertilization (Fig. 3). These data provide experimental evidence<br />

for an embryo-specific telomere elongation program, which<br />

leads to restoration of telomere length in both cloned embryos and<br />

embryos produced by in vivo or in vitro fertilization.<br />

In line with these observations, telomere length analysis on 1- to<br />

2-year-old cloned cattle derived from fetal or adult donor cells<br />

revealed a similar telomere length when measured in WBC or ear<br />

biopsies compared to those of age-matched control animals (data<br />

not shown). Because WBC are derived from the telomerasepositive<br />

hematopoietic compartment, these data indicate that telomere<br />

elongation in cloned cattle is restricted to a defined stage of<br />

embryogenesis and does not result in abnormally long telomeres in<br />

telomerase-positive compartments. The telomere length of WBC<br />

and ear biopsies was comparable between clones derived from adult<br />

or fetal fibroblasts, indicating that the cell source did not significantly<br />

affect telomere length in the cloned offspring.<br />

Telomere Elongation During Early Mouse Embryogenesis Is Telomerase-Dependent.<br />

To test whether the observed telomere elongation<br />

is a general mechanism in mammalian development and whether it<br />

8036 www.pnas.orgcgidoi10.1073pnas.0402400101 Schaetzlein et al.


Fig. 4. Telomere length in embryos from mTERC / and mTERC / mice by<br />

Q-FISH. (A) The distribution of mean telomere fluorescence units (fu) of all<br />

analyzed nuclei from wild-type embryos at the 8-cellmorula stage (1118 484<br />

fu, n 274) and at the blastocyst stage (1501 753 fu, n 681). (B) Mean value<br />

of the mean telomere fluorescence intensity determined for individual embryos<br />

at the 8-cellmorula stage (1,066 406 fu, n 32) and at the blastocyst stage<br />

(1,350 565 fu, n31). (C) The distribution of mean telomere fu of all nuclei from<br />

telomerase knockout embryos (mTERC / ) at the 8-cellmorula stage (1,060 <br />

503 fu, n 96) and at the blastocyst stage (949 388 fu, n 199). (D) Mean value<br />

of the telomere fluorescence intensity determined for individual embryos at the<br />

8-cellmorula stage (995 356 fu, n 14) and at the blastocyst stage (877 348<br />

fu, n 10). Note that mTERC / embryos showed telomere elongation at morulablastocyst<br />

transition, whereas embryos from telomerase-deficient mice<br />

(mTERC / ) did not show a lengthening of telomeres but rather a slight decrease<br />

in telomere length. (E and F) Representative photographs of telomere spots in<br />

nuclei of mTERC / (E) and mTERC / (F) embryo at 8-cell stagemorula and<br />

blastocyst stage of embryonic development. Magnification bar, 10 m. (G) Mean<br />

value of the telomere fluorescence intensity determined for individual embryos<br />

at the 8-cellmorula stage, blastocyst stage, day 8.510.5 (1,329 353 fu, n 5),<br />

and day 13.5 (1,573 133 fu, n 5). Analysis revealed no significant increase in<br />

telomere length at these postimplantation stages (P 0.05). Dotted line in A and<br />

C indicates mean value.<br />

is telomerase-dependent, we analyzed telomere length in early<br />

embryogenesis of mTERC / and mTERC / mice. In agreement<br />

with the findings from bovine embryos, the telomeres of<br />

mTERC / embryos were significantly elongated at the blastocyst<br />

stage compared to morulae and 8-cell stages (Fig. 4 A, B, and E).<br />

1. Mantell, L. L. & Greider, C. W. (1994) EMBO J. 13, 3211–3217.<br />

2. Wright, W. E., Piatyszek, M. A., Rainey, W. E., Byrd, W. & Shay, J. W. (1996)<br />

Dev. Genet. 18, 173–179.<br />

3. Xu, J. & Yang, X. (2000) Biol. Reprod. 63, 1124–1128.<br />

4. Wright, D. L., Jones, E. L., Mayer, J. F., Oehninger, S., Gibbons, W. E. &<br />

Lanzendorf, S. E. (2001) Mol. Hum. Reprod. 7, 947–955.<br />

In contrast, in mTERC / embryos, telomeres were not elongated<br />

during the morula–blastocyst transition (Fig. 4 C, D, and F),<br />

indicating that telomerase activity is required for telomere elongation<br />

at this stage of embryogenesis. To determine whether telomere<br />

elongation is restricted to the morula–blastocyst transition or is a<br />

continuous process during early development, we analyzed telomere<br />

length of primary cultures of adherent growing cells derived<br />

from embryos at days 8.5 (Theiler’s stage 13), 10.5 (Theiler’s stage<br />

17), and 13.5 of mouse embryogenesis. Analysis revealed no<br />

significant increase in telomere length at these postimplantation<br />

stages compared with length at the morula–blastocyst transition<br />

(Fig. 4G), indicating that telomere elongation during embryogenesis<br />

is restricted to the morula–blastocyst transition.<br />

Discussion<br />

The discovery of telomere elongation at the morula–blastocyst<br />

transition in embryos from two mammalian species as well as<br />

embryos derived from either in vivo or in vitro fertilization and<br />

even somatic nuclear transfer indicates a general program in<br />

preimplantation development. Morula–blastocyst transition is a<br />

critical step in preimplantation development, leading to first<br />

differentiation into two cell lineages, the inner cell mass and the<br />

trophoblast. This coincides with a dramatic change in morphology<br />

from the compacted morula to the cavity-filled blastocyst<br />

stage and is associated with massive changes in embryonic gene<br />

expression (32). Our study shows that telomeres of cloned<br />

embryos derived from fetal or adult fibroblasts are normalized<br />

by the telomere elongation program at the morula–blastocyst<br />

transition, indicating that this program resets telomere length to<br />

a specific set point. Fibroblasts are a predominantly used donor<br />

cell type in somatic nuclear transfer in farm animals. It remains<br />

to be determined why telomere length in cloned embryos derived<br />

from epithelial cell lines is not always restored (27).<br />

Telomeres are not elongated in mTERC / mouse embryos,<br />

indicating that the telomerase enzyme is required for telomere<br />

elongation during early embryogenesis. These data coincide with<br />

previous reports on haploinsufficiency of the telomerase RNA<br />

component (TERC) in mTERC / mice, leading to offspring from<br />

intercrosses of mouse strains with different telomere lengths (20,<br />

21). However, because telomerase activation alone does not necessarily<br />

result in telomere elongation in vitro (34), other factors such<br />

as telomere binding proteins could be involved in telomere length<br />

regulation at this stage of embryogenesis (35). Interestingly, the<br />

telomere elongation program discovered in the present study<br />

correlates with activation of telomerase at the same stage of human<br />

embryogenesis (4). In bovine-cloned and parthenogenetic embryos,<br />

telomerase activity followed the same pattern with a significant<br />

up-regulation at the blastocyst stage (36). Telomere elongation<br />

during embryogenesis could be critical to ensure normal telomere<br />

length segregation from one generation to the next and could have<br />

direct effects on regeneration, aging, and carcinogenesis during<br />

postnatal life. A detailed understanding of this embryonic telomere-elongation<br />

program could ultimately be important for the<br />

use of cell transplantation and gene therapy approaches for the<br />

treatment of degenerative disorders.<br />

We thank Ande Satyanarayana for critical discussion and K.-G. Hadeler<br />

for artificial insemination and bovine embryo flushing. Parts of this study<br />

were supported through funding from the Deutsche Forschungsgemeinschaft<br />

(Ni 25618-1). K.L.R. is supported by the Emmy–Noether Program<br />

of the Deutsche Forschungsgemeinschaft (Ru 7452-1) and the<br />

Deutsche Krebshilfe e.V. (10-1809-Ru1).<br />

5. Betts, D. H. & King, W. A. (1999) Dev. Genet. 25, 397–403.<br />

6. Hemann, M. T., Rudolph, K. L., Strong, M. A., DePinho, R. A., Chin, L. &<br />

Greider, C. W. (2001) Mol. Biol. Cell 12, 2023–2030.<br />

7. Liu, L., Blasco, M., Trimarchi, J. & Keefe, D. (2002) Dev. Biol. 249, 74–84.<br />

8. Djojosubroto, M. W., Choi, Y. S., Lee, H. W. & Rudolph, K. L. (2003) Mol.<br />

Cells 15, 164–175.<br />

Schaetzlein et al. PNAS May 25, 2004 vol. 101 no. 21 8037<br />

DEVELOPMENTAL<br />

BIOLOGY


9. Rudolph, K. L., Chang, S., Lee, H. W., Blasco, M., Gottlieb, G. J., Greider, C.<br />

& DePinho, R. A. (1999) Cell 96, 701–712.<br />

10. Rudolph, K. L., Millard, M., Bosenberg, M. W. & DePinho, R. A. (2001) Nat.<br />

Genet. 28, 155–159.<br />

11. Satyanarayana, A., Wiemann, S. U., Buer, J., Lauber, J., Dittmar, K. E.,<br />

Wustefeld, T., Blasco, M. A., Manns, M. P. & Rudolph, K. L. (2003) EMBO<br />

J. 22, 4003–4013.<br />

12. Cawthon, R. M., Smith, K. R., O’Brien, E., Sivatchenko, A. & Kerber, R. A.<br />

(2003) Lancet 361, 393–395.<br />

13. Rudolph, K. L., Chang, S., Millard, M., Schreiber-Agus, N. & DePinho, R. A.<br />

(2000) Science 287, 1253–1258.<br />

14. Lee, H., Blasco, M., Gottlieb, G., Horner, J. N., Greider, C. & DePinho, R.<br />

(1998) Nature 392, 569–574.<br />

15. Artandi, S. E., Chang, S., Lee, S. L., Alson, S., Gottlieb, G. J., Chin, L. &<br />

DePinho, R. A. (2000) Nature 406, 641–645.<br />

16. Samper, E., Flores, J. & Blasco, M. (2001) EMBO Rep. 2, 800–807.<br />

17. de Lange, T., Shiue, L., Myers, R. M., Cox, D. R., Naylor, S. L., Killery, A. M.<br />

& Varmus, H. E. (1990) Mol. Cell. Biol. 10, 518–527.<br />

18. Achi, M. V., Ravindranath, N. & Dym, M. (2000) Biol. Reprod. 63, 591–598.<br />

19. Hemann, M. T., Strong, M. A., Hao, L. Y. & Greider, C. W. (2001) Cell 107, 67–77.<br />

20. Hathcock, K. S., Hemann, M. T., Opperman, K. K., Strong, M. A., Greider,<br />

C. W. & Hodes, R. J. (2002) Proc. Natl. Acad. Sci. USA 99, 3591–3596.<br />

21. Zhu, L., Hathcock, K. S., Hande, P., Lansdorp, P. M., Seldin, M. F. & Hodes,<br />

R. J. (1998) Proc. Natl. Acad. Sci. USA 95, 8648–8653.<br />

22. Wakayama, T., Shinkai, Y., Tamashiro, K. L., Niida, H., Blanchard, D. C.,<br />

Blanchard, R. J., Ogura, A., Tanemura, K., Tachibana, M., Perry, A. C., et al.<br />

(2000) Nature 407, 318–319.<br />

23. Tian, X. C., Xu, J. & Yang, X. (2000) Nat. Genet. 26, 272–273.<br />

24. Betts, D., Bordignon, V., Hill, J., Winger, Q., Westhusin, M., Smith, L. & King,<br />

W. (2001) Proc. Natl. Acad. Sci. USA 98, 1077–1082.<br />

25. Lanza, R. P., Cibelli, J. B., Blackwell, C., Cristofalo, V. J., Francis, M. K.,<br />

Baerlocher, G. M., Mak, J., Schertzer, M., Chavez, E. A., Sawyer, N., et al.<br />

(2000) Science 288, 665–669.<br />

26. Shiels, P. G., Kind, A. J., Campbell, K. H., Waddington, D., Wilmut, I., Colman,<br />

A. & Schnieke, A. E. (1999) Nature 399, 316–317.<br />

27. Miyashita, N., Shiga, K., Yonai, M., Kaneyama, K., Kobayashi, S., Kojima, T.,<br />

Goto, Y., Kishi, M., Aso, H., Suzuki, T., et al. (2002) Biol. Reprod. 66,<br />

1649–1655.<br />

28. Wiemann, S. U., Satyanarayana, A., Tsahuridu, M., Tillmann, H. L., Zender,<br />

L., Klempnauer, J., Flemming, P., Franco, S., Blasco, M. A., Manns, M. P. &<br />

Rudolph, K. L. (2002) FASEB J. 16, 935–942.<br />

29. Wrenzycki, C., Herrmann, D., Keskintepe, L., Martins, A., Jr., Sirisathien, S.,<br />

Brackett, B. & Niemann, H. (2001) Hum. Reprod. 16, 893–901.<br />

30. Eckert, J. & Niemann, H. (1995) Theriogenology 43, 1211–1225.<br />

31. Bungartz, L. & Niemann, H. (1994) J. Reprod. Fertil. 101, 583–591.<br />

32. Niemann, H., Wrenzycki, C., Lucas-Hahn, A., Brambrink, T., Kues, W. A. &<br />

Carnwath, J. W. (2002) Cloning Stem Cells 4, 29–38.<br />

33. Hogan, B., Beddington, R., Costantini, F. & Lacy, E. (1994) Manipulating the<br />

Mouse Embryo (Cold Spring Harbor Lab. Press, Plainview, NY).<br />

34. Zhu, J., Wang, H., Bishop, J. M. & Blackburn, E. H. (1999) Proc. Natl. Acad.<br />

Sci. USA 96, 3723–3728.<br />

35. Loayza, D. & De Lange, T. (2003) Nature 424, 1013–1018.<br />

36. Xu, J. & Yang, X. (2001) Biol. Reprod. 64, 770–774.<br />

8038 www.pnas.orgcgidoi10.1073pnas.0402400101 Schaetzlein et al.


Review TRENDS in Biotechnology Vol.22 No.6 June 2004<br />

The contribution of farm animals to<br />

human health<br />

Wilfried A. Kues and Heiner Niemann<br />

Department of Biotechnology, Institut für Tierzucht, Mariensee, D-31535 Neustadt, Germany<br />

Farm animals and their products have a longstanding<br />

and successful history of providing significant contributions<br />

to human nutrition, clothing, facilitation of<br />

labour, research, development and medicine and have<br />

thus been essential in improving life expectancy and<br />

human health. With the advent of transgenic technologies<br />

the potential of farm animals for improving<br />

human health is growing and many areas remain to be<br />

explored. Recent breakthroughs in reproductive technologies,<br />

such as somatic cloning and in vitro embryo<br />

production, and their merger with molecular genetic<br />

tools, will further advance progress in this field. Here,<br />

we have summarized the contribution of farm animals<br />

to human health, covering the production of antimicrobial<br />

peptides, dietary supplements or functional foods,<br />

animals used as disease models and the contribution of<br />

animals to solving urgent environmental problems and<br />

challenges in medicine such as the shortage of human<br />

cells, tissues and organs and therapeutic proteins.<br />

Some of these areas have already reached the level of<br />

preclinical testing or commercial application, others<br />

will be further advanced only when the genomes of the<br />

animals concerned have been sequenced and annotated.<br />

Provided the necessary precautions are <strong>bei</strong>ng<br />

taken, the transmission of pathogens from animals to<br />

humans can be avoided to provide adequate security.<br />

Overall, the promising perspectives of farm animals and<br />

their products warrant further research and development<br />

in this field.<br />

Farm animals have made significant contributions to<br />

human health and well-<strong>bei</strong>ng throughout mankind’s<br />

history. The pioneering work of Edward Jenner with<br />

cowpox in the 18th century paved the way for modern<br />

vaccination programs against smallpox, as well as other<br />

human and animal plagues. To date, more than 250 million<br />

people have benefited from drugs and vaccines produced<br />

by recombinant technologies in bacteria and various types<br />

of mammalian cells and many more will benefit in the<br />

future (New Medicines in Development for Biotechnology,<br />

2002; www.phrma.org/newmedicines/biotech/). Further<br />

examples of the significant contribution of farm animals<br />

to human health are the longstanding use of bovine and<br />

porcine insulin for treatment of diabetes as well as horse<br />

antisera against snake venoms and antimicrobial peptides.<br />

In addition, farm animals are models for novel<br />

surgical strategies, testing of biodegradable implants and<br />

sources of tissue replacements, such as skin and heart<br />

valves.<br />

Progress in transgenic technologies has allowed the<br />

generation of genetically modified large animals for<br />

applications in agriculture and biomedicine, such as the<br />

production of recombinant proteins in the mammary gland<br />

and the generation of transgenic pigs with expression of<br />

human complement regulators in xenotransplantation<br />

research [1]. Further promising application perspectives<br />

will be developed when somatic cloning with genetically<br />

modified donor cells is further improved and the genomes<br />

of farm animals are sequenced and annotated. The first<br />

transgenic livestock were born less than 20 years ago with<br />

the aid of microinjection technology [2]. Recently the first<br />

animals with knockout of one or even two alleles of a<br />

targeted gene were reported (Table 1). Somatic nuclear<br />

transfer has been successful in 10 species, but the overall<br />

Table 1. Milestones (live offspring) in transgenesis and reproductive technologies in farm animals<br />

Year Milestone Strategy Refs<br />

1985 First transgenic sheep and pigs Microinjection of DNA into one pronucleus of a zygote [2]<br />

1986 Embryonic cloning of sheep Nuclear transfer using embryonic cells as donor cells [91]<br />

1997 Cloning of sheep with somatic donor cells Nuclear transfer using adult somatic donor cells [92]<br />

1997 Transgenic sheep produced by nuclear transfer Random integration of the construct [93]<br />

1998 Transgenic cattle produced from fetal fibroblasts and nuclear transfer Random integration of the construct [94]<br />

1998 Generation of transgenic cattle by MMLV injection Injection of oocytes with helper viruses [95]<br />

2000 Gene targeting in sheep Gene replacement and nuclear transfer [96]<br />

2002 Trans-chromosomal cattle Additional artificial chromosome [15]<br />

2002 Heterozygous knockout in pigs One allele of a-galactosyl-transferase knocked out [34,35]<br />

2003 Homozygous gene knockout in pigs Both alleles of a-galactosyl-transferase knocked out [36]<br />

2003 Transgenic pigs via lentiviral injection Gene transfer into zygotes via lentiviruses [97]<br />

Corresponding author: Heiner Niemann (niemann@tzv.fal.de).<br />

www.sciencedirect.com 0167-7799/$ - see front matter q 2004 Elsevier Ltd. All rights reserved. doi:10.1016/j.tibtech.2004.04.003


Review TRENDS in Biotechnology Vol.22 No.6 June 2004 287<br />

Table 2. Efficiency of somatic cloning of mammals<br />

Species Number of animals a<br />

% Viable offspring Comments<br />

Cattle ,3000 15–20 Up to 30% of the cloned calves showed abnormalities, such as increased birth weight<br />

Sheep ,400 5–8 Same problems as with cattle clones<br />

Goat ,400 3 Minor health problems reported<br />

Mouse ,300 ,2 Some adult mice clones showed obesity and a reduced life span<br />

Pig ,200 ,1 Some cloned piglets had reduced birth weights<br />

Cat 1 ,1<br />

Rabbit 6 ,1<br />

Mule 1 ,1<br />

Horse 1 ,1<br />

Rat 2 ,1<br />

a Estimated total numbers of mammals derived from somatic cloning since 1997.<br />

efficiency is low and few cloned offspring have been born<br />

worldwide (Table 2). Compared with microinjection of<br />

DNA constructs into pronuclei of zygotes, somatic nuclear<br />

transfer is superior for the generation of transgenic<br />

animals (Table 3).<br />

Here, we have summarized the contribution of farm<br />

animals to human health covering (i) the production of<br />

pharmaceuticals; (ii) production of xenografts for overcoming<br />

the severe shortage of human organs and tissues;<br />

(iii) the use of farm animals as disease models; (iv) the<br />

production of dietary supplements or functional foods; and<br />

(v) the contribution of farm animals to solving environmental<br />

problems.<br />

Farm animals for pharmaceutical production<br />

Gene ‘pharming’: production of recombinant human<br />

proteins in the mammary gland of transgenic animals<br />

The conventional production of rare human therapeutic<br />

proteins from blood or tissue extracts is an inefficient,<br />

expensive, labour and time consuming process, which in<br />

addition bears the risk of contamination with human<br />

pathogens. The production of human therapeutic proteins<br />

by recombinant bacteria or cell cultures has alleviated<br />

these problems and has made several therapeutic proteins<br />

available for patients. However, these recombinant systems<br />

have several limitations. They are only suitable for<br />

‘simple’ proteins, the amount of protein produced is<br />

limited, and post-translational modifications are often<br />

incorrect leading to immune reactions against the protein.<br />

In addition, the technical prerequisites are challenging<br />

and production costs are high.<br />

Farm animals such as cattle, sheep, goats, pigs and<br />

even rabbits [3,4] have several significant advantages for<br />

the production of recombinant proteins over other systems,<br />

including their potential for large-scale production,<br />

Table 3. Advantages and disadvantages of two gene transfer methodologies a<br />

correct glycosylation patterns and post-translational<br />

modifications, low running costs, rapid propagation of<br />

the transgenic founders and high expression stability.<br />

These attractive perspectives led to the development of the<br />

‘gene pharming’ concept, which has been advanced to the<br />

level of commercial application [5]. The most promising<br />

site for production of recombinant proteins is the mammary<br />

gland, but other body fluids including blood, urine<br />

and seminal fluid have also been explored [6]. The<br />

mammary gland is the preferred production site mainly<br />

because of the quantities of protein that can be produced<br />

and the ease of extraction or purification of the respective<br />

protein.<br />

Based on the assumption of average expression levels,<br />

daily milk volumes and purification efficiency, 5400 cows<br />

would be needed to produce the 100 000 kg of human<br />

serum albumin (HSA) that are required per year worldwide,<br />

4500 sheep would be required for the production of<br />

5000 kg a-antitrypsin (a-AT), 100 goats for 100 kg of<br />

monoclonal antibodies, 75 goats for the 75 kg of antithrombin<br />

III (ATIII) and two pigs to produce 2 kg human<br />

clotting factor IX. All these values are calculated on a<br />

yearly basis [3].<br />

Large amounts of numerous heterologous recombinant<br />

proteins have been produced by targeting expression to the<br />

mammary gland via mammary gland-specific promoter<br />

elements. Proteins were purified from the milk of<br />

transgenic rabbits, pigs, sheep, goats and cattle. The<br />

biological activity of the recombinant proteins was<br />

assessed and therapeutic effects have been characterized<br />

[3,7]. Products such as ATIII, a-AT or tissue plasminogen<br />

activator (tPA) are advanced to clinical trials (Table 4) [5].<br />

Phase III trials for ATIII have been completed and the<br />

protein is expected to be on the market within the next 2–3<br />

years. In February 2004 an application was submitted to<br />

Microinjection Somatic nuclear transfer<br />

Integration efficiency þ þþþ<br />

Integration site Random Random or targeted<br />

Gene deletion 2 þþþ<br />

Construction size .50 kb (Artificial chromosomes) , 30–50 kb<br />

Technical feasibility Technically demanding Technically demanding<br />

Mosaicism þþþ 2<br />

Expression screen in vitro þ þþþ<br />

Expression pattern Variable Controlled, consistent<br />

Multi-transgenics þ þþþ<br />

a Abbreviations: 2, not possible; þ, weak advantage; þþ, moderate advantage; þþþ, strong advantage<br />

www.sciencedirect.com


288<br />

Table 4. Proteins produced in the mammary gland of transgenic farm animals a<br />

Protein Developmental phase Production species Therapeutic application Potential market introduction date<br />

AT III Phase III Goat Genetic heparin resistance 2005<br />

TPA Phase II/III Goat Dissolving coronary clots . 2006<br />

a-AT Phase II/III Goat and/or sheep Lung emphysema<br />

Cystic fibrosis<br />

. 2007<br />

hFVIII Experimental Sheep Hemophilia A . 2008<br />

HAS Phase I Cattle Blood substitute . 2008<br />

Various antibodies Phase I/II Goat . 2007<br />

a Abbreviations: a-AT, a-A1-antitrypsin; AT III, antithrombin III; hFVIII, human clotting factor VIII; HSA, human serum albumin; TPA, tissue plasminogen activator<br />

the European Market Authorization to allow Atrynw, the<br />

recombinant ATIII from the milk of transgenic dairy goats,<br />

to enter the market as a fully registered drug. The enzyme<br />

a-glucosidase from the milk of transgenic rabbits has<br />

Orphan drug registration and has been successfully used<br />

for the treatment of Pompe’s disease [8]. This is a rare<br />

glycogen storage disorder, which is fatal in children under<br />

2 years and currently application with recombinant<br />

a-glucosidase is the only way to treat this metabolic defect.<br />

Biologically active human lactoferrin has been produced in<br />

large amounts in the mammary glands of transgenic cows<br />

and will probably be developed as a biopharmaceutical for<br />

prophylaxis and treatment of infectious diseases [9].<br />

Guidelines developed by the Food and Drug Administration<br />

(FDA) of the USA require monitoring of the<br />

animals’ health, validation of the gene construct, characterization<br />

of the isolated recombinant protein, as well as<br />

performance of the transgenic animals over several<br />

generations. This has been taken into account when<br />

developing ‘gene pharming’, for example by using only<br />

animals from prion disease-free countries (New Zealand)<br />

and keeping the animals in very hygienic conditions.<br />

Successful drug registration of Atrynw will demonstrate<br />

the usefulness and solidity of this approach and will<br />

accelerate registration of further products from this<br />

process, as well as stimulate research and commercial<br />

activity in this area.<br />

When considering the ‘gene pharming’ concept, one has<br />

to bear in mind that not every protein can be expressed at<br />

the desired levels. Erythropoietin (EPO) could not be<br />

expressed in the mammary gland of transgenic cattle [10]<br />

and was even detrimental to the health of rabbits<br />

transgenic for EPO [11]. We have shown that human<br />

clotting factor VIII (hFVIII) cDNA constructs can be<br />

expressed in the mammary gland of transgenic mice,<br />

rabbits and sheep [1,12]. However, the yields of biologically<br />

active recombinant hFVIII protein from ovine milk were<br />

low because hFVIII was rapidly sequestered into ovine<br />

milk. [13]. These results show that the technology needs<br />

further improvements to achieve high-level expression<br />

with large genes having complex regulation, such as that<br />

coding for hFVIII, although higher levels of hFVIII have<br />

been reported in transgenic swine [14]. With the advent of<br />

transgenic crops that produce pharmacologically active<br />

proteins, there is an array of recombinant technologies<br />

available that will allow the most appropriate production<br />

system for a specific protein to be targeted.<br />

An interesting new development is the generation of<br />

transchromosomal animals (Table 1). A human artificial<br />

www.sciencedirect.com<br />

Review TRENDS in Biotechnology Vol.22 No.6 June 2004<br />

chromosome (HAC) containing the entire sequences of<br />

the human immunoglobulin heavy and light chain loci<br />

has been introduced into bovine fibroblasts, which<br />

were then used in nuclear transfer. Transchromosomal<br />

offspring were obtained that expressed human<br />

immunoglobulin in their blood. This system could be<br />

a significant step forward in the production of human<br />

therapeutic polyclonal antibodies [15]. Further studies<br />

will show whether the additional chromosome will be<br />

maintained over future generations and how stable<br />

expression will be.<br />

Production of a new class of antibiotics: cationic antimicrobial<br />

peptides<br />

With increasing antibiotic resistance in bacterial species,<br />

there is a growing need to develop new classes of antimicrobial<br />

agents. Cationic anti-microbial peptides (AMP)<br />

have many of the desired features [16] because they<br />

possess a broad spectrum of activity, kill gram-positive and<br />

gram-negative bacteria rapidly, are unaffected by classical<br />

resistance genes and are active in animal models [17–19].<br />

AMPs belong to the innate immune defense, which acts as<br />

a first barrier ahead of humoral and cellular immune<br />

systems, and neutralizes bacteria by interacting specifically<br />

with their cell membranes. Their low transmembrane<br />

potential of ,-100 mV, and the abundant anionic<br />

phospholipids are essential for this selective interaction. It<br />

is proposed that AMPs physically disintegrate the cell<br />

membrane, and in addition can interact with several<br />

intracellular target molecules [16]. More than 500 such<br />

peptides have been discovered in plants, insects, invertebrates,<br />

fish, amphibians, birds and mammals [16–20].<br />

AMPs from livestock species would be superior antimicrobial<br />

drugs because they would lack cytotoxic effects<br />

that were found for insect peptides; the evolution of<br />

resistance would not affect the human specific innate<br />

immunity [20].<br />

Prominent examples of cationic anti-microbial peptides<br />

from farm animals already in advanced clinical trials<br />

(Phases II–III) are Iseganan (derived from protegrin-1<br />

peptide of pig leucocytes; Intrabiotics; http://intrabiotics.<br />

com/) and MBI-594 (similar to indolicidin peptide from<br />

bovine neutrophils; Micrologix, http://mbiotech.com/). To<br />

date, there are no documented cases of antimicrobial<br />

peptide-resistance for AMPs and combinatorial<br />

approaches in peptides (20 possible amino acids in each<br />

position) provide great potential for rational drug design<br />

[21]. Recombinant production will keep the production<br />

costs low.


Xenotransplantation of porcine organs to human<br />

patients<br />

Solid organs<br />

Today .250 000 people are alive only because of the<br />

successful transplantation of an appropriate human organ<br />

(allotransplantation). On average, 75–90% of patients<br />

survive the first year after transplantation and the<br />

average survival of a patient with a transplanted heart,<br />

liver or kidney is 10–15 years. This progress in organ<br />

transplantation technology has led to an acute shortage of<br />

appropriate organs, and cadaveric or live organ donation<br />

cannot cover the demand in western societies. The 2001<br />

figures from the United Network for Organ Sharing (www.<br />

unos.org) in the USA show that the ratio of patients with<br />

organ transplantation to those on the waiting list is ,1:4<br />

(Table 5). A similar ratio is found in other countries such as<br />

the UK, France and Germany. A new person is added to the<br />

waiting list every 14 minutes. This has led to the sad and<br />

ethically challenging situation in which several thousand<br />

patients who could have survived if appropriate organs<br />

had been available die every year.<br />

To close the growing gap between demand and<br />

availability of appropriate organs, porcine xenografts are<br />

considered the solution of choice [22,23]. Today the<br />

domesticated pig is considered the optimal donor animal<br />

because (i) the organs are similar in size to human organs;<br />

(ii) porcine anatomy and physiology are not too different<br />

from that of humans; (iii) pigs have short reproduction<br />

cycles and large litters; (iv) pigs grow rapidly;<br />

(v) maintenance of high hygienic standards is possible at<br />

relatively low costs; and (vi) transgenic techniques for<br />

modifying the immunogenicity of porcine cells and organs<br />

are well established.<br />

The process of generating and evaluating transgenic<br />

pigs as potential donors for xenotransplants involves a<br />

variety of complex steps and is time-, labour- and resourceintensive.<br />

Essential prerequisites for successful xenotransplantation<br />

are:<br />

(i) Overcoming the immunological hurdles.<br />

(ii) Prevention of transmission of pathogens from the<br />

donor animal to the human recipient.<br />

(iii) Compatibility of the donor organs with the human<br />

organ in terms of anatomy and physiology.<br />

The immunological obstacles in a porcine-to-human<br />

xenotransplantation are the hyperacute rejection<br />

response (HAR), acute vascular rejection (AVR), cellular<br />

rejection and potentially chronic rejection [24]. The HAR<br />

Table 5. Overview of transplanted organs and demand for<br />

organ transplantation (USA) a<br />

Type of transplant Transplantations in 2001 Waiting patients<br />

Kidney 14 152 52 772<br />

Liver 5177 17 520<br />

Pancreas 468 1318<br />

Kidney/Pancreas 884 2520<br />

Heart 2202 4163<br />

Heart/Lung 27 209<br />

Lung 1054 3799<br />

Total 23 964 79 845<br />

a Data from United Network for Organ Sharing, 2001 (www.unos.org).<br />

www.sciencedirect.com<br />

Review TRENDS in Biotechnology Vol.22 No.6 June 2004 289<br />

occurs within seconds or minutes. In the case of a<br />

discordant organ (e.g. in transplanting from pig to<br />

human) naturally occurring antibodies react with antigenic<br />

structures on the surface of the porcine organ and<br />

induce HAR by activating the complement cascade via the<br />

antigen–antibody complex. Ultimately, this results in the<br />

formation of the membrane attack complex (MAC).<br />

However, the complement cascade can be shut down at<br />

various points by expression of regulatory genes that<br />

prevent the formation of the MAC. Regulators of the<br />

complement cascade are CD55 (decay accelerating factor,<br />

DAF), CD46 (membrane cofactor protein, MCP) or CD59.<br />

MAC disrupts the endothelial cell layer of the blood<br />

vessels, which leads to lysis, thrombosis, loss of vascular<br />

integrity and ultimately to rejection of the transplanted<br />

organ [24].<br />

Induced xenoreactive antibodies are thought to be<br />

responsible for AVR, which occurs within days of a<br />

transplantation of a xenograft; disseminated intravascular<br />

coagulation (DIC) is a predominant feature of AVR.<br />

Despite severe immunosuppressive treatment, a disturbed<br />

thrombocyte function and DIC were observed in a pig-toprimate<br />

xenotransplant model [25,26]. The endothelial<br />

cells of the graft’s microvasculature loose their antithrombic<br />

properties, attract leucocytes, monocytes and<br />

platelets leading to anemia and organ failure. The<br />

underlying mechanism for DIC and thrombotic microangiopathy<br />

is thought to be activation of the endothelial<br />

cells attributed to incompatibilities between human and<br />

porcine coagulation factors [25]. At least three incompatibilities<br />

between human and porcine coagulation systems<br />

have been identified; the first is the failure of porcine<br />

thrombomodulin (TM) to activate human anticoagulant<br />

protein C, the second is that the porcine tissue factor<br />

pathway inhibitor fails to inhibit human clotting factor Xa<br />

and the third is that porcine von Willebrand factor (vWF)<br />

binds and activates human platelets [26]. Human thrombomodulin<br />

(hTM) and heme-oxygenase 1 (hHO-1) are<br />

crucially involved in the etiology of DIC and might be good<br />

targets for future transgenic studies to improve long-term<br />

survival of porcine xenografts by creating multi-transgenic<br />

pigs.<br />

The cellular rejection occurs within weeks after<br />

transplantation. In this process the blood vessels of the<br />

transplanted organ are damaged by T-cells, which invade<br />

the intercellular spaces and destroy the organ. This<br />

rejection is observed after allotransplantation and is<br />

normally suppressed by life-long administration of immunosuppressive<br />

drugs [24].<br />

When using a discordant donor species such as the pig,<br />

overcoming the HAR and AVR are the preeminent goals.<br />

The most promising strategy for overcoming the HAR is<br />

the synthesis of human complement regulatory proteins<br />

(RCAs) in transgenic pigs [22,23,27,28]. Following transplantation,<br />

the porcine organ would produce the complement<br />

regulatory protein and can thus prevent the<br />

complement attack of the recipient. Pigs transgenic for<br />

DAF or MCP have been generated by microinjection of<br />

DNA constructs into pronuclei of zygotes. Hearts and<br />

kidneys from these animals have been transplanted either<br />

heterotopically, (in addition to the recipient’s own organ) or


290<br />

Table 6. Success rates of RCA-transgenic porcine organs after transplantation to primate recipients a<br />

RCA Organ/kind of transplant Recipient Immuno-suppression Survival (days)<br />

hDAF Heart/heterotropic Cynomologus þþþ , 135<br />

hDAF Heterotopic Cynomologus þþ , 90<br />

hDAF Orthopic Baboon þþþ , 28<br />

hDAF Kidney/orthotopic Cynomologus þþ , 90<br />

hCD59 Kidney/orthotopic, heterotopic Cynomologus þþþ , 20<br />

hCD46 Heart, heterotopic Baboon þþ , 23<br />

a Abbreviations: þ, weak immunosuppression; þþ, moderate immunosuppression; þþþ, heavy immunosuppression; RCA, regulator of complement activity<br />

orthotopically (life supportive) into non-human primates.<br />

Survival rates reached 23–135 days with porcine xenografts<br />

transgenic for one of the two complement regulators;<br />

survival rates were heavily affected by the strength of<br />

the immune-suppressive protocol (Table 6) [29–31].<br />

Similarly, transgenic expression of hCD59 was compatible<br />

with an extended survival of porcine hearts in a perfusion<br />

model or following transfer into primates [32,33]. These<br />

data show that HAR can be overcome in a clinically<br />

acceptable manner by expressing human complement<br />

regulators in transgenic pigs [22].<br />

Another promising strategy towards successful xenotransplantation<br />

is the knockout of the antigenic structures<br />

on the surface of the porcine organ that cause HAR. These<br />

structures are known as 1,3-a-gal-epitopes and are<br />

primarily produced by activity of 1,3-a-galactosyltransferase<br />

(a-gal). Piglets in which one allele of a-gal locus had<br />

been knocked out by homologous recombination in<br />

primary donor cells that were employed in nuclear<br />

transfer were recently generated [34,35]. The birth of<br />

four healthy piglets with disruption of both allelic loci for<br />

a-gal has meanwhile also been published. Applying toxin A<br />

from Clostridium difficile to cells that already carried one<br />

deleted a-gal allele selected a cell clone, which carried an<br />

inactivating point mutation on the second allele. This cell<br />

clone was then used in nuclear transfer [36]. The<br />

usefulness of these animals for xenotransplantation has<br />

recently been reported [37].<br />

Further improvements in the success of xenotransplantation<br />

will arise from the possibility of inducing a<br />

permanent tolerance across xenogenic barriers [38,39]. A<br />

particularly promising strategy for long-term graft acceptance<br />

is the induction of a permanent chimerism via<br />

intraportal injection of embryonic stem cells [40].<br />

Prevention of transmission of zoonoses from the donor<br />

animal to the human recipient is crucial for clinical<br />

application of porcine xenografts. This aspect gained<br />

particular significance when a few years ago it was<br />

shown that porcine endogenous retroviruses (PERV) can<br />

be produced by porcine cell lines and can even infect<br />

human cell lines in vitro [41]. However, until today no<br />

infection has been found in patients that had received<br />

various forms of living porcine tissues (e.g. islet cells,<br />

insulin, skin, extracorporal liver) for up to 12 years [42].<br />

Recent intensive research has shown that porcine<br />

endogenous retroviruses probably do not present a risk<br />

for recipients of xenotransplants provided all necessary<br />

precautions are taken [43–46]. In addition, a strain of<br />

miniature pigs has been identified that does not produce<br />

infective PERV [47]. Although xenotransplantation poses<br />

numerous further challenges to research, it is expected<br />

www.sciencedirect.com<br />

Review TRENDS in Biotechnology Vol.22 No.6 June 2004<br />

that transgenic pigs will be available as organ donors<br />

within the next five to ten years. Guidelines for the clinical<br />

application of porcine xenotransplants are already available<br />

in the USA and are currently <strong>bei</strong>ng developed in<br />

several other countries.<br />

The ethical challenges of xenotransplantation have<br />

been a matter of a worldwide intensive debate. A general<br />

consensus has been reached that the technology is<br />

ethically acceptable provided the individual well-<strong>bei</strong>ng<br />

does not compromise public health by producing and<br />

transmitting new pathogens. Economically xenotransplantation<br />

might be viable if the enormous costs caused<br />

by patients suffering from severe kidney disease, needing<br />

dialysis or those suffering from chronic heart diseases<br />

could be avoided by a functional kidney or heart xenograft.<br />

Xenogenic cells and tissue<br />

Another promising area of application for transgenic<br />

animals will be the supply of xenogenic cells and tissue.<br />

Several intractable diseases, disorders and injuries are<br />

associated with irreversible cell death and/or aberrant<br />

cellular function. Despite numerous attempts, primary<br />

human cells cannot yet be expanded well enough in<br />

culture. In the future, human embryonic stem cells could<br />

be a source for specific differentiated cell types that can be<br />

used in cell therapy [48,49]. Xenogenic cells, in particular<br />

from the pig, hold great promise for successful cell<br />

therapies for human patients because cells can be<br />

implanted at the optimal therapeutic location (i.e. immunoprivileged<br />

sites, such as the brain), genetically or<br />

otherwise modified before transplantation to enhance<br />

cell function, banked and cryopreserved, or combined<br />

with different cell types in the same graft [50].<br />

Xenogenic cell therapy has been advanced to preclinical<br />

studies. Porcine islet cells have been transplanted to<br />

diabetic patients and were shown to be at least partially<br />

functional over a limited period of time [51]. Porcine fetal<br />

neural cells were transplanted into the brain of<br />

patients suffering from Parkinson’s and Huntington’s<br />

disease [52,53]. In a single autopsied patient the graft<br />

survived for .7 months and the transplanted cells formed<br />

dopaminergic neurons and glial cells. Pig neurons<br />

extended axons from the graft site into the host brain<br />

[52]. Further examples for the potential use of porcine<br />

neural cells are in cases of stroke and focal epilepsy [54].<br />

Olfactory ensheathing cells (OECs) or Schwann cells<br />

derived from hCD59 transgenic pigs promoted axonal<br />

regeneration in rat spinal cord lesion [55]. Thus, cells from<br />

genetically modified pigs could restore electrophysiologically<br />

functional axons across the site of a spinal cord<br />

transsection. Xenogenic porcine cells could also be useful


as a novel therapy for liver diseases. On transplantation of<br />

porcine hepatocytes to Watanabe heritable hyperlipidemic<br />

(WHHL) rabbits (a model for familial hypercholesterolemia)<br />

the xenogenic cells migrated out of the vessels and<br />

integrated into the hepatic parenchyma. The integrated<br />

porcine hepatocytes provided functional low density<br />

lipoprotein (LDL) receptors and thus reduced cholesterol<br />

levels by 30–60% for at least 100 days [56].<br />

A clone of bovine adrenocortical cells restored adrenal<br />

function upon transplantation to adrenalectomized severe<br />

combined immunodeficient (SCID) mice indicating that<br />

functional endocrine tissue can be derived from a single<br />

somatic cell [57]. Bovine neuronal cells were collected from<br />

transgenic fetuses, and when transplanted into the brain<br />

of rats resulted in significant improvements in symptoms<br />

of Parkinson’s disease [58]. Furthermore, xenotransplantation<br />

of retinal pigment epithelial cells holds promise for<br />

treating retinal diseases such as macular degeneration,<br />

which is associated with photoreceptor losses. Porcine or<br />

bovine fetal cardiomyocytes or myoblasts might provide a<br />

therapeutic approach for the treatment of ischemic heart<br />

disease. Similarly, xenogenic porcine cells might be<br />

valuable for the repair of skin or cartilage damage [50].<br />

In light of the emergence of significantly improved<br />

protocols for genetic modification of donor animals and<br />

new powerful immunosuppressive drugs xenogenic cell<br />

therapy will evolve as an important therapeutic option for<br />

the treatment of human diseases.<br />

Farm animals as models for human diseases<br />

In mouse genetics the generation of knockout animals is a<br />

standard procedure and several thousand strains carrying<br />

gene knockouts or transgenes have been developed [Mouse<br />

Knockout and Mutation Database (MKMD); http://<br />

research.bmn.com/mkmd]. Models have been developed<br />

for several human diseases. However, mouse physiology,<br />

anatomy and life span differ significantly from those in<br />

humans, making the rodent model inappropriate for<br />

several human diseases. Farm animals, such as pigs,<br />

sheep or even cattle could be more appropriate models to<br />

study human diseases in particular non-insulin-dependent<br />

diabetes, cancer and neurodegenerative disorders,<br />

which require longer observation periods than those<br />

possible in mice [59–61]. With the aid of the microinjection<br />

technology an important pig model for the rare human eye<br />

disease Retinitis pigmentosa (PR) has been developed [62].<br />

Patients with PR develop night blindness early in life<br />

attributed to a loss of photoreceptors. The transgenic pigs<br />

express a mutated rhodopsin gene and show a great<br />

similarity with the human phenotype. Treatment models<br />

with value for human patients are <strong>bei</strong>ng developed [63].<br />

The development of the somatic cloning technology and<br />

the merger with targeted genetic modifications and<br />

conditional gene expression will enhance the possibilities<br />

for creating useful models for human diseases in large<br />

animals. A good example is the knockout of the prion gene<br />

that would make sheep and cattle non-susceptible to<br />

spongiform encephalopathies (scrapie and BSE). Mice<br />

models showed that the knockout of the prion protein is the<br />

only secure way to prevent infection and transmission of<br />

the disease [64]. The first successful targeting of the ovine<br />

www.sciencedirect.com<br />

Review TRENDS in Biotechnology Vol.22 No.6 June 2004 291<br />

prion locus has been reported; however the cloned lambs<br />

carrying the kockout locus died several days after birth<br />

[65]. Prion knockout animals could be an appropriate<br />

model for studying the epidemiology of spongiform<br />

encephalopathies in humans and are crucial for<br />

developing strategies to eliminate prion carriers from a<br />

farm animal population.<br />

The pig could be a useful model to study defects of<br />

growth hormone releasing hormone (GHRH), which is<br />

characterized by a variety of conditions such as Turner<br />

syndrome, hypochrondroplasia, Crohn’s disease, intrauterine<br />

growth retardation or renal insufficiency. Application<br />

of recombinant GHRH in an injectable form and its<br />

myogenic expression has been shown to alleviate these<br />

problems in a porcine model [66].<br />

An important aspect of large animal models for human<br />

diseases is the recent finding that somatic cloning per se<br />

does not result in shortening of the telomeres. Telomeres<br />

are highly repetitive DNA sequences at the end of the<br />

chromosomes that are crucial for their structural integrity<br />

and function and are thought to be related to lifespan.<br />

Telomere shortening is usually correlated with severe<br />

limitations of the regenerative capacity of cells, the onset<br />

of cancer, ageing and chronic disease with significant<br />

impact on human lifespan [67–70]. Expression of the<br />

enzyme telomerase, which is primarily responsible for the<br />

formation and rebuilding of telomeres, is suppressed in<br />

most somatic tissues postnatally. Although telomeres in<br />

cloned sheep (Dolly) derived from epithelial cells were<br />

shorter than those of naturally bred age matched control<br />

animals, telomere lengths in cloned mice and cattle were<br />

not different from those determined in age matched<br />

controls even when senescent donor cells had been<br />

employed [71–75]. Recent studies in our laboratory have<br />

revealed that telomere length is established already early<br />

in preimplantation development by a specific genetic<br />

programme and is dependent on telomerase activity [76].<br />

Dietary modifications of animal products<br />

Application of gene and biotechnology for nutrition and<br />

biomedicine is more developed for plants than farm<br />

animals [77]. The term nutriceuticals in the farm animal<br />

context means that gene and biotechnology are used to<br />

enhance farm animal products to improve diet and have<br />

concomitant medical applications. Because these products<br />

have a proven pharmacological effect on the body, they are<br />

regarded as ‘drugs’ and must be tested for efficacy and<br />

safety. Functional foods are those designed to provide<br />

specific and beneficial physiological effects on human<br />

health and welfare and should prevent diet-related<br />

diseases. Functional foods from animals could be used to<br />

lower cholesterol levels in an effort to battle cardiovascular<br />

diseases, to reduce high blood pressure by adding<br />

angiotensin converting enzyme (ACE) inhibitors or to<br />

increase immunity by adding specific immunostimulatory<br />

peptides [78]. However, data showing that nutriceuticals<br />

are really beneficial for human health are rare.<br />

Regarding the production of improved quality animal<br />

products, interesting observations have been made in<br />

several beef cattle breeds, such as Belgian Blue and<br />

Piedmontese. These breeds were accidentally bred for


292<br />

mutations of the myostatin gene, which renders it nonfunctional<br />

or less functional than the wild-type gene [79].<br />

The mutated genes cause muscle hypertrophy and led to<br />

improved meat quality. This observation makes targeted<br />

modifications of the myostatin gene an interesting option<br />

for the meat industry. A diet rich in non-saturated fatty<br />

acids is correlated with a reduced risk of stroke and<br />

coronary diseases. One transgenic approach to this is the<br />

generation of pigs with increased amounts of nonsaturated<br />

fatty acids. Pigs producing a higher ratio of<br />

unsaturated versus saturated fatty acids in their muscles<br />

are currently under development in Japan.<br />

An attractive example for targeted genetic modification<br />

could be dairy production [80,81]. Apart from conventional<br />

dairy products, it could be possible to produce fat-reduced<br />

or even fat-free milk or milk with a modified lipid<br />

composition via modulation of enzymes involved in lipid<br />

metabolism; to increase curd and cheese production by<br />

enhancing expression of the casein gene family in the<br />

mammary gland; to create ‘hypoallergenic’ milk by knockout<br />

of the b-lactoglobulin gene; to generate lactose-free<br />

milk via knockout of the a-lactalbumin locus that is the<br />

key molecule in milk sugar synthesis; to produce ‘infant<br />

milk’ in which human lactoferrin is abundantly available<br />

or to produce milk with a highly improved hygienic<br />

standard via an increased level of lysozyme or other<br />

anti-microbiological substances in the udder. Lactosereduced<br />

or lactose-free milk could make dairy products<br />

suitable for consumption by a large proportion of the<br />

world’s population who do not possess an active lactase<br />

enzyme in their gut system. However, one has to bear in<br />

mind that lactose is the main osmotically active substance<br />

in milk and a lack thereof could interfere with milk<br />

secretion. A lactase construct has been expressed in the<br />

mammary gland of transgenic mice and reduced lactose<br />

contents by 50–85% without altering milk secretion [82].<br />

However, mice with a homozygous knockout for alactalbumin<br />

could not feed their offspring because of the<br />

high viscosity of the milk [83]. These diverging findings<br />

demonstrate the feasibility of obtaining significant alterations<br />

of milk composition by applying the appropriate<br />

strategy.<br />

The physicochemical properties of milk are mainly<br />

affected by the ratio of casein variants. Therefore, casein is<br />

a prime target for the improvement of milk composition.<br />

Mouse models have been developed for most of the above<br />

modifications indicating the feasibility of obtaining significant<br />

alterations in milk composition but at the same<br />

time showing that unwanted side effects cannot be ruled<br />

out [83,84]. Only one full-scale study in livestock has been<br />

reported as yet [85]. The recent report showed that the<br />

casein ratio can be altered by overexpression of b- and<br />

k-casein in cattle clearly underpinning the potential for<br />

improvements in the functional properties of bovine<br />

milk [85].<br />

Towards environmentally friendly farm animals<br />

Phosphorus pollution by animal production is a serious<br />

problem in agriculture and excess phosphate from manure<br />

promotes eutrophication. Phytase transgenic pigs have<br />

been developed to address the problem of manure-based<br />

www.sciencedirect.com<br />

Review TRENDS in Biotechnology Vol.22 No.6 June 2004<br />

environmental pollution. These pigs carry a bacterial<br />

phytase gene under the transcriptional control of a<br />

salivary gland specific promoter, which allows the pigs to<br />

digest plant phytate. Without the bacterial enzyme, the<br />

phytate phosphorus passes undigested into manure and<br />

pollutes the environment. With the bacterial enzyme, the<br />

fecal phosphorus output was reduced up to 75% [86]. These<br />

environmentally friendly pigs are expected to enter the<br />

commercial production chains within the next few years.<br />

Precautions and perspectives<br />

Throughout mankind’s history farm animals have made<br />

significant contributions to human health and well-<strong>bei</strong>ng.<br />

The convergence of the recent advances in reproductive<br />

technologies with the tools of molecular biology opens a<br />

new dimension for this area [1]. Major prerequisites will be<br />

the continuous refinement of reproductive biotechnologies<br />

and a rapid completion of livestock genome sequencing<br />

and annotation. The technology developed in the deciphering<br />

of the human genome will improve and accelerate<br />

sequencing of genomes from livestock [87]. We anticipate<br />

genetically modified animals will play a significant role in<br />

the biomedical arena, in particular via the production of<br />

valuable pharmaceutical proteins and the derivation of<br />

xenografts, within the next 5–7 years. Agricultural<br />

application might be further away (.10 years) given the<br />

complexity of some of the economically important traits<br />

and the public skepticism of genetic modification related to<br />

food production [1].<br />

A crucial aspect of animal-derived products is the<br />

prevention of transmission of pathogens from animals to<br />

humans. This requires sensitive and reliable diagnostic<br />

and screening methods for the various types of pathogenic<br />

organisms. The recent findings (see above) that the risk of<br />

PERV transmission is negligible are promising and show<br />

that with targeted and intensive research such important<br />

questions can be answered within a limited period of time,<br />

paving the way for preclinical testing of xenografts.<br />

Furthermore, it should be kept in mind that the biomedical<br />

applications of farm animals will require strict standards<br />

of ‘genetic security’ and reliable and sensitive methods for<br />

the molecular characterization of the products. A major<br />

contribution towards the goal of well-defined products will<br />

come from array technology (cDNA, peptide or protein<br />

arrays), which establishes ‘fingerprint’ profiles at the<br />

transcriptional and/or protein level [88,89]. Meanwhile<br />

improvements of RNA isolation and unbiased amplification<br />

of tiny amounts of mRNA (picogram) enable researchers<br />

to analyse RNA from single embryos [90]. With the aid<br />

of this technology one can gain in-depth insight into the<br />

proper functioning of a transgenic organism and thereby<br />

ensure the absence of unwanted side effects [88,89]. This<br />

would also be required to maintain the highest possible<br />

levels of animal welfare in cases of genetic modification.<br />

References<br />

1 Niemann, H. and Kues, W.A. (2003) Application of transgenesis on<br />

livestock for agriculture and biomedicine. Anim. Reprod. Sci. 79,<br />

291–317<br />

2 Hammer, R.E. et al. (1985) Production of transgenic rabbits, sheep and<br />

pigs by microinjection. Nature 315, 680–683


3 Rudolph, N.S. (1999) Biopharmaceutical production in transgenic<br />

livestock. Trends Biotechnol. 17, 367–374<br />

4 Bösze, Zs. et al. (2003) The transgenic rabbit as model for human<br />

diseases and as a source of biologically active recombinant proteins.<br />

Transgenic Res. 12, 541–553<br />

5 Ziomek, C.A. (1998) Commercialization of proteins produced in the<br />

mammary gland. Theriogenology 49, 139–144<br />

6 Dyck, M.K. et al. (2003) Making recombinant proteins in animals –<br />

different systems, different applications. Trends Biotechnol. 21,<br />

394–399<br />

7 Meade, H.M. et al. (1999) Expression of recombinant proteins in the<br />

milk of transgenic animals. In: Gene Expression Systems. Fernandez,<br />

J.M., Hoeffler, J.P., (eds.) Academic Press, San Diego, U. S. A., pp. 399–<br />

427<br />

8 Van der Hout, J.M.P. et al. (2001) Enzyme therapy for Pompe disease<br />

with recombinant human a-glucosidase from rabbit milk. J. Inherit.<br />

Metab. Dis. 24, 266–274<br />

9 van Berkel, P.H. et al. (2002) Large scale production of recombinant<br />

human lactoferrin in the mik of trangenic cows. Nat. Biotechnol. 20,<br />

484–487<br />

10 Hyttinen, J.M. et al. (1994) Generation of transgenic dairy cattle from<br />

transgene-analyzed and sexed embryos produced in vitro. Biotechnology<br />

(N. Y.) 12, 606–608<br />

11 Massoud, M. et al. (1996) The deleterious effects of human erythropoietin<br />

gene driven by the rabbit whey acidic protein gene promoter in<br />

transgenic rabbits. Reprod. Nutr. Dev. 36, 555–563<br />

12 Hiripi, L. et al. (2003) Expression of active human blood clotting factor<br />

VIII in the mammary gland of transgenic rabbits. DNA Cell Biol. 22,<br />

21–25<br />

13 Niemann, H. et al. (1999) Expression of human blood clotting factor<br />

VIII in the mammary gland of transgenic sheep. Transgenic Res. 8,<br />

237–247<br />

14 Paleyanda, R.K. et al. (1997) Transgenic pigs produce functional<br />

human factor VIII in milk. Nat. Biotechnol. 15, 971–975<br />

15 Kuroiwa, Y. et al. (2002) Cloned transchromosomic calves producing<br />

human immunoglobulin. Nat. Biotechnol. 20, 889–894<br />

16 Hancock, R.E.E. and Scott, M.G. (2000) The role of antimicrobial<br />

peptides in animal defense. Proc. Natl. Acad. Sci. U. S. A. 97,<br />

8856–8861<br />

17 Goldstein, B.P. (1998) Activity of nisin against Streptococcus pneunoniae,<br />

in vitro, and in a mouse infection model. J. Antimicrob.<br />

Chemother. 42, 277–278<br />

18 Gamelli, R. et al. (1998) Improvement in survival with peptidyl<br />

membraned interactive molecule D4B treatment after burn wound<br />

infection. Arch. Surg. 133, 715–720<br />

19 Kirikae, T. et al. (1998) Protective effects of a human 18-kilodalton<br />

cationic antimicrobial protein (CAP18)-derived peptide against murine<br />

endotoxemia. Infect. Immun. 66, 1861–1868<br />

20 Zhang, G. et al. (2001) Porcine antimicrobial peptided: new prospects<br />

for ancient molecules of host defense. Vet. Res. 31, 277–296<br />

21 Lee, K.H. (2002) Development of short antimicrobial peptides derived<br />

from host defense peptides or by combinatorial libraries. Curr. Pharm.<br />

Des. 8, 795–813<br />

22 Bach, F.H. (1998) Xenotransplantation: Problems and prospects.<br />

Annu. Rev. Med. 49, 301–310<br />

23 Platt, J.L. and Lin, S.S. (1998) The future promises of xenotransplantation.<br />

In: Xenotransplantation. Ann. NY Acad. Sci. 862, 5–18.<br />

24 White, D. (1996) Alteration of complement activity: a strategy for<br />

xenotransplantation. Trends Biotechnol. 14, 3–5<br />

25 Cowan, P.J. et al. (2000) Renal xenografts from triple-transgenic pigs<br />

are not hyperacutely rejected but cause coagulopathy in nonimmunosuppressed<br />

baboons. Transplantation 69, 2504–2515<br />

26 Gaca, J.G. et al. (2002) The role of the porcine von Willebrand factor:<br />

Baboon platelet interactions in pulmonary xenotransplantation.<br />

Transplantation 74, 1596–1603<br />

27 Cozzi, E. and White, D.J.G. (1995) The generation of transgenic pigs as<br />

potential organ donors for humans. Nat. Med. 1, 964–966<br />

28 White, D. (1996) Alteration of complement activity: a strategy for<br />

xenotransplantation Trends. Biotechnol. 14, 3–5<br />

29 Cozzi, E. et al. (2000) Progress in xenotransplantation. Clin. Nephrol.<br />

53, 13–18<br />

30 Diamond, L.E. et al. (2001) A human CD46 transgenic pig model<br />

www.sciencedirect.com<br />

Review TRENDS in Biotechnology Vol.22 No.6 June 2004 293<br />

system for the study of discordant xenotransplantation. Transplantation<br />

71, 132–142<br />

31 Zaidi, A. et al. (1998) Life-supporting pig to primate renal xenotransplantation<br />

using genetically modified donors. Transplantation<br />

65, 1584–1590<br />

32 Fodor, W.L. et al. (1994) Expression of a functional human complement<br />

inhibitor in a transgenic pig as a model for the prevention of<br />

xenogeneic hyperacute organ rejection. Proc. Natl. Acad. Sci. U. S. A.<br />

91, 11153 – 11157<br />

33 Niemann, H. et al. (2001) CMV early promoter induced expression of<br />

hCD59 in porcine organs provides protection against hyperacute<br />

rejection. Transplantation 72, 1898–1906<br />

34 Dai, Y. et al. (2002) Targeted disruption of the alpha1,3-galactosyltransferase<br />

gene in cloned pigs. Nat. Biotechnol. 20, 251–255<br />

35 Lai, L. et al. (2002) Production of alpha-1,3-galactosyltransferase<br />

knockout pigs by nuclear transfer cloning. Science 295, 1089–1092<br />

36 Phelps, C.J. et al. (2003) Production of alpha 1,3-galactosyltransferasedeficient<br />

pigs. Science 299, 411–414<br />

37 Yamada, K. et al. (2003) An initial report of alpha-gal deficient pig-tobaboon<br />

renal xenotransplantation: evidence for the benefit of cotransplanting<br />

vascularized donor thymic tissue. Xenotransplantation<br />

10, 480<br />

38 Greenstein, J. and Sachs, D.H. (1997) The use of tolerance for<br />

transplantation across xenogeneic barriers. Nat. Biotechnol. 15,<br />

235–238<br />

39 Auchincloss, H. Jr and Sachs, D.H. (1998)) Xenogenic transplantation.<br />

Annu. Rev. Immunol. 16, 433–470<br />

40 Fändrich, F. et al. (2002) Preimplantation-stage stem cells induce longterm<br />

allogeneic graft acceptance without supplementary host conditioning.<br />

Nat. Med. 8, 171–177<br />

41 Patience, C. et al. (1997) Infection of human cells by an endogenous<br />

retrovirus of pigs. Nat. Med. 3, 282–286<br />

42 Paradis, K. et al. (1999) Search for cross-species transmission of<br />

porcine endogenous retrovirus in patients treated with living pig<br />

tissue. Science 285, 1236–1241<br />

43 Patience, C. et al. (1998) No evidence of pig DNA of retroviral infection<br />

in patients with short-term extracorporal connection of pig kidneys.<br />

Lancet 352, 699–701<br />

44 Dinsmore, L.E. et al. (2000) No evidence for infection of human cells<br />

with porcine endogenous retrovirus (PERV) after exposure to porcine<br />

fetal neuronal cells. Transplantation 70, 1382–1389<br />

45 Switzer, W.M. et al. (2001) Lack of cross-species transmission of<br />

porcine endogenous retrovirus infection to nonhuman primate<br />

recipients of porcine cells, tissues and organs. Transplantation 71,<br />

959–965<br />

46 Martin, U. et al. (2002) Absence of PERV specific humoral immune<br />

response in baboons after transplantation of porcine cells of organs.<br />

Transplant. Int. 15, 361–368<br />

47 Oldmixon, B.A. et al. (2002) Porcine endogenous retrovirus transmission<br />

characteristics of an inbred herd of miniature swine. J. Virol.<br />

76, 3045–3048<br />

48 Rossant, J. (2001) Stem cells from the mammalian blastocyst. Stem<br />

Cells 19, 477–482<br />

49 Reubinoff, B.E. et al. (2000) Embryonic stem cell lines from human<br />

blastocysts: somatic differentiation in vitro. Nat. Biotechnol. 18,<br />

399–404<br />

50 Edge, A.S.B. et al. (1998) Xenogeneic cell therapy: Current progress<br />

and future developments in porcine cell transplantation. Cell<br />

Transplant. 7, 525–539<br />

51 Groth, C.G. et al. (1994) Transplantation of porcine fetal pancreas to<br />

diabetic patients. Lancet 344, 1402–1404<br />

52 Deacon, T. et al. (1997) Histological evidence of fetal pig neutral cell<br />

survival after transplantation into a patient with Parkinson’s disease.<br />

Nat. Med. 3, 350–353<br />

53 Fink, J.S. et al. (2000) Porcine xenografts in Parkinson’s disease and<br />

Huntington’s disease patients: preliminary results. Cell Transplant. 9,<br />

273–278<br />

54 Björklund, A. (1991) Neural transplantation – An experimental tool<br />

with clinical possibilities. Trends Neurosci. 14, 319–322<br />

55 Imaizumi, T. et al. (2000) Xenotransplantation of transgenic pig<br />

olfactory ensheathing cells promotes axonal regeneration in rat spinal<br />

cord. Nat. Biotechnol. 18, 949–953<br />

56 Gunsalus, J.R. et al. (1997) Reduction of serum cholesterol in


294<br />

Review TRENDS in Biotechnology Vol.22 No.6 June 2004<br />

Watanabe rabbits by xenogeneic hepatocellular transplantation. Nat.<br />

Med. 3, 48–53<br />

57 Thomas, M. et al. (1997) Adrenocortical tissue formed by transplantation<br />

of normal clones of bovine adrenocortical cells in scid mice<br />

replaces the essential functions of the animals’ adrenal glands. Nat.<br />

Med. 3, 978–983<br />

58 Zawada, W.M. et al. (1998) Somatic cell cloned transgenic bovine<br />

neurons for transplantation in parkinson rats. Nat. Med. 4, 569–574<br />

59 Milan, D. et al. (2000) A mutation in PRKAG3 associated with excess<br />

glycogen content in pig skeletal muscle. Science 288, 1248–1251<br />

60 Palmarini, M. and Fan, H. (2001) Retrovirus-induced ovine pulmonary<br />

adenocarcinoma, an animal model for lung cancer. J. Natl. Cancer Inst.<br />

93, 1603–1614<br />

61 Theuring, F. et al. (1997) Transgenic animals as models of neurodegenerative<br />

disease in humans. Trends Biotechnol. 15, 320–325<br />

62 Petters, R.M. et al. (1997) Genetically engineered large animal model<br />

for studying cone photoreceptor survival and degeneration in retinitis<br />

pigmentosa. Nat. Biotechnol. 15, 965–970<br />

63 Mahmoud, T.H. (2003) Lensectomy and vitrectomy decrease the rate of<br />

photoreceptor loss in rhodopsin P347L transgenic pigs. Graefes Arch.<br />

Clin. Exp. Ophtalmol. 241, 298–308<br />

64 Weissmann, C. et al. (2002) Transmission of prions. Proc. Natl. Acad.<br />

Sci. U. S. A. 99 (Suppl. 4), 16378–16383<br />

65 Denning, C. et al. (2001) Deletion of the alpha (1,3)galactosyl<br />

transferase (GGTA1) and the prion protein (PrP) gene in sheep. Nat.<br />

Biotechnol. 19, 559–562<br />

66 Draghia-Akli, R. et al. (1999) Myogenic expression of an injectable<br />

protease-resistant growth hormone-releasing hormone augments<br />

long-term growth in pigs. Nat. Biotechnol. 17, 1179–1183<br />

67 Djojosubroto, M.W. et al. (2003) Telomeres and telomerase in aging,<br />

regeneration and cancer. Mol. Cells 15, 164–175<br />

68 Rudolph, K.L. et al. (1999) Longevity, stress response, and cancer in<br />

aging telomerase-deficient mice. Cell 96, 701–712<br />

69 Rudolph, K.L. et al. (2001) Telomere dysfunction and evolution of<br />

intestinal carcinoma in mice and humans. Nat. Genet. 28, 155–159<br />

70 Cawthon, R.M. et al. (2003) Association between telomere length in<br />

blood and mortality in people aged 60 years or older. Lancet 361,<br />

393–395<br />

71 Shiels, P.G. et al. (1999) Analysis of telomere lengths in cloned sheep.<br />

Nature 399, 316–317<br />

72 Wakayama, T. et al. (2000) Cloning of mice to six generations. Nature<br />

407, 318–319<br />

73 Tian, X.C. et al. (2000) Normal telomere lengths found in cloned cattle.<br />

Nat. Genet. 26, 272–273<br />

74 Lanza, R.P. et al. (2000) Extension of cell life-span and telomere length<br />

in animals cloned from senescent somatic cells. Science 288, 665–669<br />

75 Betts, D. et al. (2001) Reprogramming of telomerase activity and<br />

rebuilding of telomere length in cloned cattle. Proc. Natl. Acad. Sci.<br />

U. S. A. 98, 1077–1082<br />

76 Schätzlein, S. et al. (2004) Telomere length is reset during early<br />

mammalian embryogenesis. Proc. Natl. Acad. Sci. U. S. A. (accepted)<br />

www.sciencedirect.com<br />

77 Man<strong>zur</strong>, B.J. (2001) Developing transgenic grains with improved oils,<br />

proteins and carbohydrates. Novartis Found. Symp. 236, 233–239<br />

78 German, B. et al. (1999) The development of functional foods: lessions<br />

from the gut. Trends Biotechnol. 17, 492–499<br />

79 Grobet, L. et al. (1997) A deletion in the bovine myostatin gene causes<br />

the double-muscled phenotype in cattle. Nat. Genet. 17, 71–74<br />

80 Yom, H.C. and Bremel, R.D. (1993) Genetic engineering of milk<br />

composition: modification of milk components in lactating transgenic<br />

animals. Am. J. Clin. Nutr. 58 (Suppl.), 299S–306S<br />

81 Karatzas, C.N. and Turner, J.D. (1997) Toward altering milk<br />

composition by genetic manipulation: current status and challenges.<br />

J. Dairy Sci. 80, 2225–2232<br />

82 Jost, B. et al. (1999) Production of low-lactose milk by extopic<br />

expression of intestinal lactase in the mouse mammary gland. Nat.<br />

Biotechnol. 17, 160–164<br />

83 Stinnakre, M.G. et al. (1994) Creation and phenotypic analysis of<br />

a-lactalbumin-deficient mice. Proc. Natl. Acad. Sci. U. S. A. 91,<br />

6544–6548<br />

84 Kumar, S. et al. (1994) Milk composition and lactation of b-caseindeficient<br />

mice. Proc. Natl. Acad. Sci. U. S. A. 91, 6138–6142<br />

85 Brophy, B. et al. (2003) Cloned transgenic cattle produce milk with<br />

higher levels of b-casein and k-casein. Nat. Biotechnol. 21, 157–162<br />

86 Golovan, S.P. et al. (2001) Pigs expressing salivary phytase produce<br />

low-phosphorus manure. Nat. Biotechnol. 19, 741–745<br />

87 O’Brien, S.J. et al. (1999) The promise of comparative genomics in<br />

mammals. Science 286, 458–481<br />

88 Hughes, T.R. et al. (2000) Widespread aneuploidy revealed by DNA<br />

microarray expression profiling. Nat. Genet. 25, 333–337<br />

89 Templin, M.F. et al. (2002) Protein microarray technology. Trends<br />

Biotechnol. 20, 160–166<br />

90 Brambrink, T. et al. (2002) Application of cDNA arrays to monitor<br />

mRNA profiles in single preimplantation mouse embryos. Biotechniques<br />

33, 376–378<br />

91 Willadsen, S.M. (1986) Nuclear transplantation in sheep embryos.<br />

Nature 320, 63–65<br />

92 Wilmut, I. et al. (1997) Viable offspring derived from fetal and adult<br />

mammalian cells. Nature 385, 810–813<br />

93 Schnieke, A.E. et al. (1997) Human factor IX transgenic sheep<br />

produced by transfer of nuclei from transfected fetal fibroblasts.<br />

Science 278, 2130–2133<br />

94 Cibelli, J.B. et al. (1998) Cloned transgenic calves produced from<br />

nonquiescent fetal fibroblasts. Science 280, 1256–1258<br />

95 Chan, A.W.S. et al. (1998) Transgenic cattle produced by reversetranscribed<br />

gene transfer in oocytes. Proc. Natl. Acad. Sci. U. S. A. 95,<br />

14028–14033<br />

96 McCreath, K.J. et al. (2000) Production of gene-targeted sheep by<br />

nuclear transfer from cultured somatic cells. Nature 405,<br />

1066–1069<br />

97 Hofmann, A. et al. (2003) Efficient transgenesis in farm animals by<br />

lentiviral vectors. EMBO Rep. 4, 1054–1060


BIOLOGY OF REPRODUCTION 72, 1020–1028 (2005)<br />

Published online before print 22 December 2004.<br />

DOI 10.1095/biolreprod.104.031229<br />

Isolation of Murine and Porcine Fetal Stem Cells from Somatic Tissue 1<br />

Wilfried A. Kues, Björn Petersen, Wiebke Mysegades, Joseph W. Carnwath, and Heiner Niemann 2<br />

Department of Biotechnology, Institut für Tierzucht, Mariensee, D-31535 Neustadt, Germany<br />

ABSTRACT<br />

Adult stem cells have been previously isolated from a variety<br />

of somatic tissues, including bone marrow and the central nervous<br />

system; however, contribution of these cells to the germ<br />

line has not been shown. Here we demonstrate that fetal somatic<br />

explants contain a subpopulation of somatic stem cells<br />

(FSSCs), which can be induced to display features of lineageuncommitted<br />

stem cells. After injection into blastocysts, these<br />

cells give rise to a variety of cell types in the resultant chimeric<br />

fetuses, including those of the mesodermal lineage; they even<br />

migrate into the genital ridge. In vitro, FSSCs exhibit characteristics<br />

of embryonic stem cells, including extended self-renewal;<br />

expression of stem cell marker genes, such as Pou5f1 (Oct4),<br />

Stat3, and Akp2 (Tnap) and growth as multicellular aggregates.<br />

We report that fetal tissue contains somatic stem cells with<br />

greater potency than previously thought, which might form a<br />

new source of stem cells useful in somatic nuclear transfer and<br />

cell therapy.<br />

assisted reproductive technology, early development, embryo,<br />

environment, fibroblasts, germ line, Oct4, pluripotency, stem<br />

cells<br />

INTRODUCTION<br />

Adult stem cells with a surprisingly high degree of plasticity<br />

have been identified among hematopoietic and mesenchymal<br />

cell populations from bone marrow and the ependymal<br />

and subventricular zones of the central nervous<br />

system, as well as in vitro-derived cultures thereof [1–8].<br />

These cells are capable of contributing to organs of the<br />

three germ layers, but contribution to the germ line has not<br />

yet been shown. Recently, it has become a matter of debate<br />

whether these cells have inherent pluripotency or whether<br />

fusion with differentiated cells can account for the observed<br />

plasticity [9–12].<br />

Fibroblasts are the principal cell type in connective (mesodermal)<br />

tissue, where they mediate structural and functional<br />

processes, such as formation of a collagen meshwork<br />

and wound healing [13–15]. Primary fibroblasts can be obtained<br />

from subdermal and other connective tissues [16,<br />

17]. These cells represent a poorly characterized mixture of<br />

cell lineages [18], in which the composition varies with the<br />

developmental stage of the donor and the tissue of origin<br />

[14]. The regenerative ability of connective tissue suggests<br />

the presence of tissue-specific progenitor cells.<br />

In the present study, we have tested whether primary<br />

fibroblast cultures contain stem cells with the potential for<br />

multilineage differentiation. To identify rare stem cells<br />

1 Supported by DFG.<br />

2 Correspondence: FAX: 0 5034 871 101; e-mail: niemann@tzv.fal.de<br />

Received: 26 April 2004.<br />

First decision: 11 May 2004.<br />

Accepted: 9 December 2004.<br />

2005 by the Society for the Study of Reproduction, Inc.<br />

ISSN: 0006-3363. http://www.biolreprod.org<br />

1020<br />

within a large population of primary somatic cells, we used<br />

the OG2 transgenic mouse line, which carries the Oct4<br />

(also known as Pou5f1) promoter driving a green fluorescent<br />

protein (GFP) marker [19, 20]. Oct4 is a transcription<br />

factor of the POU family that is crucial for maintenance of<br />

embryonic stem cells in the undifferentiated state. It plays<br />

a major role in mouse embryogenesis [21–23]. Oct4 regulates<br />

expression of several genes, including Fgf4, Rex1<br />

(also known as Zpf42), Sox2, Ssp1 (osteopontin), Hand1<br />

and Ifnt1 [23, 24]. The Oct4 promoter is active in oocytes,<br />

zygotes, early cleavage-stage embryos, the inner cell mass<br />

of the blastocyst and in embryonic carcinoma and stem<br />

cells [21, 24, 25]. Oct4 is downregulated when cells leave<br />

the germ line and is found exclusively in germ cells in fetal<br />

and adult mice [21–23]. (Re)activation of the Oct4 gene in<br />

somatic cells would be a strong indication that these cells<br />

fulfill an essential molecular precondition for pluripotency.<br />

To test the potential for multilineage differentiation and<br />

functional contribution to organogenesis, the most informative<br />

assay is injection of the putative stem cells into<br />

recipient blastocysts and the generation of chimeras.<br />

Here, we report the identification of a fetal somatic stem<br />

cell (FSSC) population in primary fibroblast cultures derived<br />

from two mammalian species (mouse and pig) and<br />

show that simple changes of the culture conditions allow<br />

selective enrichment of the FSSCs. The FSSCs could be<br />

more amenable to reprogramming and would thus be a new<br />

source of donor cells in somatic nuclear transfer or celltherapy<br />

approaches.<br />

MATERIALS AND METHODS<br />

Cell Culture of Fetal and Adult Somatic Tissues<br />

Primary cultures were prepared by somatic explant (1 mm 3 ) cultures<br />

of connective tissue. For isolation of fetal cells, porcine fetuses of Postcoitus<br />

(PC) Day 25 were eviscerated, the extremities and the heads were<br />

removed. Cell cultures from adult pigs were isolated from subdermal tissues<br />

of ear clips. The tissues were isolated under sterile conditions and<br />

washed twice in phosphate buffered saline (PBS) containing antibiotics<br />

and were then cut into small pieces and pasted to cell culture dishes by<br />

employing recalcified microdrops of bovine plasma. Bovine plasma was<br />

isolated from EDTA-blood of healthy cows by centrifugation (4000 g,<br />

15 min) and stored frozen in aliquots at 80C. For use, an aliquot was<br />

thawed, supplemented with 80 mM CaCl 2 and microdrops of 5–10 l were<br />

pipetted into a cell culture dish. Small pieces of tissue were placed into<br />

the plasma drops, and culture dishes were incubated for 10–30 min at<br />

37C until plasma coagulation was completed. Subsequently, tissues were<br />

maintained in Dulbecco modified Eagles medium (DMEM) medium supplemented<br />

with 2 mM glutamine, 1% nonessential amino acids, 1% vitamin<br />

solution, 0.1 mM mercaptoethanol, 100 U/ml penicillin, 100 mg/ml<br />

streptomycin (Sigma, Deisenhofen, Germany) containing 10% fetal calf<br />

serum (FCS; batch numbers 40G321K, 40G2810K; Gibco, Karlsruhe, Germany)<br />

and incubated in humidified 95% air with 5% CO 2 at 37C [16,<br />

17]. Outgrowing cells were trypsinized and subpassaged once before cryopreservation.<br />

Typically, first outgrowing cells from fetal explants became<br />

visible within 24 h of incubation, and subpassaging was done after 7 days<br />

of culture. For high serum culture, the serum content of the standard medium<br />

was increased to 30% FCS. For suspension culture, colonies were<br />

selectively isolated and completely dissociated in a trypsin solution, then


10 4 cells were seeded into 35-mm bacteriological dishes. Every second<br />

day, 50% of the medium was replaced with new medium. To determine<br />

the maximal replicative limit, cultures were serially subpassaged and 12.5<br />

10 3 cells were seeded per square centimeter in six-well dishes, trypsinized<br />

after 5–7 days, counted, and reseeded. The number of accumulated<br />

population doublings per passage was determined using the equation, PD<br />

log(A/B)/log 2, in which A is the number of collected cells and B is the<br />

number of plated cells. Murine fibroblasts were obtained from Day 11.5–<br />

15.5 fetuses or adult OG2 mice [14] (homozygous for the Oct4-GFP transgene)<br />

or from double transgenic fetuses of crosses of OG2 with Rosa26<br />

mice. Confocal microscopy was applied to detect GFP using a Zeiss Axiomat<br />

LSM and an excitation wavelength of 488 nm. Embryonic stem<br />

(ES) cells (wildtype GS1 129/Sv) were cultured as described [26]. Animal<br />

handling was in accordance with institutional guidelines.<br />

In some experiments, porcine FSSCs were cultured in standard medium<br />

(DMEM 10% FCS) supplemented with growth factors. As controls, high<br />

serum and standard cultures were run in parallel and, after 7 days, the<br />

three-dimensional-colony formation and alkaline phosphatase (AP) induction<br />

were determined. The following growth factors were supplemented:<br />

human leukemia inhibitory factor (LIF, 1000 U/ml; Chemicon International),<br />

epidermal growth factor (EGF, 0.1 ng/ml; TEBU GmbH, Offenbach,<br />

Germany), platelet-derived growth factor (PDGF, 1.0 ng/ml; TEBU<br />

GmbH), basic fibroblast growth factor (bFGF, 1.0 ng/ml; TEBU GmbH),<br />

endothelial cell growth factor (ECGF, 500 ng/ml; Roche Diagnostics,<br />

Mannheim, Germany), and insulin (250 ng/ml; Sigma).<br />

In some experiments, porcine FSSCs were cultured in high serum medium<br />

supplemented with one of the following inhibitory peptides: 1) bFGF<br />

inhibitory peptide H-1948 (5 M; Bachem, Weil am Rhein, Germany), 2)<br />

insulin-like growth factor (IGF-I) analog H-1356 (10 M; Bachem), 3)<br />

PDGF antagonist H-8650 (10 M; Bachem), and 4) insulin receptor (689-<br />

702) H-4212 (10 M; Bachem). In parallel, high serum and standard cultures<br />

were run as controls and, after 7 days, three-dimensional-colony<br />

formation and AP induction were determined.<br />

DNA contents were measured by fluorescent activated cell sorting as<br />

described [16]. In addition, nuclei from FSSCs were karyotyped and the<br />

ploidy status was assessed.<br />

RT-PCR of Specific mRNAs<br />

In brief, total RNA was isolated from cells grown in six-well dishes<br />

and 0.5 g RNA was reverse transcribed (RT) into cDNA using random<br />

hexamers in a total volume of 40 l. Aliqouts of 2.5 l of the RT-reaction<br />

were used as templates for polymerase chain reaction (PCR). Murine Oct4<br />

and Stat3 were amplified by PCR with the following primers and conditions:<br />

5-GGC GTT CTC TTT GGA AAG GTG TTC and 5-CTC GAA<br />

CCA CAT CC TTC TCT (35 cycles, annealing temperature 57C) for the<br />

murine Oct4; 5-TCA AGC ACC TGA CCC TTA GG and 5-CTG AAG<br />

CGC AGT AGG AAG GT (37 cycles, 58C annealing temperature) for<br />

Stat3 (U06922). For the amplification of TNAP (AF285233), the following<br />

primer pair was used: 5-ATG AGG GTA AGG CCA AGC AG and 5-<br />

CCA CCA TGG AGA CAT TTT CC (34 cycles, 58C annealing temperature).<br />

Porcine OCT4 was amplified with intron-spanning 5-AGG TGT<br />

TCA GCC AAA CGA CC and 5-TGA TCG TTT GCC CTT CTG GC<br />

primers (AJ251914) and 36 cycles. Glyceraldehyde phosphate dehydrogenase<br />

(GAPD) was amplified with the primer pair 5-CCT CCA CTA<br />

CAT GGT CTA CAT GTT CCA GTA and 5-GCC TGC TTC ACC ACC<br />

TTC TTG ATG TCA TCA (25 cycles, annealing temperature 60C). The<br />

GAPD primers matched completely with the porcine ortholog (U48832)<br />

and showed a single nucleotide mismatch with the mouse ortholog<br />

(BC083065). The PCR reactions were performed in 20-l volumes, consisting<br />

of 20 mM TrisHCl (pH 8.4), 50 mM KCl, 1.5 mM MgCl 2, 200<br />

M deoxynucleoside triphosphates, 1 M of specific primer pairs and 0.5<br />

U of Taq DNA polymerase (Gibco). Reactions without RT or without<br />

template were used as controls.<br />

Measurement of Cell Proliferation by BrdU Incorporation<br />

DNA synthesis was measured by 5-bromo-2deoxy-uridine (BrdU) incorporation<br />

as described [17]. Incorporated BrdU was detected by a chromogenic<br />

immunoassay employing an anti-BrdU antibody conjugated with<br />

alkaline phosphatase.<br />

Immunohistology<br />

For antibody staining, cells were cultured on gelatinized coverslips.<br />

Sterile coverslips were placed in six-well cell culture dishes and covered<br />

with a 1% gelatine solution (in PBS) for 5 min. The gelatine solution was<br />

FETAL SOMATIC STEM CELLS<br />

1021<br />

then removed and the semidry coverslips were seeded with cells. After 3–<br />

4 days of culture, the cells were fixed in cold 80% methanol. The following<br />

monoclonal antibody dilutions were used: anti-vimentin (AMF-17b, 1:<br />

200; Developmental Studies Hybridoma Bank, IA) and anticytokeratin<br />

(peptide17, 1:100; Sigma). A rhodamine-labeled secondary anti-mouse antibody<br />

(1:2000; Molecular Probes, Netherlands) was used. In some cases,<br />

the nuclei were counterstained with 1 mM Hoechst 33342 [27]. The samples<br />

were examined with an Olympus BX60 microscope equipped with<br />

phase-contrast and epifluorescence optics, using band-pass rhodamine and<br />

Hoechst filter sets.<br />

Staining of Endogenous Alkaline Phosphatase Activity<br />

Standard and high-density cell cultures were prepared in six-well cell<br />

cultures dishes. The cultures were washed with PBS, fixed in 3.7% paraformaldehyde<br />

for 15 min, washed in PBS, and then incubated in a solution<br />

containing 25 mM TrisHC (pH 9.0), 4 mM MgCl 2, 0.4 mg sodium-naphtylphosphate,<br />

1 mg/ml Fast Red TR (Sigma), and 0.05% Triton X-<br />

100 for 60 min.<br />

Chimera Generation by FSSC Injection<br />

into Host Blastocysts<br />

Rosa26 homozygous mice were obtained from Jackson Laboratory<br />

(Bar Harbor, ME) and mated with homozygous OG2 animals to generate<br />

double-transgenic fetuses carrying both marker genes, which were used to<br />

isolate FSSCs. Day 11.5—15.5 fetuses were isolated and employed for<br />

fetal cell cultures using the explant method described above.<br />

For blastocyst injections, 6- to 10-wk-old female CD2F1 mice were<br />

superovulated with 10 U eCG at noon on Day 2, followed by 10 U hCG<br />

on Day 0, and were then mated with CD2F1 males. The next day, females<br />

were checked for plug formation. At Day 3.5, females were sacrificed and<br />

the uterine tracts were isolated and flushed with PBS containing 1% bovine<br />

serum albumin. Blastocysts were isolated and incubated in PBS/1% albumin<br />

at 37C. Single blastocysts were transferred into a micromanipulation<br />

unit (Zeiss) and fixed with a holding pipette. On average, 2–15<br />

double transgenic cells (OG2/Rosa26) were injected into the blastocoel<br />

with the aid of a microcapillary. In total, 8–10 blastocysts were transferred<br />

into the uteri of Day 2.5 or Day 3.5 pseudopregnant NMRI females that<br />

had been mated with vasectomized males. Fetuses were recovered at Day<br />

10.5–15.5 and either stained for lacZ-positive cells [28] as whole mounts,<br />

or dissected and screened for GFP expression in genital ridges and other<br />

organs.<br />

RESULTS<br />

Activation of the Germline-Specific Oct4 Promoter<br />

in Somatic Cells<br />

Fetal OG2-transgenic mice, which express the GFP gene<br />

under transcriptional control of the germ-line-specific Oct4<br />

promoter, were employed for explant cultures. Explants of<br />

somatic tissue with an average size of 1 mm3were isolated<br />

from fetuses at PC Days 11.5, 13.5, and 14.5; attached<br />

with microdrops of bovine plasma to cell culture dishes,<br />

and cultured in DMEM. Special care was taken to isolate<br />

explants from mesodermal tissue of the neck and shoulder<br />

regions. As expected, no GFP-positive cells were found in<br />

the tissue explants and in the first outgrowths after 2 days<br />

in culture (Fig. 1, C and D). However, after 8 days of culture,<br />

GFP-positive cells were detected within the outgrowing<br />

primary cells (Fig. 1, E and F), indicating (re)activation<br />

of the germ-line-specific Oct4-GFP marker cassette. As<br />

positive control, the genital ridges were isolated in parallel<br />

and primordial germ cells could be unequivocally detected<br />

by their strong expression of the Oct4-GFP marker (Fig. 1,<br />

A and B). Additional experiments suggested that the microenvironment<br />

of the plasma microdrops played an important<br />

role in stimulating the proliferation of GFP-positive<br />

cells (Fig. 1, G–I). Therefore, subpassages of the outgrowing<br />

cells were cultured in DMEM containing high fetal calf<br />

serum supplementation (30% FCS) and a population of<br />

103 GFP-positive cells (Fig. 1, G–I) could be main-


1022 KUES ET AL.<br />

FIG. 1. Activation of Oct4-promoter in<br />

somatic explants of Oct4-eGFP tg mice.<br />

Genital ridge of a male fetus (PC Day<br />

14.5) with tissue-specific expression of<br />

GFP in the primordial germ cells is shown<br />

under fluorescent (A) and bright-field optics<br />

(B). Bar 150 m. The primordial<br />

germ cells show expression of GFP for 8<br />

days in culture (A, B), no outgrowing GFPpositive<br />

cells could be detected. Outgrowing<br />

cells of a somatic explant, under fluorescent<br />

(C) and bright-field (D) optics after<br />

2 days of culture. No GFP-positive cells<br />

were found. In the upper left, the explant<br />

is visible. After 8 days in culture, several<br />

GFP-positive cells were detectable within<br />

the outgrowing cells (E, F). Bar 140 m.<br />

Confocal analysis of murine FSSCs cultured<br />

in medium with high serum supplementation:<br />

(G) fluorescent, (H) bright-field,<br />

and (I) merged images; bar 10 m. The<br />

GFP is preferentially located in the cytoplasm,<br />

probably because it doesn’t contain<br />

a nuclear localization motif. J) RT-PCR detection<br />

of endogenous Oct4, Stat3, and<br />

Tnap transcripts in murine FSSCs; M, DNA<br />

ladder; lane 1: FSSCs; lane 2: no RT; lane<br />

3: ES cells; lane 4: no RT; lane 5: no template<br />

control.<br />

tained. Expression of the endogenous Oct4 gene was confirmed<br />

by RT-PCR (Fig. 1J). Oct4 is one of the best characterized<br />

genes for stemness. The presence of Oct4 transcripts<br />

has been shown by two independent assays in this<br />

study. The absence of a PCR product from fibroblast cultures<br />

argues against the amplification of an Oct4 pseudogene.<br />

In addition, the tissue nonspecific variant of alkaline<br />

phosphatase (TNAP) and Stat3 were expressed in high serum-supplemented<br />

cultures as determined by RT-PCR (Fig.<br />

1); in contrast, no expression of the embryonic or intestinal<br />

AP genes could be detected (not shown).<br />

In Vivo Differentiation Potential by Injection of FSSCs<br />

into Blastocysts<br />

To determine the developmental potential, FSSCs were<br />

injected into murine blastocysts, which were subsequently<br />

transferred to pseudopregnant recipients. FSSCs of both<br />

sexes were isolated from double transgenic fetuses of OG2<br />

and Rosa26 mouse strains. These cells carried the germlinespecific<br />

Oct4 GFP and the ubiquitously active lacZ reporter<br />

gene constructs and thus allowed distinguishing them from<br />

the cells of the recipient blastocysts.<br />

Day 10.5–15.5 fetuses derived from the injected blas-<br />

tocysts were isolated and analyzed for chimerism either by<br />

staining for lacZ activity or by fluorescence microscopy to<br />

identify GFP-positive cells. Of a total of 19 analyzed fetuses,<br />

7 contained progeny cells from the injected FSSCs<br />

(Table 1). Chimerism was detected in mesenchymal organs,<br />

such as liver, muscle, and tongue, but also in the genital<br />

ridges. Figure 2A shows an example of a chimeric fetus<br />

with massive lacZ staining in liver, tongue, and genital ridges,<br />

suggesting that at least parts of these organs were derived<br />

from the injected cells. Chimeric and wildtype fetuses<br />

were derived from embryo transfers that had been performed<br />

on the same day, were stained for LacZ activity in<br />

parallel, and photographed on the same slide. It is unclear<br />

whether the apparent oversize of the chimeric fetus is related<br />

to the cell injection. The summarized data for the<br />

blastocyst transfer suggest that development of embryos after<br />

FSSC injection is compromised (Table 1). Figure 2B<br />

shows the presence of GFP-positive cells in the genital<br />

ridges of a male PC Day 15.5 fetus. In total, 16 GFP-positive<br />

cells were counted in the squeeze preparation, and<br />

these cells behaved like prospermatogonia in that they floated<br />

within the ducts of the genital ridges. GFP-positive cells<br />

were not found in other organs, such as heart, liver, brain,<br />

or connective tissue.


FETAL SOMATIC STEM CELLS<br />

TABLE 1. Generation of chimeric fetuses by injection of FSSCs into recipient blastocysts.*<br />

No. injected FSSCs<br />

6–8<br />

10–15<br />

2–5<br />

6–8<br />

10–15<br />

Control blastocysts<br />

w/o cell injection<br />

* NA, not applicable.<br />

Transgenic<br />

background<br />

Rosa26<br />

Rosa26<br />

Rosa26/OG2<br />

Rosa26/OG2<br />

Rosa26/OG2<br />

Isolation of FSSCs from Porcine Tissues<br />

No. of<br />

blastocysts<br />

transferred<br />

57<br />

6<br />

24<br />

49<br />

20<br />

Recovered<br />

fetuses<br />

4<br />

1<br />

5<br />

9<br />

0<br />

Chimeric<br />

fetuses Assay Positive cells found in<br />

0of4<br />

1of1<br />

3of5<br />

2of7<br />

1of2<br />

wt 29 14 0 of 9<br />

0of5<br />

The isolation of FSSCs with stem-like properties from<br />

murine somatic explants raised the question whether this is<br />

a species-specific phenomenon or whether similar cells<br />

could be obtained from other species. Therefore, somatic<br />

explants of porcine fetuses (PC Day 25) were established<br />

and subpassaged once using standard protocols. Immunostaining<br />

showed uniform labeling for vimentin and no labeling<br />

for cytokeratin (data not shown), suggesting that<br />

most of the cells were fibroblasts. However, culture in<br />

DMEM with high serum supplementation activated the porcine<br />

OCT4 gene (Fig. 3M). Porcine OCT4 was amplified<br />

with intron spanning porcine-specific primers; the obtained<br />

—<br />

LacZ:<br />

LacZ:<br />

LacZ:<br />

LacZ:<br />

OG2:<br />

LacZ:<br />

OG2:<br />

none<br />

liver, genital ridge, tongue<br />

mesoderm, sev. organs, low chimerism<br />

mesoderm, sev. organs, low chimerism<br />

genital ridge<br />

NA<br />

some background in spinal cord<br />

—<br />

1023<br />

PCR product was purified, sequenced, and found to be<br />

identical to the corresponding OCT4 cDNA sequence. Control<br />

cultures maintained in DMEM/10% FCS did not express<br />

detectable OCT4 mRNA levels<br />

With high serum concentrations, the cultures grew faster<br />

and lost contact inhibition, resulting in the formation of<br />

three-dimensional colonies (Fig. 3, A and B). Only cells<br />

within the three-dimensional colonies continued to proliferate<br />

as measured by BrdU incorporation, whereas cells in<br />

the surrounding monolayers were mitotically inactive (Fig.<br />

3D). Staining for endogenous AP activity revealed a massive<br />

induction of AP-positive cells, almost exclusively associated<br />

with the three-dimensional colonies (Fig. 3, G–J).<br />

FIG. 2. Generation of chimeric fetuses by<br />

FSSC injection into blastocysts. A) Wholemount<br />

staining for LacZ activity in a control<br />

fetus (left) and a fetus (PC Day 15.5)<br />

derived from a FSSC (Rosa26/OG2)-injected<br />

blastocyst (right). Note the -galactosidase<br />

staining in liver (arrow) and genital<br />

ridge (arrowheads) of the chimeric fetus.<br />

B) Oct4 promoter driven expression of<br />

GFP in the genital ridges of a chimeric fetus<br />

(PC Day 15.5) derived from a FSSC<br />

(Rosa26/OG2)-injected blastocyst (left).<br />

Genital ridge from a control OG2/Rosa26<br />

fetus (right). Bar 20 m. C) Higher magnification<br />

of a GFP-positive cell (top) in<br />

the genital ridge of a chimeric fetus (PC<br />

Day 15.5) and corresponding phase-contrast<br />

image (bottom). Bar 20 m.


1024 KUES ET AL.<br />

FIG. 3. Induction of three-dimensional<br />

growth and AP-positive cells in porcine<br />

FSSCs. A) Three-dimensional-colony<br />

growth of porcine FSSCs (passage 3, 5<br />

days) in culture medium with high serum<br />

supplementation, (B) same view as in A<br />

with Hoechst 33324-stained nuclei under<br />

ultraviolet excitation. Bar 20 m for A–<br />

F; it is displayed in D. C) Control culture<br />

of the same cell batch cultured in standard<br />

medium (10% FCS, 5 days). D) BrdU incorporation<br />

in high serum cultures (5 days,<br />

30% FCS). Note that only cells within the<br />

three-dimensional colonies (arrows) incorporated<br />

BrdU, the surrounding monolayer<br />

is unlabelled; inset: another three-dimensional<br />

colony. E) BrdU incorporation in<br />

proliferating fibroblasts (three-dimensional,<br />

standard medium with 10% FCS); the majority<br />

of the cells are labeled. F) BrdU incorporation<br />

in confluent fibroblasts (5<br />

days, standard medium), the majority of<br />

the cells became contact inhibited and<br />

stopped proliferating. G–J) Induction of<br />

AP-positive cells, accompanied with thredimensional-colony<br />

growth after 2, 4, 6, 8<br />

days in high serum culture. K) Higher<br />

magnification of AP-positive cells aggregated<br />

in three-dimensional colony (4<br />

days). L) Individual AP-positive cells within<br />

the fibroblast monolayer. Bars 20 m.<br />

M) Species-specific RT-PCR detection of<br />

OCT4 transcripts in porcine FSSCs (top); 1,<br />

DNA ladder with prominent 600-base pair<br />

fragment; lane 2: no template; lane 3: murine<br />

ES cells; lane 4: porcine fibroblasts;<br />

lane 5: porcine FSSCs. As control, transcripts<br />

of the housekeeping gene GAPD<br />

were amplified by RT-PCR (bottom).<br />

AP-positive cells differed (Fig. 3, K and L) from typical<br />

fibroblasts in that they displayed dendritic projections.<br />

These changes were not found in control cultures from<br />

identical batches of cells when grown under standard conditions<br />

with 10% FCS supplementation (Fig. 3C, Table 2).<br />

In total, 6.7% of microwells seeded with 10 cells from high<br />

serum cultures resulted in continuously growing cultures,<br />

suggesting that 1 out of 150 cells was able to initiate clonal<br />

growth. Induction of colony growth and AP expression by<br />

high serum supplementation were abolished by pretreatment<br />

with heat (90C) or trypsin (data not shown).<br />

Adult porcine fibroblasts derived from subdermal tissue<br />

explants of three animals did not display three-dimensionalcolony<br />

growth (Fig. 4) when cultured for 7 days in DMEM/<br />

30% FCS, but the frequency of AP-expressing cells increased<br />

2- to 10-fold over this period (Table 2). In comparison,<br />

when fetal cells were used, a 100- to 1000-fold<br />

increase in AP-positive cells was observed. Induction of<br />

three-dimensional-colony growth and AP expression in porcine<br />

cultures was at least one order of magnitude higher<br />

than that seen in murine cultures (Table 2).<br />

When high serum cultures were split and one part of the<br />

population was returned to standard medium, colony<br />

growth ceased and AP-positive cells disappeared in the


FETAL SOMATIC STEM CELLS<br />

TABLE 2. High serum induction of AP-positive cells and three-dimensional (3D)-colonies in porcine and murine cell isolates.*<br />

Age of donors Tissue source Method n<br />

porcine<br />

Day 25–27 PC<br />

Day 25 PC<br />

0.5–1.5 y<br />

murine<br />

Day 11.5 PC<br />

Day 13.5 PC<br />

Day 14.5 PC<br />

4mo<br />

Mesoderm<br />

Mesoderm<br />

Ear biopsy<br />

Mesoderm<br />

Mesoderm<br />

Mesoderm<br />

Subdermal tissue<br />

Try<br />

Expl<br />

Expl<br />

2<br />

12<br />

3<br />

High serum induced<br />

1025<br />

AP-positive cells (fold<br />

increase compared with<br />

standard cultures) 3D colonies (no./6-well)<br />

250–850<br />

100–1000<br />

2–10<br />

* Try, Trypsinisation of pooled fetuses (n 6–8); Expl, explant cultures from individual fetuses or adult subdermal tissues.<br />

Expl<br />

Expl<br />

Expl<br />

Expl<br />

standard culture within 2–3 subpassages, suggesting that<br />

the induction and proliferation of FSSCs are dependent on<br />

high levels of not-yet-identified serum factors (Fig. 4B).<br />

When high serum cultures were subpassaged five times<br />

in standard medium, the colony growth phenotype was lost<br />

and a switch to high serum medium was no longer associated<br />

with the appearance of three-dimensional growth,<br />

suggesting that the FSSC population had been lost due to<br />

unfavorable conditions.<br />

To narrow down the possibilities of potential factors in<br />

serum responsible for these effects, it was tested whether<br />

purified growth factors could substitute for high serum supplementation.<br />

Porcine FSSC cultures in standard medium<br />

(DMEM 10% FCS) were supplemented with either hLIF,<br />

PDGF, FGF1, ECGF, or insulin. However, none of these<br />

growth factors induced three-dimensional growth or induction<br />

of AP expression compared with approximately 200–<br />

300 colonies in the positive controls (not shown).<br />

In addition, no inhibitory effect of the inhibitory peptides<br />

FGF1 inhibitory peptide, IGF1 analog, PDGF antagonist,<br />

or insulin-receptor [609–702] [29–32] on the FSSC<br />

phenotype could be detected; the peptide supplemented cultures<br />

developed the same number of FSSC colonies as the<br />

positive controls. These data indicate that the examined<br />

growth factors and receptors are not involved in the apparent<br />

pluripotent phenotype induced by the high serum supplementation.<br />

FSSCs Have Increased Proliferative Potential<br />

Culture medium supplemented with high serum resulted<br />

in a dramatically altered growth curve (Fig. 4C). Porcine<br />

FSSC cultures maintained under high serum conditions<br />

grew continuously over a period of 120 days and exceeded<br />

more than 100 cell doublings without reaching a<br />

plateau phase (Fig. 4C). In contrast, standard cultures, i.e.,<br />

DMEM with 10% FCS, were compatible with only 50–60<br />

cell doublings before mitotic activity ceased after approximately<br />

70 days. The FSSCs maintained a diploid status, as<br />

measured by fluorescence-activated cell sorting (Fig. 4D)<br />

and metaphase spreads of cells from passage 20. Thirtytwo<br />

karyotyped nuclei of porcine FSSC cultures contained<br />

the euploid number of 38 chromosomes; only 1 nucleus<br />

with 37 chromosomes was found.<br />

FSSCs Are Capable of Forming Spheroids and Exhibit<br />

Anchorage-Independent Growth<br />

To investigate the growth potential of the colony-forming<br />

cells, porcine three-dimensional colonies of 200–300<br />

3<br />

1<br />

3<br />

1<br />

8<br />

3<br />

11<br />

1.5<br />

100–290<br />

65–340<br />

3–5<br />

0<br />

3–5<br />

0<br />

m diameter were isolated and trypsinized to obtain singlecell<br />

suspensions. Subsequently, 10 4 cells were seeded into<br />

bacteriological culture dishes. In DMEM/30% FCS, irregular<br />

aggregates consisting of only few cells (2–20), were<br />

detected. These cells did not divide and the majority apparently<br />

underwent cell death (Fig. 4E). Culture medium<br />

supplemented with retinoic acid induced the formation of<br />

tiny aggregates, which, after 2–4 days, attached to the surface<br />

(Fig. 4F). Supplementation of the culture medium<br />

(DMEM/30% FCS) with dexamethasone resulted in aggregation<br />

of small multicellular spheroids within 24 h, which<br />

continued to grow up to a diameter of 400 m after 10–<br />

15 days and contained nearly exclusively AP-positive cells<br />

(Fig. 4, G and H). Dexamethasone is an antiinflammatory<br />

glucocorticoid, which regulates T cell survival, growth, and<br />

differentiation and inhibits the induction of nitric oxide synthase.<br />

Several protocols for differentiation of stem cells also<br />

employed dexamethasone. The dexamethasone-derived<br />

spheroids were capable of forming secondary spheroids if<br />

trypsinized and reseeded in bacteriological dishes. If plated<br />

on gelatinized coverslips, dexamethasone spheroids reattached<br />

and monolayer cells grew out. These outgrowing<br />

cells were vimentin negative, whereas control cultures kept<br />

in standard medium with 10% FCS were strongly positive<br />

(Fig. 4, I and J).<br />

DISCUSSION<br />

The present findings provide compelling evidence for the<br />

presence of a somatic stem cell population in explant cultures<br />

derived from mouse and pig fetuses. The FSSC cells<br />

are characterized by extended proliferative capacity, altered<br />

morphology, expression of stem cell-specific markers, including<br />

Oct4, Stat3, and Tnap and by contact- and anchorage-independent<br />

growth. Chimeric fetuses, obtained by injection<br />

of murine FSSCs into recipient blastocysts, showed<br />

that the FSSCs were able to contribute to various mesenchymal<br />

organs and in particular the genital ridges. Whether<br />

FSSCs show a similar broad spectrum of differentiation potential<br />

as multipotent adult progenitor cells [6] remains to<br />

be shown. As some, al<strong>bei</strong>t few, cells expressed GFP fluorescence<br />

driven by the Oct4 promoter in the chimeric genital<br />

ridges, germ-line transmission might be possible. The<br />

final proof for germ-line transmission by FSSCs would be<br />

the generation of live and fertile chimeric animals. The<br />

finding that GFP-positive cells were not found outside of<br />

the genital ridges indicates that the Oct4 marker was correctly<br />

activated in cells committed to the germ line. It also<br />

suggests that at least some of the FSSC descendants were


1026 KUES ET AL.<br />

FIG. 4. Growth characteristics of FSSCs<br />

in vitro. A) Porcine cultures from fetal and<br />

adult origin of the same batches, respectively,<br />

were split and cultured with high<br />

serum (30%) or standard (10% FCS) conditions<br />

in six-well plates; after 5 days, the<br />

cultures were fixed and stained for endogenous<br />

AP activity. Note the massive induction<br />

of AP-positive three-dimensional colonies<br />

in the fetal culture (red dots). B) The<br />

induction of three-dimensional-colony<br />

growth and AP expression is serum dependent.<br />

After six passages with constant<br />

three-dimensional-colony formation and<br />

AP expression in high serum (30% FCS),<br />

fetal cells were trypsinized, replated, and<br />

cultured for two passages with standard<br />

medium (10% FCS) before AP staining. C)<br />

Growth curves of fetal somatic cells cultured<br />

in standard medium (□) containing<br />

10% FCS and high serum medium ()<br />

containing 30% FCS. D) Cell cycle status<br />

in standard and high serum culture. Note<br />

that the high serum culture displays a normal<br />

ploidy. E, F) Anchorage-independent<br />

growth of FSSCs in suspension culture.<br />

High serum-induced three-dimensional<br />

colonies were isolated, trypsinized to single<br />

cell suspensions, and seeded into bacteriological<br />

dishes to prevent attachment.<br />

E) In DMEM with high serum supplementation,<br />

tiny aggregates formed. F) In high<br />

serum medium with retinoic acid (10 7<br />

M), the initial aggregates reattach and<br />

show outgrowing cells on the surface. G)<br />

In high serum medium with dexamethasone<br />

(10 7 M), spheroids of 300-m size<br />

grow over 10–15 days; inset: lower magnification.<br />

H) Dexamethasone-spheroids<br />

stained for endogenous AP, bar 230<br />

m. I) Expression of vimentin in fibroblasts<br />

cultured in standard medium (passage<br />

5), merged image of antibody (red)<br />

and nuclei (blue) staining (left). J) Loss of<br />

vimentin epitopes in cells derived from<br />

dexamethasone spheroids. After 15 days of<br />

suspension culture, the spheroids were allowed<br />

to reattach to gelatinized coverslips<br />

and probed with a monoclonal antivimentin<br />

antibody.<br />

capable of migrating into the genital ridge. The relatively<br />

low percentage of chimerism might be due to the fact that<br />

the cells used for blastocyst injection were not preselected<br />

for Oct4-GFP expression.<br />

Chimerism was found preferentially in liver, muscle, and<br />

tongue. The LacZ-positive cells in the tongue may actually<br />

represent lymphatic tissue. It is well known that the fetal<br />

liver contains hepatic as well as migratory hematopoietic<br />

progenitor cells. However, we have not yet determined<br />

which cell types in these three organs are derived from the<br />

injected cells. This is the subject of future studies. No chimerism<br />

was detected in heart and brain, two organs that<br />

showed a high rate of spontaneous cell fusions in a recent<br />

study [12]. However, we cannot fully exclude the possibility<br />

that fusion with differentiated cells might have contributed<br />

to the observed chimerism.<br />

FSSCs represent a rare subpopulation in fetal fibroblast<br />

cultures, and the proliferation of FSSCs critically depends<br />

on the culture conditions, including an autologous feedercell<br />

layer and factors present in bovine serum. The explant<br />

culture is based on microdrops of bovine plasma as an adhesive<br />

support, which seems to be essential for the initial<br />

stimulation of FSSC proliferation. Standard culture media<br />

compositions, with no or low amounts of FCS supplementation,<br />

are associated with a progressive loss of the FSSC<br />

phenotype. These observations show that appearance of<br />

FSSCs is critically dependent on cell-collection method<br />

and/or culture conditions.<br />

At present, it is not clear whether FSSCs represent a<br />

preexisting subpopulation in the fetus or whether they are<br />

the result of a de novo formation induced by the culture<br />

conditions. Potential sources are ectopic primordial germ<br />

cells [19] or fetal tissue-specific stem cells. In humans, it<br />

has been shown that mesenchymal stem cells are present in<br />

a variety of tissues during development, whereas in adults,<br />

they are predominantly found in the bone marrow [33].


Activation of the Oct4 promoter in murine explant outgrowths,<br />

although at lower expression levels than in murine<br />

ES cells, after a few days in culture suggests that the culture<br />

conditions induce this phenotype. We cannot completely<br />

rule out that this reprogramming is attributed to dedifferentiation<br />

of a susceptible, non-lineage-committed cell population<br />

present in fetal tissues. The observed frequency of<br />

1/1000 GFP-positive cells in the in vitro cultures may<br />

actually underestimate the occurrence of FSSCs, as GFP<br />

expression was low and might be around the detection limit.<br />

Clonal growth was found for 1 out of 150 cells and<br />

chimerism was detected in 7 out of 19 fetuses.<br />

In line with the expression of stem cell marker genes,<br />

FSSCs show three-dimensional-colony growth in adherent<br />

culture and spheroid growth in suspension culture, whereas<br />

fibroblasts cease dividing and arrest in the G1 phase of the<br />

cell cycle under these conditions [34–36]. Loss of contactinhibition<br />

and anchorage-independent growth is characteristic<br />

for ES cells or transformed cells [37]. Our data provide<br />

convincing evidence that, unlike many cell lines derived<br />

from tumors or cells transformed by oncogenic agents, the<br />

FSSC subpopulation does not result from spontaneous immortalization<br />

or transformation. FSSCs do not exhibit a crisis<br />

followed by clonal outgrowth, and they show no chromosomal<br />

abnormalities or aneuplodies. The altered growth<br />

characteristics of FSSCs are reversed after exposure to standard<br />

cell culture conditions. Unlike rodent cell cultures,<br />

cells from humans and livestock rarely, if ever, immortalize<br />

spontaneously [38]. Our findings show that FSSCs can be<br />

efficiently and reliably isolated from porcine, but also from<br />

bovine and ovine, fetal explants (data not shown) and with<br />

a lower efficiency from murine fetal explants. The diminished<br />

efficiency of FSSC stimulation from murine explant<br />

cultures may be attributed to the species specificity of factors<br />

present in bovine serum.<br />

In conclusion, this study shows that explants from fetal<br />

somatic tissues contain a subpopulation of potentially pluripotent<br />

stem cells, which fulfill most of the essential criteria<br />

for stem cells, including self-renewal ability, multilineage<br />

differentiation, and differentiation in vivo [39]. Therefore,<br />

the FSSCs might hold great promise for regenerative<br />

therapies.<br />

ACKNOWLEDGMENTS<br />

The authors thank Erika Lemme for advice on blastocyst injection. We<br />

gratefully acknowledge critical discussion and the gift of the OG2-transgenic<br />

mouse line from Hans Schöler.<br />

REFERENCES<br />

1. Weissman IL. Stem cells: units of development, units of regeneration<br />

and units of evolution. Cell 2000; 100:157–168.<br />

2. Shih CC, Weng Y, Mamelak A, LeBon T, Hu MC, Forman SJ. Identification<br />

of a candidate human neurohematopoietic stem-cell population.<br />

Blood 2001; 98:2412–2422.<br />

3. Tropepe V, Coles BL, Chiasson BJ, Horsford DJ, Elia AJ, McInnes<br />

RR, van der Kooy D. Retinal stem cells in the adult mammalian eye.<br />

Science 2000; 287:2032–2036.<br />

4. Bjornson C, Rietze RL, Reynolds BA, Magli MC, Vescovi AL. Turning<br />

brain into blood: a hematopoietic fate adopted by adult neural<br />

stem cells in vivo. Science 1999; 283:354–357.<br />

5. Lagasse E, Connors H, Al-Dhalimy M, Reitsma M, Dohse M, Osborne<br />

L, Wang X, Finegold M, Weissman IL, Grompe M. Purified<br />

hematopoietic stem cells can differentiate into hepatocytes in vivo.<br />

Nat Med 2000; 6:1229–1234.<br />

6. Jiang Y, Jahagirdar BN, Reinhardt RL, Schwartz RE, Keene CD, Ortiz-Gonzalez<br />

XR, Reyes M, Lenvik T, Lund T, Blackstad M, Du J,<br />

Aldrich S, Lisberg A, Low WC, Largaespada DA, Verfaillie CM.<br />

FETAL SOMATIC STEM CELLS<br />

1027<br />

Pluripotency of mesenchymal stem cells derived from adult marrow.<br />

Nature 2002; 418:41–49.<br />

7. Clarke DL, Johansson CB, Wilbertz J, Veress B, Nilsson E, Karlstrom<br />

H, Lendahl U, Frisen J. Generalized potential of adult neural stem<br />

cells. Science 2000; 288:1660–1663.<br />

8. Raff M. Adult stem cell plasticity: fact or artifact? Annu Rev Cell<br />

Dev Biol 2003; 19:1–22.<br />

9. Terada N, Hamazaki T, Oka M, Hoki M, Mastalerz DM, Nakano Y,<br />

Meyer EM, Morel L, Petersen BE, Scott EW. Bone marrow cells adopt<br />

the phenotype of other cells by spontaneous cell fusion. Nature 2002;<br />

416:542–545.<br />

10. Ying QL, Nichols J, Evans EP, Smith AG. Changing potency by spontaneous<br />

fusion. Nature 2002; 416:545–548.<br />

11. Wang X, Willenbring H, Akkari Y, Torimaru Y, Foster M, Al-Dhalimy<br />

M, Lagasse E, Finegold M, Olson S, Grompe M. Cell fusion is the<br />

principal source of bone-marrow-derived hepatocytes. Nature 2003;<br />

422:897–901.<br />

12. Alvarez-Dolado M, Pardal R, Garcia-Verdugo JM, Fike JR, Lee HO,<br />

Pfeffer K, Lois C, Morrison SJ, Alvarez-Buylla A. Fusion of bonemarrow-derived<br />

cells with Purkinje neurons, cardiomyocytes and hepatocytes.<br />

Nature 2003; 425:968–973.<br />

13. Moulin V, Garrel D, Auger FA, O’Connor-McCourt M, Castilloux G,<br />

Germain L. What’s new in human wound-healing myofibroblasts?<br />

Curr Top Pathol 1999; 93:123–133.<br />

14. Chang HY, Chi JT, Dudoit S, Bondre C, van de Rijn M, Botstein D,<br />

Brown PO. Diversity, topographic differentiation, and positional<br />

memory in human fibroblasts. Proc Natl Acad Sci U S A 2002; 99:<br />

12877–12882.<br />

15. Han X, Amar S. Identification of genes differentially expressed in cultured<br />

human periodontal ligament fibroblasts vs. human gingival fibroblasts<br />

by DNA microarray analysis. J Dent Res 2002; 81:399–405.<br />

16. Kues WA, Anger M, Carnwath JW, Paul D, Motlik J, Niemann H.<br />

Cell cycle synchronization of porcine fetal fibroblasts. Effects of serum<br />

deprivation and reversible cell cycle inhibitors. Biol Reprod 2000;<br />

62:412–419.<br />

17. Kues WA, Carnwath JW, Paul D, Niemann H. Cell cycle synchronization<br />

of porcine fetal fibroblasts by serum deprivation initiates an<br />

unconventional form of apoptosis. Clon Stem Cells 2002; 4:231–243.<br />

18. Oback B, Wells D. Donor cells for nuclear cloning: many are called,<br />

but few are chosen. Clon Stem Cells 2002; 4:147–168.<br />

19. Anderson R, Copeland TK, Scholer H, Heasman J, Wylie C. The onset<br />

of germ cell migration in the mouse embryo. Mech Dev 2000; 91:<br />

61–68.<br />

20. Boiani M, Eckardt S, Scholer HR, McLaughlin KJ. Oct4 distribution<br />

and level in mouse clones: consequences for pluripotency. Genes Dev<br />

2002; 16:1209–1219.<br />

21. Scholer HR, Balling R, Hatzopoulos AK, Suzuki N, Gruss P. Octamer<br />

binding proteins confer transcriptional activity in early mouse embryogenesis.<br />

EMBO J 1989; 8:2543–2550.<br />

22. Yeom YI, Fuhrmann G, Ovitt CE, Brehm A, Ohbo K, Gross M, Hubner<br />

K, Scholer HR. Germline regulatory element of Oct4 specific for the<br />

totipotent cycle of embryonal cells. Development 1996; 122:881–894.<br />

23. Pesce M, Schöler HR. Oct4: gatekeeper in the beginnings of mammalian<br />

development. Stem Cells 2001; 19:271–278.<br />

24. Niwa H, Miyazaki J, Smith A. Quantitative expression of Oct3/4 defines<br />

differentiation, dedifferentiation or self-renewal of ES cells. Nat<br />

Genet 2000; 24:372–376.<br />

25. Rosner MH, Vigano MA, Ozato K, Timmons PM, Poirier F, Rigby<br />

PW, Staudt LM. A POU-domain transcription factor in early stem cells<br />

and germ cells of the mammalian embryo. Nature 1990; 345:686–<br />

692.<br />

26. Gotz J, Probst A, Ehler E, Hemmings B, Kues W. Delayed embryonic<br />

lethality in mice lacking protein phosphatase 2A catalytic subunit Calpha.<br />

Proc Natl Acad Sci U S A 1998; 95:12370–12375.<br />

27. Kues WA, Brenner HR, Sakmann B, Witzemann V. Local neurotrophic<br />

repression of gene transcripts encoding fetal AChRs at rat neuromuscular<br />

synapses. J Cell Biol 1995; 130:949–957.<br />

28. Friedrich G, Soriano P. Promoter traps in embryonic stem cells: a<br />

genetic screen to identify and mutate developmental genes in mice.<br />

Genes Dev 1991; 5:1513–1523.<br />

29. Yayon A, Aviezer D, Safran M, Gross JL, Heldman Y, Cabilly S,<br />

Givol D, Katchalski-Katzir E. Isolation of peptides that inhibit binding<br />

of basic fibroblast growth factor to its receptor from a random phageepitope<br />

library. Proc Natl Acad Sci U S A 1993; 90:10643–10647.<br />

30. Pietrzkowski Z, Wernicke D, Porcu P, Jameson BA, Baserga R. Inhibition<br />

of cellular proliferation by peptide analogues of insulin-like<br />

growth factor 1. Cancer Res 1992; 52:6447–6451.


1028 KUES ET AL.<br />

31. Engstrom U, Engstrom A, Ernlund A, Westermark B, Heldin CH.<br />

Identification of a peptide antagonist for platelet-derived growth factor.<br />

J Biol Chem 1992; 267:16581–16587.<br />

32. Naranda T, Goldstein A, Olsson L. A peptide derived from an extracellular<br />

domain selectively inhibits receptor internalization: target sequences<br />

on insulin and insulin-like growth factor 1 receptors. Proc<br />

Natl Acad Sci U S A 1997; 94:11692–11698.<br />

33. Caplan AI, Bruder SP. Mesenchymal stem cells: building blocks for<br />

molecular medicine in the 21st century. Trends Mol Med 2001; 7:<br />

259–264.<br />

34. Holley RW, Kiernan JA. ‘‘Contact inhibition’’ of cell division in 3T3<br />

cells. Proc Natl Acad Sci U S A 1968; 60:300–304.<br />

35. Todaro GJ, Lazar GK, Green H. The initiation of cell division in a<br />

contact-inhibited mammalian cell line. J Cell Physiol 1965; 66:325–<br />

333.<br />

36. Otsuka H, Moskowitz M. Arrest of 3T3 cells in G1 phase in suspension<br />

culture. J Cell Physiol 1975; 87:213–220.<br />

37. Layer PG, Robitzki A, Rothermel A, Willbold E. Of layers and<br />

spheres: the reaggregate approach in tissue engineering. Trends Neurosci<br />

2002; 25:131–134.<br />

38. Rubin H. Multistage carcinogenesis in cell culture. Dev Biol (Basel)<br />

2001; 106:61–67.<br />

39. Verfaillie C. Adult stem cells: assessing the case for pluripotency.<br />

Trend Cell Biol 2002; 12:502–508.


Bovine ICM Derived Cells Express<br />

the Oct4 Ortholog<br />

MOLECULAR REPRODUCTION AND DEVELOPMENT<br />

PREM S. YADAV, 1,2 WILFRIED A. KUES, 1 DORIS HERRMANN, 1 JOSEPH W. CARNWATH, 1<br />

AND HEINER NIEMANN 1 *<br />

1 Department of Biotechnology, Institute for Animal Breeding (FAL) Mariensee, Neustadt, Germany<br />

2 Department of Physiology and Reproduction, Central Institute for Research on Buffaloes, Hisar, Haryana, India<br />

ABSTRACT The goal of this study was to<br />

define conditions for the successful isolation of embryonic<br />

stem cells from bovine blastocysts. Expression<br />

of the Pit-Oct-Unc (POU) transcription factor Oct4 was<br />

employed to monitor the pluripotent status of cultured<br />

cells. No expression of the previously identified bovine<br />

Oct4 pseudogene was found, and transcription of the<br />

Oct4 ortholog correlated with the proliferative potential<br />

of bovine ICM derived cells. Two methods to isolate<br />

pluripotent inner cell mass were compared; 90% of<br />

trypsin isolated ICMs formed growing cultures,<br />

whereas only 12%–23% of the ICMs isolated by immunosurgery<br />

attached and grew. Colony formation<br />

from complete blastocysts was 55%. The bovine ICM<br />

derived cells could be grown for 4–7 passages. However,<br />

Oct4 transcripts were only present in the primary<br />

cultures, indicating that the initial culture period of<br />

bovine ICM derived cells is critical and needs to be<br />

optimized to yield true ES cells. In contrast to bovine<br />

ICMs, murine ICMs yielded rapidly growing cells,<br />

which proliferated for more than 60 passages. Mol.<br />

Reprod. Dev. ß 2005 Wiley-Liss, Inc.<br />

Key Words: embryonic stem cells; bovine Oct4;<br />

pseudogene<br />

INTRODUCTION<br />

Embryonic stem cells are derived from the inner cell<br />

mass of blastocysts and are pluripotent, that is, can<br />

differentiate into any cell type in the body including<br />

gametes (Weissman, 2000; Rossant, 2001). Application<br />

of these cells to study fundamental events in embryonic<br />

development, to produce therapeutic delivery systems,<br />

and to test drugs in pharmaceutical research has made<br />

this a rapidly growing area of research (Lovell-Badge,<br />

2001). Evans and Kaufman (1981) were the first to<br />

establish a method for growing stem cells from mouse<br />

embryos indefinitely in the laboratory, while Thomson<br />

et al. (1998) developed the first human embryonic stem<br />

cell line. No true pluripotent ES cells with germ line<br />

contribution have been established from mammals<br />

other than mouse (Thomson and Marshall, 1998;<br />

Niemann and Kues, 2000; Saito et al., 2004; Wolf et al.,<br />

2004). Bovine ES-like cells have been reported by<br />

ß 2005 WILEY-LISS, INC.<br />

several laboratories. The time period over which these<br />

cells could be maintained in culture varied from 13<br />

weeks to 3 years and chimera formation has been<br />

described with cells from passages 3–10 (Evans et al.,<br />

1990; Sims and First, 1993; Stice et al., 1996; Cibelli<br />

et al., 1998; Talbot et al., 2000; for review see Gjørret and<br />

Maddox-Hyttel, 2005).<br />

The correct expression level of Oct4 is considered a key<br />

marker for pluripotent cells in the mouse because it is<br />

essential for maintenance of an undifferentiated state<br />

(Niwa et al., 2000; Pesce and Schöler, 2001). Oct4, a<br />

transcription factor of the Pit-Oct-Unc (POU) family, is<br />

expressed in pluripotent cells of preimplantation<br />

embryos and in all cells of the germ line. In the mouse,<br />

Oct4 is present in all embryonic cells up to the morula<br />

stage; and becomes restricted to the ICM cells in<br />

blastocysts (Nordhoff et al., 2001; Pesce and Schöler,<br />

2001). Orthologous genes have been identified in the<br />

genome of humans, primates, and domestic species.<br />

Whereas the bovine Oct4 mRNA follows the same<br />

pattern of expression as in the mouse during preimplantation<br />

development, bovine, and porcine Oct4 proteins<br />

are apparently not restricted to the pluripotent<br />

ICM cells of blastocysts, but have also been detected in<br />

trophectoderm and in all cells of primary blastocyst<br />

outgrowths (van Eijk et al., 1999; Kirchhof et al., 2000;<br />

Kurosaka et al., 2004). Moreover, van Eijk et al. (1999)<br />

isolated a bovine Oct4 pseudogene, which was not<br />

known from the mouse system and RT-PCR data<br />

suggested that the pseudogene might be expressed in<br />

bovine embryos (van Eijk et al., 1999). These latter data<br />

compromised the usefulness of Oct4 as a marker for<br />

pluripotency in livestock species.<br />

The availability of ES cells from farm animals would<br />

permit the extension of studies using murine cells, as<br />

human embryos and ES cells are constrained by legal<br />

and ethical restrictions. A procedure for the isolation of<br />

bovine ES cells would facilitate transgenic approaches<br />

Grant sponsor: Deutsche Forschungsgemeinschaft (DFG).<br />

*Correspondence to: Prof. Heiner Niemann, Department of Biotechnology,<br />

Institute of Animal Breeding (FAL), Mariensee, 31535<br />

Neustadt, Germany. E-mail: niemann@tzv.fal.de<br />

Received 7 February 2005; Accepted 1 June 2005<br />

Published online in Wiley InterScience<br />

(www.interscience.wiley.com).<br />

DOI 10.1002/mrd.20343


2 P.S. YADAV ET AL.<br />

in agriculture and biomedicine (Kues and Niemann,<br />

2004). An important precondition for the isolation of<br />

livestock ES cells is the availability of a suitable marker<br />

for pluripotency. In the present study, bovine ICM cells<br />

were isolated and cultured, and the usefulness of Oct4 as<br />

a bovine stem cell marker was assessed. For comparison,<br />

murine ES cells were derived from mouse blastocysts<br />

and cultured in parallel.<br />

MATERIALS AND METHODS<br />

Isolation of Feeder Cells<br />

A bovine fetus (estimated age 110 days p.c.) was<br />

obtained from a commercial slaughterhouse and kept in<br />

PBS. The fetus was decapitated, eviscerated, and small<br />

(>1 mm 3 ) pieces of sub-dermal tissue were prepared.<br />

These tissue explants were placed in microdrops of recalcified<br />

bovine plasma and incubated for 1 hr to allow<br />

coagulation of the plasma prior to adding culture<br />

medium (87% Dulbecco’s Modified Eagles Medium<br />

(DMEM, PAA, Cölbe, Germany) and 10% fetal calf<br />

serum (batch: 40F2342K, untreated, Gibco, Karlsruhe,<br />

Germany), 1% nonessential amino acids (Sigma, Deisenhofer,<br />

Germany), 1% vitamin solution (Sigma), 1%<br />

antibiotics, 2 mM glutamine, and 0.1 mM mercaptoethanol.<br />

Subconfluent cultures were trypsin-EDTA treated<br />

and subpassaged for cell multiplication (Kues et al.,<br />

2000, 2005). The harvested cells were frozen in aliquots<br />

using a mixture of 90% culture medium and 10%<br />

dimethylsulfoxid; frozen aliquots were thawed and<br />

grown to confluency and then treated with 10 mg/ml<br />

mitomycin C (Sigma) for 3 hr. The mitomycin C treated<br />

cells were frozen in small aliquots, and seeded 1 day<br />

prior usage as feeder cells. Sufficient stocks of the feeder<br />

fibroblasts were created and frozen in liquid nitrogen to<br />

provide an identical feeder layer for all experiments.<br />

In Vitro Production of Bovine Blastocysts<br />

Bovine blastocysts were produced in vitro (Eckert and<br />

Niemann, 1995; Holm et al., 1999; Wrenzycki et al.,<br />

2001). Briefly, ovaries were collected from a local<br />

slaughterhouse and transported to the laboratory in<br />

an insulated container at 25–308C in PBS. Oocytes were<br />

isolated by slicing the ovaries in PBS containing heparin<br />

and 1% BSA. The oocytes were matured in TCM 199<br />

(Sigma) containing 0.2 mM L-glutamine and 25 mM<br />

HEPES. For oocyte maturation, TCM199 containing<br />

0.1% fatty acid free BSA (Sigma, A7030) medium was<br />

supplemented with 10 IU/ml eCG and 5 IU hCG/ml<br />

(Suigonan, Intervet, Germany) for 24 hr in a CO2 incubator. Cumulus oocyte complexes (COC) were<br />

cultured in groups of 15–20 in 100 ml maturation drops<br />

under silicone oil in a humidified atmosphere composed<br />

of 5% CO2 in air at 398C for 24 hr. Matured oocytes were<br />

rinsed in fertilization medium (Fert-TALP supplemented<br />

with 6 mg/ml BSA) and fertilized in Fert-TALP<br />

containing 10 mM hypotaurine (Sigma), 1 mM epinephrine<br />

(Sigma), 0.1 IU/ml heparin (Serva, Heidelberg,<br />

Germany), and 6 mg/ml BSA. Frozen thawed semen<br />

from bulls of proven fertility was used for IVF. The final<br />

sperm concentration added per drop of 100 ml containing<br />

10–15 oocytes was 1 10 6 sperm cells/ml. Fertilization<br />

took place over 18 hr co-culture of sperm and oocytes.<br />

Presumptive zygotes were cultured in 30 ml of synthetic<br />

oviduct fluid (SOF) supplemented with 1% BSA. Embryos<br />

were cultured in mixture of 5% CO 2,7%O 2, 88%<br />

N 2 (Air Products, GmbH Hattingen, Germany) in<br />

modular incubator chambers (615300 ICN Biomedical,<br />

Inc., Aurora OH). Expanded blastocysts were harvested<br />

on day 8 of development (day 0 ¼ IVF).<br />

Isolation of Inner Cell Masses<br />

From Blastocysts<br />

The inner cell masses (ICM) were isolated either by<br />

trypsin treatment or by immunosurgery. In both cases,<br />

the zona pellucida was first removed by incubation in<br />

1.0% pronase (Sigma, P6911) in PBS (w/v) until it<br />

dissolved. For trypsin isolation, zona free blastocysts<br />

were transferred into 0.25% trypsin-EDTA solution and<br />

observed under a microscope until the trophectoderm<br />

cells became loose, and were shed from the ICM by<br />

pipetting. These ICMs were transferred into PBS and<br />

then seeded on feeder cells. For a modified immunosurgical<br />

isolation of ICMs (Solter and Knowles, 1975), the<br />

zona-free blastocysts were washed three times in PBS<br />

with 1% polyvinyalcohol (PVA) (Sigma P 8136) and<br />

placed in 10 mM trinitrobenzene-sulphonic acid (TNBS)<br />

(Sigma P 2297) in PBS containing 3 mg/ml polyvinylpyrrolidone<br />

for 10 min. After washing in 1% PBS-PVA,<br />

they were treated with rabbit anti dinitrophenyl-BSA<br />

(anti DNP BSA 61-006 ICN Biochemicals, Eschwege,<br />

Germany) diluted 3:7 in PBS for 10 min. The anti-DNP-<br />

BSA crossreacts with trinitrophenol groups. Outer<br />

trophectodermal cells were then lysed with guinea pig<br />

complement serum (Sigma S1639, 1:4) for 15 min. The<br />

isolated ICMs were washed with PBS-PVA and then<br />

seeded on feeder layers. The culture medium for bovine<br />

embryonic cells consisted of 77% DMEM, 20% FCS, 1%<br />

nonessential amino acids, 1% vitamin solution, and 1%<br />

penicillin-streptomycin solution. In some experiments,<br />

the culture medium was supplemented with either LIF<br />

(40–80 ng/ml), hIGF (82 ng/ml), TNF-b (0.12 ng/ml), and<br />

bFGF (4 ng/ml). Culture media were changed every<br />

second day.<br />

Suspension cultures were maintained in hanging<br />

drops. Hanging drops were prepared by placing drops<br />

of 30 ml media on the inner surface of the cover of 6-well<br />

plates. The wells of the 6-well plates were filled with PBS<br />

to prevent drying of drops.<br />

Alkaline Phosphatase Staining<br />

Expression of alkaline phosphatase (AP) was analyzed<br />

in blastocysts, ICMs, and cultured cells. The<br />

specimens were washed twice with PBS before fixing<br />

them in 3.7% paraformaldehyde for 10 min. After<br />

washing with PBS, they were incubated in AP staining<br />

solution (25 mM Tris HCl pH 9.0, 150 mM NaCl, 4 mM<br />

MgCl2; 0.4 mg/ml sodium naphthylphosphate; 1.0 mg/<br />

ml fast red) for 1 hr (Kues et al., 2000). AP positive<br />

stained cells appeared red, while negative cells were<br />

colorless or brownish.


RT-PCR Amplification of Selected Transcripts<br />

Poly(A) þ RNA from groups of 5–10 blastocysts was<br />

isolated using a Dynabeads mRNA DIRECT Kit (Dynal,<br />

Norway, Product number 610.11). Nearly confluent ICM<br />

derived cells in 6-well culture dishes at the 2nd, 3rd, and<br />

4th passage (or 24-well dishes with primary colonies)<br />

were washed twice with PBS, and then stored at 808C<br />

prior to isolation of total RNA. Isolation of RNA from<br />

cells was accomplished with Trizol (1 ml/6-well culture<br />

plate) (Trizol Invitrogen Cat No. 15596026). The Trizol<br />

extracts were transferred to 1.5 ml safe lock tubes and<br />

200 ml of chloroform were added, the aqueous and<br />

organic phase were mixed by shaking prior to centrifugation<br />

for 15 min at 9,700g. The supernatant was<br />

collected and 0.5 ml isopropylalcohol was added per ml<br />

Trizol. The samples were then centrifuged for 15 min,<br />

the supernatant was removed and the RNA pellet was<br />

washed with 1 ml 70% cold ethanol. The pellet was air<br />

dried for several minutes and dissolved in 60 ml RNase<br />

free water. The RNA was treated with RNase-free<br />

DNase for removal of DNA, then the RNA concentration<br />

was measured using a Nanodrop spectrophotometer<br />

(ND-1000 spectrophotometer, Rockland, MD).<br />

The mRNAs were reverse transcribed in a reaction<br />

mixture consisting of 1 RT buffer (50 mM KCl, 10 mM<br />

Tris HCl pH 8.3, Perkin Elmer), 5 mM MgCl2, 1 mM each<br />

of dNTP (Amersham, Brunswick, Germany) 20 IU<br />

RNase inhibitor (Perkin Elmer), and 50 U reverse<br />

transriptase (RT, Perkin Elmer) in a total volume of<br />

20 ml using 1.0 ml of50mMrandom hexamer primers<br />

(Perkin-Elmer, Vaterstetten, Germany). The RT reaction<br />

was carried out at 258C for 10 min and then 428C for<br />

1 hr followed by denaturation at 998C for 5 min. The<br />

polymerase chain reaction was performed with the<br />

reverse transcribed product using one or two embryo<br />

equivalents or 10% of the RT-reaction from cultured<br />

cells. The final volume of the PCR reaction mixture was<br />

50 ml (1 PCR Buffer, 20 mM Tris buffer HCl pH 8.4,<br />

1.5 mM MgCl2, 200 mM dNTPs, 0.25 mM of each primer).<br />

RT-PCR products were separated by electrophoresis on<br />

a 2% agarose gel in 1 TBE buffer (90 mM Tris, 90 mM<br />

borate, 2 mM EDTA, pH 8.3 containing 0.2 mg/ml<br />

ethidium bromide). An image of each gel was recorded<br />

using a CCD camera (Quantix, Photometrics, Munich<br />

Germany). Details of primers are given in Table 1. The<br />

amplicon of the control poly(A) polymerase gene does not<br />

contain an intron, therefore the PCR products using<br />

RNA or genomic DNA as templates are identical. For<br />

sequencing, the DNA bands were cut out under long<br />

wave length UV illumination (365 nm), purified using a<br />

DNA gel band purification kit (Amersham Biosciences,<br />

Freiburg, Germany) and sequenced with the Oct-26<br />

primer (van Eijk et al., 1999).<br />

Isolation and Culture of Mouse<br />

Embryo Derived Cells<br />

Mouse blastocysts were collected by flushing the uteri<br />

of superovulated wildtype or Oct4-GFP transgenic<br />

(OG2) mice at day 3.5 after mating. Oct4 positive transgenic<br />

blastocysts were used for culture or ICM isolation.<br />

ICMs were isolated and cultured as described for bovine<br />

embryos.<br />

RESULTS<br />

Expression of the Bovine Oct4 Gene but not<br />

of the Pseudogene in Bovine Blastocysts<br />

Expression of Oct4 was assessed in bovine embryos<br />

and somatic tissues by RT-PCR with splice-site spanning<br />

primers (Fig. 1A). To determine if the obtained<br />

RT-PCR fragments resulted from amplification of the<br />

Oct4 ortholog or the Oct4 pseudogene, PCR products<br />

were isolated and directly sequenced. The sequences of<br />

the Oct4 ortholog (AF022987) and the Oct4 pseudogene<br />

(AF022988, van Eijk et al., 1999) differ by nucleotide<br />

exchanges at positions 754 and 1,245 (numbering<br />

according to AF022987). The obtained sequences from<br />

bovine blastocysts revealed that blastocysts exclusively<br />

expressed the bovine Oct4 ortholog, whereas control<br />

PCRs with bovine genomic DNA isolated from white<br />

blood cells (Fig. 1B), liver, and kidney amplified exclusively<br />

the pseudogene. No amplification of Oct4 fragments<br />

was obtained from mRNAs isolated from white<br />

blood cells, liver, and kidney (Fig. 1A). Thus the presence<br />

of a bovine Oct4 pseudogene could be verified,<br />

however, no evidence of pseudogene transcription was<br />

found. Together with recent data indicating that Oct4<br />

mRNA expression patterns are conserved between<br />

mammalian species (Kurosaka et al., 2004), this<br />

suggests that Oct4 transcription can indeed be used as<br />

a marker for pluripotent cells in cattle.<br />

TABLE 1. Details of Primers Used for Identification of Bovine Oct-4 Expression<br />

Name of primer Primer sequence<br />

Bovine Oct4 Upper 50-GTT CTC TTT GGA AAG GTG TTC<br />

Lower 50-ACA CTC GGA CCA CGT CTT TC<br />

PolyA polymerase Uppper 50-CCT CAC TGC ACA ATA TGC CGT<br />

CTT CAC CTG<br />

STEM CELLS FROM BOVINE AND MURINE BLASTOCYSTS 3<br />

Gene bank<br />

accession<br />

number<br />

Annealing temp.<br />

(PCR cycles) Reference<br />

AF022987 57 (40) Van Eijk et al.,<br />

1999<br />

X 63436 54 (36) Wrenzycki et al.,<br />

2001<br />

Lower 5 0 -AGC CAA TCC ACT GTT CTC CCA<br />

CCT TTA CTC<br />

Mouse Oct4 Upper 5 0 -GAA CAG TTT GCC AAG CTG CTG X 52437 54 (40) Schöler, 1991<br />

Lower 5 0 -CCG GTT ACA GAA CCATAC TCG


4 P.S. YADAV ET AL.<br />

Fig. 1. Oct4 but not its pseudogene is expressed in bovine<br />

blastocysts. A: Lane 1, RT-PCR amplification of Oct4 from bovine<br />

blastocysts; lane 2 and 3, controls without RT and without template;<br />

lanes 4, 6, 8, and 10, RT-PCR of total RNA from bovine white blood cells,<br />

liver, kidney, and lung; lane 11, PCR Oct4-fragment from genomic<br />

DNA; lanes 5, 7, 9 controls without RT; and lane 12 without template.<br />

Control reactions with primers specific for poly(A) polymerase are<br />

shown below. B: The amplified Oct4 products obtained from blastocyst<br />

mRNAs and genomic DNA were isolated (indicated by white boxes in A)<br />

and sequenced in pools. Short stretches of the sequences containing the<br />

Isolation and Initial Cultures<br />

of Bovine ICM Derived Cells<br />

In the first experiment, 20 intact blastocysts and<br />

50 ICMs isolated by trypsin treatment were cultured<br />

on feeder cells. Intact blastocysts hatched prior to attachment,<br />

and the trypsin isolated ICMs regained a<br />

blastocyst-like shape due to outgrowth of the remaining<br />

trophectodermal cells (Fig. 2A). Different trypsin treatment<br />

times and different enzyme concentrations were<br />

tested. Under conditions, which allowed the survival of<br />

regions of nucleotide mismatches (boxed nt positions) between the Oct4<br />

gene and its pseudogene are shown, the numbering is according to<br />

GenBank accession no AF022987. The sequences derived from<br />

blastocysts match 100% with the Oct4 gene. Both nucleotide exchanges<br />

of the described pseudogene were present in genomic DNA at positions<br />

1,245 and 754. For sequence reactions, the reverse Oct-26 primer was<br />

used, therefore, the noncoding strand is shown. C: Control sequences of<br />

bovine Prnp genes of cattle with known single nucleotide polymorphisms<br />

(SNP). Arrows indicate the SNP positions where signal peaks for<br />

both allele variants are present.<br />

ICMs, the removal of trophectodermal cells was not<br />

complete. The remaining trophectodermal cells produced<br />

blastocyst-like structures with diameters of up<br />

to 500–600 mm before attaching to the feeder cells.<br />

Primary colonies formed after 2 weeks and grew as a<br />

single sheet of cells (Fig. 2B). These sheet-like colonies<br />

were physically separated into small pieces for subpassaging<br />

since trypsin treatment was not effective. In<br />

subcultures, the pieces increased in size before attaching<br />

(Fig. 2C). The typical time for each passage was 10–<br />

12 days. It was found that 55% of the intact blastocysts


Fig. 2. Different growth phenotypes of bovine ICM derived cultures.<br />

A–F: Bovine ICMs and initial cultures. (A) Reexpanded bovine trypsinisolated<br />

ICM before attachment (40 ), (B) Trypsin isolated ICM,<br />

passage 2, sheet like colony (40 ), (C) Passage 2 colony after physical<br />

detachment (40 ), (D) Donut shaped growth in LIF supplemented<br />

formed primary colonies, which could be passaged 4–<br />

5 times. By comparison, 90% of the cells isolated from<br />

trypsin digested ICMs formed primary colonies which<br />

could be passaged 6–7 times over a period of 60–65 days<br />

(Table 2) before proliferation ceased.<br />

STEM CELLS FROM BOVINE AND MURINE BLASTOCYSTS 5<br />

When trypsin isolated ICMs were cultured in medium<br />

with LIF (40 ng/ml) supplementation, they grew into<br />

donut shaped structures (Fig. 2D). The size of the<br />

donut-like structure increased to >1 mm and cells of<br />

two different morphologies became visible. Round<br />

TABLE 2. Comparison of Culture Behavior of Embryo Derived Cells From Intact<br />

Bovine Blastocysts and Trypsin-EDTA Isolated ICMs<br />

Starting material No. of Embryos<br />

Primary colony<br />

formation (%) Passages a<br />

Blastocysts 20 11 (55%) >4<br />

ICMs (trypsin isolated) 50 45 (90%) >7<br />

a Sub-passaging was done when the cells reached 50%–70% confluency.<br />

medium (40 ), (E) Bovine ICM derived by immunosurgery, attached on<br />

feeder cells (200 ), (F) Primary colony growth from immunoisolated<br />

ICM. G–H: Murine ICMs and derived cell cultures: (G) Murine trypsin<br />

isolated ICM, day 13 of attachment (40 ), (H) Proliferating ICM<br />

derived cells at passage 6.


6 P.S. YADAV ET AL.<br />

individual cells were present on the bottom of the<br />

structure, while the surface was composed of tightly<br />

bound cells.<br />

Immunosurgically Isolated ICMs<br />

In the second experiment, bovine ICMs were isolated<br />

by immunosurgery and cultured in DMEM/20% FCS<br />

supplemented with (1) LIF (40 ng/ml), hIGF(2 ng/ml),<br />

TNF-ß (0.12 ng/ml), and bFGF (4 ng/ml); (2) LIF (40 ng/<br />

ml); (3) LIF (80 ng/ml); or (4) without growth factor<br />

supplement. The rate of primary colony formation and<br />

the proportion of proliferating colonies are presented in<br />

Table 3.<br />

Immunosurgically isolated ICMs attached within 3–<br />

4 days after seeding on the feeder layer in all tested<br />

media (Fig. 2E). The initial colonies remained very small<br />

for 1 week, but after 2 weeks in culture, large primary<br />

colonies could be observed (Fig. 2F). In the group grown<br />

in medium containing all four growth factors, 11.5% of<br />

the seeded ICMs produced primary colonies. The group<br />

supplemented with LIF (40 ng/ml) had the highest<br />

percentage of colony formation (21.5%). The morphology<br />

of cells in these colonies was different from that of<br />

enzyme digested ICMs or intact blastocyst colonies.<br />

Trypsin could be used to detach these colonies and the<br />

cultures were passaged after gentle physical separation.<br />

Small clumps of cells were better at attachment and<br />

growth after passaging than individual cells. Most<br />

cells proliferated well in the first four passages and<br />

then grew at a slower rate up to the 6th passage, finally<br />

arresting after 50–60 days. RT-PCR analysis revealed<br />

that only the initial cultures expressed Oct4 mRNA,<br />

from the second passage Oct4 mRNA was no longer<br />

detectable (Fig. 3A,B). The average passage time was<br />

10–14 days. In hanging drops, the floating cells did not<br />

proliferate.<br />

Permanent Culture of Murine<br />

ICM-derived Cells<br />

Day 3.5 p.c. murine blastocysts from wildtype and<br />

Oct4-EGFP (OG2) transgenic mice served as controls to<br />

validate the ICM isolation protocols and culture conditions.<br />

Results were identical for transgenic and nontransgenic<br />

embryos, and ICMs (Table 4). Attachment<br />

rates and primary colony formation for trypsin isolated<br />

murine ICMs were higher than for intact blastocysts.<br />

GFP expression was observed in cells from the OG2transgenic<br />

embryos, but started to diminish from day 4<br />

of culture. In unattached blastocysts, GFP was visible<br />

for up to 10 days in culture. Primary colonies formed<br />

after 10–13 days of culture starting with intact embryos<br />

or ICMs (Fig. 2G). Primary colonies in 24-well plates<br />

were transferred to 6-well plates and then to 25 cm<br />

flasks for amplification (Fig. 2H). ICM derived cells from<br />

three different batches survived for 62, 58, 57 passages,<br />

respectively and could be maintained in culture for more<br />

than 225 days (Table 5). In hanging drops, the cells<br />

formed loose embryoid bodies within 1 week. Without a<br />

feeder cell layer, embryonic cells could attach to the<br />

plastic but growth was slow. Murine ICM derived cells<br />

were negative for AP staining, but expressed Oct4<br />

mRNA at passages 41 and 46 (Fig. 3C).<br />

DISCUSSION<br />

The transcription factor Oct4 is essential for the germ<br />

cell lineage in the mouse where it is required at critical<br />

levels to maintain embryonic stem cells in a pluripotent<br />

state (Niwa et al., 2000; Buehr et al., 2003). The bovine<br />

Oct4 shares high sequence homology with its mouse,<br />

primate, and human orthologs (Nordhoff et al., 2001;<br />

Mitalipova et al., 2003), and has 90.6% and 81.7%<br />

overall identity at the protein level to human and<br />

murine orthologs, respectively. However, in bovine cells,<br />

the usefulness of Oct4 has been questioned by the<br />

identification of a bovine Oct4 pseudogene and suggestions<br />

that this pseudogene might be transcribed in<br />

bovine embryos (van Eijk et al., 1999). The apparent<br />

discrepancy to the data presented here might be due to<br />

the fact that van Eijk et al. (1999) used an indirect<br />

method to discriminate between the pseudogene and<br />

Oct4 mRNA. They employed a mutagenised PCR primer<br />

to introduce a restriction site in the Oct4 amplicon,<br />

followed by EagI digest and gel analysis. In the present<br />

study, Oct4 expression was detected by RT-PCR with<br />

completely matching primers. Direct sequencing of the<br />

PCR product allowed unequivocal discrimination<br />

between the Oct4 ortholog and pseudogene, as they<br />

differ by nucleotide exchanges at positions 754 and<br />

1,245. Our sequence data revealed that bovine blastocysts<br />

expressed exclusively the Oct4 ortholog; expression<br />

of the pseudogene could not be detected in<br />

embryonic or adult tissues such as liver, kidney, and<br />

white blood cells. In contrast, control experiments with<br />

genomic DNA from liver, kidney, and white blood cells,<br />

resulted in successful amplification of the pseudogene,<br />

confirming presence of the pseudogene in the bovine<br />

genome and demonstrating that the PCR conditions are<br />

suitable and sensitive for its detection. Recently, it was<br />

TABLE 3. Proliferation of Bovine Immunosurgically Isolated ICMs in the<br />

Presence of Various Growth Factors<br />

Media<br />

supplements<br />

No. of<br />

trials<br />

No. of<br />

ICMs<br />

No. of<br />

primary<br />

colonies<br />

% primary<br />

colonies<br />

No. of<br />

passages<br />

Days in<br />

culture<br />

LIF, IGF, TNFb, bFGF 11 78 9 11.5 4–6 65<br />

LIF (40 ng/ml) 13 87 19 21.8 4–6 57<br />

LIF (80 ng/ml) 3 33 6 18.2 3 30<br />

None 12 79 13 16.4 4–6 56


Fig. 3. Expression of Oct4 in ICM derived cell cultures. A: Oct4 is<br />

expressed in passage 1 bovine ICM cultures. M, 100 bp ladder; lane 1,<br />

passage 1 bovine ICM cells (day 14 of culture); lane 2 and 3, controls<br />

without RT or template; lane 4, Oct4 PCR with bovine genomic DNA;<br />

lane 5, control without template. B: RT-PCR of Oct4 from passages 2–4.<br />

M, 100 bp ladder, lanes 1, 3, 5 passage 2–4 bovine ICM cells; lanes 2, 4,<br />

6, controls without RT; lane 7, control without template; lane 8, Oct4<br />

shown that the Oct4 mRNA pattern is restricted to ICM<br />

cells in bovine blastocysts as in mice (Kurosaka et al.,<br />

2004). The more widespread distribution of the Oct4<br />

protein in ICM and trophectodermal cells of bovine<br />

blastocysts (van Eijk et al., 1999; Kirchhof et al., 2000)<br />

possibly reflects the prolonged bovine preimplantation<br />

phase compared to that of the mouse. Recent data show<br />

that the Oct4 protein becomes restricted to the embryonic<br />

disc of hatched day 12 bovine embryos which<br />

confirms this view (Gjørret and Maddox-Hyttel, 2005).<br />

Taken together, these observations make Oct4 transcription<br />

a suitable marker for the identification of<br />

pluripotent cells in cattle. During culture of bovine<br />

ICMs, expression of Oct4 mRNA was rapidly lost after<br />

STEM CELLS FROM BOVINE AND MURINE BLASTOCYSTS 7<br />

PCR with bovine genomic DNA; lane 9, control without template. C: RT-<br />

PCR of Oct4 from murine ICM derived cultures employing mouse<br />

specific primers. M, 100 bp ladder, lanes 1 and 2, Oct4 RT-PCR from<br />

passages 41 and 47, lane 3, control without RT; lane 4, control without<br />

template. In parallel, control RT-PCRs with poly(A) polymerasespecific<br />

primers were performed (bottoms).<br />

the first passage, which correlates well with their<br />

limited capacity for self renewal and proliferation.<br />

AP activity has previously been used to identify<br />

pluripotent cells in cultures from livestock species<br />

(Talbot et al., 2000, 1993; Li et al., 2003). In the present<br />

study, intact bovine and murine blastocysts were<br />

positive for AP activity. However, we found that AP<br />

activity was weak or absent in isolated bovine ICMs and<br />

in bovine ICM derived cell cultures. In some colonies of<br />

later passages, a few AP positive cells could be identified.<br />

A similar heterogeneous staining pattern of AP activity<br />

in bovine ICM derived cells had been reported previously<br />

(van Stekelenburg-Hamers et al., 1995; Gjørret<br />

and Maddox-Hyttel, 2005). These observations suggest<br />

TABLE 4. Murine Embryo Derived Cells: Efficiency of Isolation and Development<br />

Starting material No. No. of attached (%) Primary Colony (%)<br />

Transgenic blastocysts 12 3 (25%) 3 (25%)<br />

Nontransgenic blastocysts 10 6 (60%) 4 (40%)<br />

Transgenic ICM 24 17 (91%) 17 (91%)<br />

Nontransgenic ICM 16 16 (100) 14 (87%)


8 P.S. YADAV ET AL.<br />

TABLE 5. Maintenance of Murine Cells in Culture on<br />

Bovine Feeders<br />

Cultures<br />

Number of<br />

days<br />

Number of<br />

passages<br />

Av. time/passage<br />

(days)<br />

1 234 62 3.8<br />

2 227 57 4.0<br />

3 220 58 3.8<br />

that AP expression is not a robust marker for pluripotency<br />

in bovine ICM derived cells.<br />

In the present study, two different methods, trypsin<br />

digestion and immunosurgery, were tested for their<br />

suitability for ICM isolation from bovine blastocysts.<br />

Trypsin digestion had previously been shown to effectively<br />

remove porcine trophectoderm cells (Li et al.,<br />

2003). In the present study, murine and bovine ICMs<br />

isolated with trypsin showed an increased attachment<br />

rate and improved primary colony formation as compared<br />

to intact blastocysts. This substantiates recent<br />

findings for porcine embryos (Li et al., 2003). The trypsin<br />

digestion method yielded a population of murine cells<br />

which could be continuously cultured, while the bovine<br />

system resulted in sheet shaped colonies, in which the<br />

cells had the appearance of trophectoderm, suggesting<br />

that lysis of trophectodermal cells was incomplete.<br />

Trypsin has been reported to be ineffective for<br />

detaching primary ES cell colonies for initial passaging<br />

(Stice et al., 1996; Talbot et al., 2000), and such cultures<br />

are commonly passaged by physical dissociation. In the<br />

present study, primary colonies derived from trypsin<br />

digested ICMs, contained cells of two distinct morphologies.<br />

Individual, round cells were only found in the<br />

central parts of the colonies. Immunosurgery is the<br />

conventional method for isolating ICMs from blastocysts<br />

(Thomson and Marshall, 1998). The morphology of the<br />

cells obtained in the present study after immunosurgery<br />

was quite different from that of cells derived from<br />

trypsin digestion in that individual cells could be easily<br />

distinguished by phase contrast microscopy.<br />

Four different culture media were used after immunosurgical<br />

isolation of ICMs to explore the effect of growth<br />

factors. For immunosurgically isolated ICMs, growth<br />

factors caused variation in attachment and growth, but<br />

only LIF seemed to stimulate colony formation. However,<br />

one has to bear in mind that the medium also<br />

contained fetal calf serum, which may have masked<br />

effects of individual growth factors. When all growth<br />

factors were used together, this resulted in low-primary<br />

colony formation rates, showing that the mixture of<br />

growth factors including LIF did not improve the growth<br />

of cells from the ICMs. The morphology of cells after<br />

growth factor supplementation with or without LIF<br />

remained similar. Recently, a synergistic effect between<br />

bone morphogenetic protein 4 (BMP4) and LIF was<br />

found in maintaining the self renewal of murine ES cells<br />

(Qi et al., 2004). This makes BMP4 a good candidate<br />

molecule for testing on bovine ICM derived cells.<br />

The culture of murine and human stem cells usually<br />

requires the presence of feeder cells (Thomson and<br />

Marshall, 1998; Richards et al., 2002; Amit et al., 2004).<br />

Feeder cells presumably produce growth factors and<br />

other substances that support proliferation and prevent<br />

the differentiation of pluripotent cultured cells (Anderson,<br />

1992). Livestock ES-like cells are also commonly<br />

cultured on feeder cells because the molecular pathways<br />

and key molecules required to maintain pluripotency in<br />

these species are unknown (Wolf et al., 2004). In the<br />

present study, mitomycin-treated bovine fetal fibroblasts<br />

(BFF) from explant cultures (Kues et al., 2005)<br />

were used as a feeder layer. These feeder cells supported<br />

the attachment of intact embryos, ICMs and the proliferation<br />

of bovine ICM cell colony formation for 6–7<br />

passages and murine ICM cell proliferation for more<br />

than 60 passages. Previously, murine embryonic fibroblasts<br />

and STO cells or bovine epithelial layers were<br />

used as feeders for bovine ICM cells (Saito et al., 1992;<br />

van Stekelenburg-Hamers et al., 1995). Successful use<br />

of BFF cells to culture murine ICM cells indicates that<br />

BFF cells could potentially be used as feeder cells for<br />

homologous as well as heterologous ES cells.<br />

In contrast to the limited proliferation of bovine ICM<br />

derived cells, cells from murine blastocysts could be<br />

grown well for more than 200 days over 60 passages on<br />

the same feeder layers and under identical culture<br />

conditions. In the mouse, ES cells appear to arise from<br />

the epiblast cells (Brook and Gardner, 1997); however,<br />

attempts to isolate ES cells from bovine embryonic<br />

discs of day 12 embryos did not result in permanent cell<br />

lines (Gjørret and Maddox-Hyttel, 2005). AP activity in<br />

murine ES cells was negative but the cells expressed<br />

Oct4 in the 41th and 46st passage. The cells formed<br />

embryoid bodies in suspension culture and differentiated<br />

into various cell types when cultured without<br />

feeder cells.<br />

In summary, bovine embryonic fibroblast cells can be<br />

used as feeder cells for both bovine and murine ICM<br />

cells. However, bovine ICM derived cells proliferated for<br />

a limited period. The bovine cells expressed Oct4 only in<br />

primary colonies for up to 2 weeks whereas murine cells<br />

still expressed Oct4 at passage 47, suggesting that the<br />

initial culture conditions for bovine ICMs are critical<br />

and require optimization before permanent bovine ES<br />

cells can be obtained.<br />

ACKNOWLEDGMENTS<br />

The overseas fellowship from the Department of<br />

Biotechnology, the Ministry of Science and Technology<br />

(Government of India) for PY, and funding by the<br />

Deutsche Forschungsgemeinschaft (DFG) to H.N. is<br />

gratefully acknowledged. We are grateful to Karin<br />

Korsawe and Erika Lemme for excellent technical help<br />

to Dieter Bunke for photographical artwork and to Hans<br />

Schöler (Münster) for the gift of OG2 mice.<br />

REFERENCES<br />

Amit M, Shariki C, Margulets V, Itskovitz-Eldor J. 2004. Feeder layerand<br />

serum-free culture of human embryonic stem cells. Biol Reprod<br />

70:837–845.


Anderson GB. 1992. Isolation and use of embryonic stem cells from<br />

livestock species. Anim Biotech 3:165–175.<br />

Brook FA, Gardner RL. 1997. The origin and efficient derivation of<br />

embryonic stem cells in the mouse. Proc Natl Acad Sci USA 94:5709–<br />

5712.<br />

Buehr M, Nichols J, Stenhouse F, Mountford P, Greenhalg C,<br />

Kantachuvesiri S, Brooker G, Mullins S, Smith AG. 2003. Rapid loss<br />

of OCT-4 and pluripotency in cultured rodents and blastocysts<br />

derived cell lines. Biol Reprod 68:222–229.<br />

Cibelli J, Stice SL, Golueke OPJ, Kane JJ, Jerry J, Blackwell W, Ponce<br />

de Leon FA, Robl JM. 1998. Transgenic bovine chimeric offspring<br />

produced from somatic cell-derived stem-like cells. Nat Biotechnol<br />

16:642–646.<br />

Eckert J, Niemann H. 1995. In vitro maturation, fertilization, and<br />

culture to blastocysts of bovine oocytes in protein-free media.<br />

Theriogenology 43:1211–1225.<br />

Evans MJ, Kaufman MH. 1981. Establishment in culture of pluripotential<br />

cells from mouse embryos. Nature 294:154–156.<br />

Evans MJ, Notarianni E, Laurie S, Moor RM. 1990. Derivation and<br />

primary characterization of pluripotent cell lines from porcine and<br />

bovine blastocysts. Theriogenology 33:125–128.<br />

Gjørret JO, Maddox-Hyttel P. 2005. Attempts towards derivation and<br />

establishment of bovine embryonic stem cell-like cultures. Reprod<br />

Fertil Dev 17:113–124.<br />

Holm P, Booth PJ, Schmidt MH, Greve T. 1999. High bovine blastocyst<br />

development in a static in vitro production system using SOF as<br />

medium supplemented with sodium lactate and myo-inositol with or<br />

without serum proteins. Theriogenology 52:683–700.<br />

Kirchhof N, Carnwath JW, Lemme E, Anastassiadis K, Scholer H,<br />

Niemann H. 2000. Expression pattern of Oct-4 in preimplantation<br />

embryos of different species. Biol Reprod 69:1698–1705.<br />

Kues WA, Niemann H. 2004. The contribution of farm animals to<br />

human health. Trends Biotech 22:286–294.<br />

Kues WA, Anger M, Carnwath JW, Paul D, Motlik J, Niemann H. 2000.<br />

Cell cycle synchronization of porcine fetal fibroblasts: Effects of<br />

serum deprivation and reversible cell cycle inhibitors. Biol Reprod<br />

62:412–419.<br />

Kues WA, Petersen B, Mysegades W, Carnwath JW, Niemann H. 2005.<br />

Isolation of murine and porcine fetal stem cells from somatic tissue.<br />

Biol Reprod 72:1020–1028.<br />

Kurosaka S, Eckhardt S, McLaughlin KJ. 2004. Pluripotent lineage<br />

definition in bovine embryos by Oct4 transcript localization. Biol<br />

Reprod 71:1578–1582.<br />

Li M, Zhang D, Hou Y, Jiao L, Zheng X, Wang WH. 2003. Isolation and<br />

culture of embryonic stem cells from porcine blastocyst. Mol Reprod<br />

Dev 65:429–434.<br />

Lovell-Badge R. 2001. Future for stem cells research. Nature 414:<br />

88–97.<br />

Mitalipova SM, Kuo HC, Hennebold JD, Wolf DP. 2003. Oct-4 expression<br />

in pluripotent cells of the rhesus monkey. Biol Reprod 69:1785–<br />

1792.<br />

Niemann H, Kues WA. 2000. Transgenic livestock: Premises and<br />

promises. Anim Reprod Sci 60–61:277–293.<br />

Niwa H, Miyazaki J, Smith AG. 2000. Quantitative expression of Oct3/4<br />

defines differentiation, dedifferentiation or self-renewal of ES cells.<br />

Nat Genet 24:372–376.<br />

STEM CELLS FROM BOVINE AND MURINE BLASTOCYSTS 9<br />

Nordhoff V, Hübner K, Bauer A, Orlova I, Malapesta A, Schöler HR.<br />

2001. Comparative analysis of human, bovine and murine Oct 4<br />

upstream sequences. Mamm Genome 12:309–317.<br />

Pesce M, Schöler HR. 2001. Oct-4: Gate keeper in beginning of<br />

mammalian development. Stem Cells 19:271–278.<br />

Qi X, Li TG, Hao J, Hu J, Wang J, Simmons H, Miura S, Mishina Y,<br />

Zhao GQ. 2004. BMP4 supports self renewal of embryonic stem cells<br />

by inhibiting mitogen-activated protein kinase pathways. Proc Natl<br />

Acad Sci USA 101:6027–6032.<br />

Richards M, Fong CY, Chan WK, Wong PC, Bongso A. 2002.<br />

Human feeders support prolonged undifferentiated growth of human<br />

inner masses and embryonic stem cells. Nat Biotechnol 20:933–<br />

936.<br />

Rossant J. 2001. Stem cells from the mammalian blastocst. Stem Cells<br />

19:477–482.<br />

Saito S, Strelchenko N, Niemann H. 1992. Bovine embryonic stem like<br />

cells cultured over several passages. Roux’s Arch Dev Biol 21:134–<br />

141.<br />

Saito S, Liu B, Yokohama K. 2004. Animal embryonic stem (ES) cells:<br />

Self renewal, pluripotency, transgenesis and nuclear transfer.<br />

Human Cell 17:7–15.<br />

Schöler HR. 1991. Octamania: the POU factors in murine development.<br />

Trends Genet 7:323–329.<br />

Sims MM, First NL. 1993. Production of fetuses from totipotent<br />

cultured bovine inner cell mass cells. Theriogenology 39:313.<br />

Solter D, Knowles BB. 1975. Immunosurgery of mouse blastocysts.<br />

Proc Natl Acad Sci USA 72: 5099–5102.<br />

Stice SL, Strelchenko NS, Keefer CL, Matthews L. 1996. Pluripotent<br />

bovine embryonic cell lines direct embryonic development following<br />

nuclear transfer. Biol Reprod 54:100–110.<br />

Talbot NC, Caperna J, Edwards JL, Gerret W, Wells KD, Ealy AD.<br />

2000. Bovine blastocyst derived trophectoderm and endoderm cell<br />

cultures: Interferon TAU and transferrin expression as respective<br />

in vitro markers. Biol Reprod 62:235–247.<br />

Thomson JA, Marshall VS. 1998. Primate embryonic stem cells. Curr<br />

Top Dev Biol 38:133–165.<br />

Thomson JA, Itskovitz-Elder J, Shapiro SS, Waknitz MA, Swiergiel JJ,<br />

Marshall VS, Jones JM. 1998. Embryonic stem cell line derived from<br />

human <strong>bei</strong>ngs. Science 282:1145–1147.<br />

van Eijk MJ, van Rooijen MA, Modina S, Scesi L, Folkers G, van Tol HT,<br />

Bevers MM, Fisher SR, Lewin HA, Rakacolli D, Galli C, de Vaureix C,<br />

Trounson AO, Mummery CL, Gandolfi F. 1999. Molecular cloning,<br />

genetic mapping, and developmental expression of Bovine POU5F1.<br />

Biol Reprod 60:1093–1103.<br />

Van Stekelenburg-Hamers AE, Van Achterberg TA, Rebel HG, Flechon<br />

JE, Campbell KH, Weima SM, Mummery CL. 1995. Isolation and<br />

characterisation of permanent cell lines from inner cell mass cells<br />

from bovine blastocysts. Mol Reprod Dev 40:444–454.<br />

Weissman IL. 2000. Stem cells: Units of development, units of<br />

regeneration and units of evolution. Cell 100:157–168.<br />

Wolf DP, Kuo HC, Pau KY, Lester L. 2004. Progress with nonhuman<br />

primate embryonic stem cells. Biol Reprod 71:1766–1771.<br />

Wrenzycki C, Herrmann D, Keskintepe L, Martins A , Jr., Sirisathien<br />

S, Brackett B, Niemann H. 2001. Effects of culture system and<br />

protein supplementation on mRNA expression in pre-implantation<br />

bovine embryos. Hum Reprod 16:893–901.


Rev. sci. tech. Off. int. Epiz., 2005, 24 (1), 285-298<br />

Transgenic farm animals: present and future<br />

Introduction: evolution of<br />

transgenic technologies<br />

The first transgenic livestock were produced almost<br />

20 years ago by microinjection of foreign deoxyribonucleic<br />

acid (DNA) into the pronuclei of zygotes (33). However, as<br />

microinjection has several significant shortcomings –<br />

including low efficiency, random integration and variable<br />

expression patterns related to the site of integration –<br />

research has focused on alternative methodologies for<br />

improving efficiency of generating transgenic livestock<br />

(Table I). These methodologies include:<br />

– sperm-mediated DNA transfer (12, 54, 55)<br />

– intracytoplasmic injection of sperm heads carrying<br />

foreign DNA (71, 72)<br />

– injection or infection of oocytes and/or embryos by<br />

different types of viral vectors (11, 35, 37)<br />

– ribonucleic acid (RNA) interference technology<br />

(small interfering RNA: siRNA) (14)<br />

– the use of nuclear transfer (5, 13, 21, 53, 86).<br />

H. Niemann, W. Kues & J.W. Carnwath<br />

Department of Biotechnology, Institute for Animal Breeding (FAL), Mariensee, 31535 Neustadt, Germany<br />

Summary<br />

Until recently, pronuclear microinjection of deoxyribonucleic acid (DNA) was the<br />

standard method for producing transgenic animals. This technique is now <strong>bei</strong>ng<br />

replaced by more efficient protocols based on somatic nuclear transfer that also<br />

permit targeted genetic modifications. Lentiviral vectors and small interfering<br />

ribonucleic acid technology are also becoming important tools for transgenesis.<br />

Transgenic farm animals are important in human medicine as sources of<br />

biologically active proteins, as donors in xenotransplantation, and for research<br />

in cell and gene therapy. Typical agricultural applications include improved<br />

carcass composition, lactational performance and wool production, as well as<br />

enhanced disease resistance and reduced environmental impact. Product safety<br />

can be ensured by standardisation of procedures and monitored by polymerase<br />

chain reaction and array technology. As sequence information and genomic<br />

maps of farm animals are refined, it becomes increasingly practical to remove or<br />

modify individual genes. This approach to animal breeding will be instrumental in<br />

meeting global challenges in agricultural production in the future.<br />

Keywords<br />

Agricultural application – Environmental benefit – Farm animal – Gene pharming –<br />

Lentiviral vector – Microinjection – Nuclear transfer – Small interfering ribonucleic acid<br />

– Transgenic – Xenotransplantation.<br />

To date, somatic nuclear transfer, which has been<br />

successful in ten species – al<strong>bei</strong>t at low efficiency (47, 103)<br />

– holds the greatest promise for significant improvements<br />

in the generation of transgenic livestock (Table I). Further<br />

qualitative improvements may be derived from<br />

technologies that allow precise modifications of the<br />

genome, including targeted chromosomal integration by<br />

site-specific DNA recombinases, such as Cre or flippase<br />

(FLP), or methods that allow temporally and/or spatially<br />

controlled transgene expression (9, 45). The first genomes<br />

of farm animals (cattle, chicken) were sequenced and<br />

annotated in 2004 (2, 3). Thus, after approximately<br />

7,000 years of domestic animal selection (16) based on the<br />

random mutations caused by radiation and oxidative<br />

injury to the genome, technology is now available to<br />

introduce or remove known genes with known functions.<br />

Despite the inherent inefficiency of microinjection<br />

technology, a broad spectrum of genetically modified large<br />

animals has been generated for applications in agriculture<br />

and biomedicine (Table II). The use of transgenic livestock<br />

for ‘gene pharming’ has now reached the level of<br />

commercial exploitation (47). This paper provides a brief<br />

summary of the current status of transgenic animal<br />

production and look at future implications. The authors


286<br />

focus on large domesticated species and do not cover the<br />

recent developments in poultry breeding or aquaculture.<br />

Biomedical applications of<br />

transgenic domestic animals<br />

Pharmaceutical production in the mammary<br />

gland of transgenic animals<br />

Gene ‘pharming’ entails the production of recombinant<br />

biologically active human proteins in the mammary glands<br />

of transgenic animals. This technology overcomes the<br />

Table I<br />

Evaluation of gene transfer methodology*<br />

Methodology<br />

Rev. sci. tech. Off. int. Epiz., 24 (1)<br />

limitations of conventional and recombinant production<br />

systems for pharmaceutical proteins (63, 82) and has<br />

advanced to the stage of commercial application (27, 107).<br />

The mammary gland is the preferred production site,<br />

mainly because of the quantities of protein that can be<br />

produced in this organ using mammary gland-specific<br />

promoter elements and established methods for extraction<br />

and purification of that protein (63, 82).<br />

Guidelines developed by the Food and Drug Administration<br />

(FDA) in the United States of America (USA) require<br />

monitoring of the animals’ health, sequence validation of the<br />

gene construct, characterisation of the isolated recombinant<br />

protein, and monitoring of the genetic stability of the<br />

Animal Site of Integration Homologous Multiple Maximum construct Founder Technical<br />

usage integration prescreening recombination transgenes size limitation mosaicism difficulty<br />

Microinjection High Random No No Inefficient Artificial chromosomes Yes<br />

Somatic nuclear transfer Low Prescreened Yes Possible Yes Transfection limited/<br />

artificial chromosomes<br />

No<br />

Implantation of transgenic<br />

spermatogonia<br />

Low Random Yes Not yet Not yet Transfection limited No<br />

Lentivirus injection Low Random No No Yes ~ 5 kb Yes<br />

Sperm-mediated gene<br />

transfer method<br />

Low Random No No Inefficient Not known Not known<br />

* all grading is based on the fully matured technology<br />

Table II<br />

Current status of transgenesis in livestock<br />

Area of application<br />

Biomedicine<br />

Gene pharming<br />

Antibody production<br />

Xenotransplantation<br />

Solid organs<br />

Cells/tissues<br />

Blood replacement<br />

Disease model<br />

Agriculture<br />

Carcass composition<br />

Leaner meat<br />

Healthful lipids<br />

Lactational performance<br />

Environmental impact<br />

Wool production<br />

Disease resistance<br />

Fertility<br />

Early experimental Experimental Advanced Transition experiments/ Practical<br />

phase stage experimental stage practical use use


Rev. sci. tech. Off. int. Epiz., 24 (1) 287<br />

transgenic animals over several generations. This has<br />

necessitated, for example, the use of animals from scrapiefree<br />

countries (New Zealand) and the maintenance of<br />

production animals under strict hygienic conditions.<br />

Products derived from the mammary gland of transgenic<br />

goats and sheep, such as antithrombin III (ATIII),<br />

α-antitrypsin or tissue plasminogen activator (tPA), have<br />

progressed to advanced clinical trials (47). Phase III trials<br />

for a recombinant human ATIII product have been<br />

completed and an application has been filed for European<br />

Market Authorisation. The protein is expected to be<br />

registered and on the market by the end of 2005. The<br />

product is employed for the treatment of heparin-resistant<br />

patients undergoing cardiopulmonary bypass procedures.<br />

At the same company that manufactures this product,<br />

11 transgenic proteins have been expressed in the<br />

mammary gland of transgenic goats at more than one gram<br />

per litre. The enzyme α-glucosidase from the milk of<br />

transgenic rabbits has orphan drug status and has been<br />

successfully used for the treatment of Pompe’s disease (94)<br />

(in the USA the term ‘orphan drug’ refers to a product that<br />

treats a rare disease affecting fewer than<br />

200,000 Americans). Similarly, recombinant C1 inhibitor<br />

produced in the milk of transgenic rabbits has completed<br />

phase III trials and is expected to receive registration within<br />

the next 24 months. A topical antibiotic against Streptococcus<br />

mutans, for prevention and treatment of dental caries, has<br />

completed phase III trials and should be launched shortly.<br />

The global market for recombinant proteins from domestic<br />

animals is expected to exceed US$1 billion in 2008 and to<br />

reach US$18.6 billion in 2013.<br />

Some gene constructs have failed to produce economically<br />

significant amounts in the milk of transgenic animals,<br />

indicating that the technology needs further refinements to<br />

achieve high-level expression. This is particularly true for<br />

genes that have complex regulation, such as those coding<br />

for erythropoietin or human clotting factor VIII (36, 41,<br />

62, 67). With the advent of transgenic crops that produce<br />

pharmacologically active proteins, there is now an array of<br />

recombinant technologies that will allow selection of the<br />

most appropriate production system for each required<br />

protein (58). The production of edible vaccines in<br />

transgenic crop plants against, for example, foot and<br />

mouth disease, might become an important application for<br />

animal health (102).<br />

Antibody production in transgenic animals<br />

Numerous monoclonal antibodies are <strong>bei</strong>ng produced in<br />

the mammary gland of transgenic goats (63). Cloned<br />

transgenic cattle produce a recombinant bispecific<br />

antibody in their blood (31). Purified from serum, the<br />

antibody is stable and mediates target cell-restricted T cell<br />

stimulation and tumour cell killing. An interesting new<br />

development is the generation of trans-chromosomal<br />

animals. A human artificial chromosome containing the<br />

complete sequences of the human immunoglobulin heavy<br />

and light chain loci was introduced into bovine fibroblasts,<br />

which were then used in nuclear transfer. Transchromosomal<br />

bovine offspring were obtained that<br />

expressed human immunoglobulin in their blood. This<br />

system could be a significant step forward in the<br />

production of human therapeutic polyclonal antibodies<br />

(51). Further studies will show whether the additional<br />

chromosome will be maintained over future generations<br />

and how stable expression will be.<br />

Blood replacement<br />

Functional human haemoglobin has been produced in<br />

transgenic swine. The transgenic protein could be purified<br />

from the porcine blood and showed oxygen-binding<br />

characteristics similar to natural human haemoglobin. The<br />

main obstacle was that only a small proportion of porcine<br />

red blood cells contained the human form of haemoglobin<br />

(90). Alternative approaches to produce human blood<br />

substitutes have focused on the chemical cross-linking of<br />

haemoglobin to the superoxide-dismutase system (20).<br />

Xenotransplantation of porcine organs to<br />

human patients<br />

Today more than 250,000 people are alive only because of<br />

the successful transplantation of an appropriate human<br />

organ (allotransplantation). However, progress in organ<br />

transplantation technology has led to an acute shortage of<br />

appropriate organs, and cadaveric or live organ donation<br />

does not meet the demand in Western societies. To close<br />

the growing gap between demand and availability of<br />

appropriate human organs, porcine xenografts from<br />

domesticated pigs are considered to be the best alternative<br />

(47, 77). Essential prerequisites for successful<br />

xenotransplantation are:<br />

– overcoming the immunological hurdles<br />

– preventing the transmission of pathogens from the donor<br />

animal to the human recipient<br />

– ensuring the compatibility of the donor organs with<br />

human anatomy and physiology.<br />

The major immunological obstacles in porcine-to-human<br />

xenotransplantation are:<br />

– hyperacute rejection (HAR)<br />

– acute vascular rejection (AVR)<br />

– cellular rejection, and eventually<br />

– chronic rejection (101).


288<br />

Hyperacute rejection occurs within seconds or minutes,<br />

when, in the case of a discordant organ (e.g. in<br />

transplanting from pig to human), pre-existing antibodies<br />

react with antigenic structures on the surface of the porcine<br />

cells and activate the complement cascade; in other words,<br />

the antigen-antibody complex triggers formation of the<br />

membrane attack complex. Induced xenoreactive<br />

antibodies are thought to be responsible for AVR, which<br />

occurs within days after transplantation of a xenograft;<br />

disseminated intravascular coagulation (DIC) is a<br />

predominant feature of AVR (17, 52). Human<br />

thrombomodulin and heme-oxygenase 1 are crucially<br />

involved in the etiology of DIC and are targets for future<br />

transgenic modifications to improve the long-term survival<br />

of porcine xenografts by creating multi-transgenic pigs.<br />

When using a discordant donor species such as the pig,<br />

overcoming HAR and AVR are the pre-eminent goals.<br />

Two main strategies have been successfully explored for<br />

long-term suppression of HAR: synthesis of human<br />

regulators of complement activity (RCAs) in transgenic<br />

pigs (18, 77) and the knockout of α-gal epitopes, the<br />

antigenic structures on the surface of the porcine cells that<br />

cause HAR (52, 74, 105). Survival rates, after the<br />

transplantation of porcine hearts or kidneys expressing<br />

transgenic RCA proteins to immunosuppressed nonhuman<br />

primates, reached 23 to 135 days, indicating that<br />

HAR can be overcome in a clinically acceptable manner<br />

(4). The successful xenotransplantation of porcine organs<br />

with a knockout of the 1,3-α-galactosyltransferase gene,<br />

eliminating the 1,3-α-gal-epitopes produced by the<br />

1,3-α-galactosyltransferase enzyme, has recently been<br />

demonstrated. Upon transplantation of porcine hearts or<br />

kidneys from these α-gal-knockout pigs to baboons,<br />

survival rates reached two to six months (52, 105). A<br />

particularly promising strategy to enhance long-term graft<br />

tolerance is the induction of permanent chimerism via<br />

intraportal injection of embryonic stem (ES) cells (28) or<br />

the co-transplantation of vascularised thymic tissue (105).<br />

Recent findings have revealed that the risk of porcine<br />

endogenous retrovirus transmission is negligible and show<br />

that this critical danger could be eliminated, paving the<br />

way for preclinical testing of xenografts (47). Despite<br />

further challenges, appropriate lines of transgenic pigs are<br />

likely to be available as organ donors within the next five<br />

to ten years. Guidelines for the clinical application of<br />

porcine xenotransplants are already available in the USA<br />

and are currently <strong>bei</strong>ng developed in other countries. The<br />

general consensus of a worldwide debate is that the<br />

technology is ethically acceptable provided that the<br />

individual’s well-<strong>bei</strong>ng does not compromise public health.<br />

Economically, xenotransplantation will be viable, as the<br />

enormous costs of maintaining patients suffering from<br />

severe kidney disease using dialysis or supporting those<br />

suffering from chronic heart disease could be reduced by a<br />

functional kidney or heart xenograft.<br />

Rev. sci. tech. Off. int. Epiz., 24 (1)<br />

Farm animals as models for human diseases<br />

Mouse physiology, anatomy and life span differ<br />

significantly from those of humans, making the rodent<br />

model inappropriate for many human diseases.<br />

Farm animals, such as pigs, sheep or even<br />

cattle, may be more appropriate models in which to<br />

study potential therapies for human diseases that<br />

require longer observation periods than those possible in<br />

mice, e.g. atherosclerosis, non-insulin-dependent diabetes,<br />

cystic fibrosis, cancer and neuro-degenerative disorders<br />

(34, 56, 70, 92). Cardiovascular disease is an increasing<br />

health problem in aging Western societies, where coronary<br />

artery diseases account for the majority of deaths. Because<br />

genetically modified mice do not manifest myocardial<br />

infarction or stroke as a result of atherosclerosis, new<br />

animal models, such as swine that exhibit these<br />

pathologies, are needed to develop effective therapeutic<br />

strategies (32, 80). An important porcine model has been<br />

developed for the rare human eye disease retinitis<br />

pigmentosa (PR) (73). Patients with PR suffer from night<br />

blindness early in life, a condition attributed to a loss of<br />

photoreceptors. Transgenic pigs that express a mutated<br />

rhodopsin gene show great similarity to the<br />

human phenotype and effective treatments are <strong>bei</strong>ng<br />

developed (61).<br />

The pig could be a useful model for studying<br />

defects of growth-hormone releasing hormone<br />

(GHRH), which are implicated in a variety of conditions<br />

such as Turner syndrome, hypochrondroplasia,<br />

Crohn’s disease, intrauterine growth retardation or renal<br />

insufficiency. Application of recombinant GHRH<br />

and its myogenic expression has been shown to alleviate<br />

these problems in a porcine model (26). Development<br />

of further ovine and porcine models of human diseases is<br />

underway (29).<br />

An important aspect of nuclear-transfer-derived<br />

large animal models for human diseases and the<br />

development of regenerative therapies is that somatic<br />

cloning per se does not result in shortening of the<br />

telomeres and thus does not necessarily lead to premature<br />

ageing (85). Telomeres are highly repetitive<br />

DNA sequences at the ends of the chromosomes that are<br />

crucial for their structural integrity and function and are<br />

thought to be related to lifespan. Telomere shortening is<br />

usually correlated with severe limitation of the regenerative<br />

capacity of cells, the onset of cancer, ageing and<br />

chronic disease with significant impacts on human<br />

lifespans (85). Expression of the enzyme telomerase, which<br />

is primarily responsible for the formation and rebuilding of<br />

telomeres, is suppressed in most somatic tissues postnatally.<br />

Recent studies have revealed that telomere length is<br />

established early in pre-implantation development by a<br />

specific genetic programme and correlates with telomerase<br />

activity (84).


Rev. sci. tech. Off. int. Epiz., 24 (1) 289<br />

Transgenic animals in<br />

agriculture<br />

Carcass composition<br />

Transgenic pigs bearing a human metallothionein<br />

promoter/porcine growth-hormone gene construct showed<br />

significant improvements in economically important traits<br />

such as growth rate, feed conversion and body fat/muscle<br />

ratio without the pathological phenotype known from<br />

previous growth hormone constructs (69, 78). Similarly,<br />

pigs transgenic for the human insulin-like growth factor-I<br />

had ~30% larger loin mass, ~10% more carcass lean tissue<br />

and ~20% less total carcass fat (79). The commercialisation<br />

of these pigs has been postponed due to the current lack of<br />

public acceptance of genetically modified foods.<br />

Recently, an important step towards the production of<br />

more healthful pork has been made by the creation of the<br />

first pigs transgenic for a spinach desaturase gene that<br />

produces increased amounts of non-saturated fatty acids.<br />

These pigs have a higher ratio of unsaturated to saturated<br />

fatty acids in striated muscle, which means more healthful<br />

meat since a diet rich in non-saturated fatty acids is known<br />

to be correlated with a reduced risk of stroke and coronary<br />

diseases (66, 83).<br />

Lactation<br />

The physicochemical properties of milk are mainly due to<br />

the ratio of casein variants, making these a prime target for<br />

the improvement of milk composition. Transgenic mice<br />

have been developed with various modifications<br />

demonstrating the feasibility of obtaining significant<br />

alterations in milk composition in larger animals, but at the<br />

same time, showing that unwanted side effects can occur<br />

(50, 89).<br />

Dairy production is an attractive field for targeted genetic<br />

modification (44, 106). It may be possible to produce milk<br />

with a modified lipid composition by modulating the<br />

enzymes involved in lipid metabolism, or to increase curd<br />

and cheese production by enhancing expression of the<br />

casein gene family in the mammary gland. The bovine<br />

casein ratio has been altered by over-expression of betaand<br />

kappa-casein, clearly underpinning the potential for<br />

improvements in the functional properties of bovine milk<br />

(8). Furthermore, it may be possible to create<br />

‘hypoallergenic’ milk by knockout or knockdown of the<br />

β-lactoglobulin gene; to generate lactose-free milk by a<br />

knockout or knockdown of the α-lactalbumin locus,<br />

which is the key molecule in milk sugar synthesis; to<br />

produce ‘infant milk’ in which human lactoferrin is<br />

abundantly available; or to produce milk with a highly<br />

improved hygienic standard by increasing the amount of<br />

lysozyme. Lactose-reduced or lactose-free milk would<br />

render dairy products suitable for consumption by the<br />

large numbers of adult humans who do not possess an<br />

active intestinal lactase enzyme system. Although lactose is<br />

the main osmotically active substance in milk and a lack<br />

thereof could interfere with milk secretion, a lactase<br />

construct has been tested in the mammary gland of<br />

transgenic mice. In hemizygous mice, this reduced lactose<br />

contents by 50% to 85% without altering milk secretion<br />

(43). Unfortunately, mice with a homozygous knockout for<br />

α-lactalbumin could not nurse their offspring because of<br />

the high viscosity of their milk (89). These findings<br />

demonstrate the feasibility of obtaining significant<br />

alterations in milk composition by applying the<br />

appropriate strategy.<br />

In the pig, transgenic expression of a bovine lactalbumin<br />

construct in sow milk has been shown to result in higher<br />

lactose contents and greater milk yields, which correlated<br />

with improved survival and development of piglets (100).<br />

Any increased survival of piglets at weaning would provide<br />

significant commercial advantages to the producer.<br />

Wool production<br />

Transgenic sheep carrying a keratin-IGF-I construct<br />

showed that expression in the skin and the amount of clear<br />

fleece was about 6.2% greater in transgenic than in nontransgenic<br />

animals. No adverse effects on health or<br />

reproduction were observed (22, 23). Approaches<br />

designed to alter wool production by transgenic<br />

modification of the cystein pathway have met with only<br />

limited success, although cystein is known to be the ratelimiting<br />

biochemical factor for wool growth (96).<br />

Environmentally friendly farm animals<br />

Phytase transgenic pigs have been developed to address the<br />

problem of manure-related environmental pollution. These<br />

pigs carry a bacterial phytase gene under the<br />

transcriptional control of a salivary-gland-specific<br />

promoter, which allows the pigs to digest plant phytate.<br />

Without the bacterial enzyme, phytate phosphorus passes<br />

undigested into manure and pollutes the environment.<br />

With the bacterial enzyme, fecal phosphorus output was<br />

reduced by up to 75% (30). Developers expect these<br />

environmentally friendly pigs to enter commercial<br />

production in Canada within the next few years.<br />

Transgenic animals and disease resistance<br />

Transgenic strategies to increase disease resistance<br />

In most cases, susceptibility to pathogens originates from<br />

the interplay of numerous genes; in other words,<br />

susceptibility to pathogens is polygenic in nature. Only a


290<br />

few loci are known that confer resistance against a specific<br />

disease. Transgenic strategies to enhance disease resistance<br />

include the transfer of major histocompatibility-complex<br />

genes, T-cell-receptor genes, immunoglobulin genes, genes<br />

that affect lymphokines or specific disease-resistance genes<br />

(64). A prominent example for a specific disease resistance<br />

gene is the murine Mx-gene. Production of the<br />

Mx1-protein is induced by interferon and was discovered<br />

in inbred mouse strains that are resistant to infection with<br />

influenza viruses (88). Microinjection of an interferon- and<br />

virus-inducible Mx-construct into porcine zygotes resulted<br />

in two transgenic pig lines that expressed the<br />

Mx-messenger RNA (mRNA); unfortunately, no Mx protein<br />

could be detected (65). Recently the bovine MxI gene was<br />

identified and shown to confer antiviral activity as a<br />

transgenic construct in Vero cells (6).<br />

Transgenic constructs bearing the immunoglobulin-A<br />

(IgA) gene have been successfully introduced into pigs,<br />

sheep and mice in an attempt to increase resistance against<br />

infections (57). The murine IgA gene was expressed in two<br />

transgenic pig lines, but only the light chains were detected<br />

and the IgA-molecules showed only marginal binding to<br />

phosphorylcholine (57). High levels of monoclonal murine<br />

antibodies with a high binding affinity for their specific<br />

antigen have been produced in transgenic pigs (97).<br />

Attempts to increase ovine resistance to Visna virus<br />

infection via transgenic production of Visna envelope<br />

protein have been reported (15). The transgenic sheep<br />

developed normally and expressed the viral gene without<br />

pathological side effects. However, the transgene was not<br />

expressed in monocytes, the target cells of the viral<br />

infection. Antibodies were detected after an artificial<br />

infection of the transgenic animals (15).<br />

It has also proved possible to induce passive immunity<br />

against an economically important porcine disease in a<br />

transgenic mouse model (10). The transgenic mice secreted<br />

a recombinant antibody in milk that neutralised the corona<br />

virus responsible for transmissible gastroenteritis (TGEV)<br />

and conferred resistance against TGEV. Strong mammarygland-specific<br />

expression was achieved for the entire<br />

duration of lactation. Verification of this work in transgenic<br />

pigs is anticipated in the near future.<br />

Knockout of the prion protein is the only secure way to<br />

prevent infection and transmission of spongiform<br />

encephalopathies like scrapie or bovine spongiform<br />

encephalopathy (98). The first successful targeting of the<br />

ovine prion locus has been reported; however, the cloned<br />

lambs carrying the knockout locus died shortly after birth<br />

(24). Cloned cattle with a knockout for the prion locus<br />

have also been generated (19). Transgenic animals with<br />

modified prion genes will be an appropriate model for<br />

studying the epidemiology of spongiform<br />

encephalopathies in humans and are crucial for developing<br />

Rev. sci. tech. Off. int. Epiz., 24 (1)<br />

strategies to eliminate prion carriers from farm animal<br />

populations.<br />

Transgenic approaches to increase disease<br />

resistance of the mammary gland<br />

The levels of the anti-microbial peptides lysozyme and<br />

lactoferrin in human milk is many times higher than in<br />

bovine milk. Transgenic expression of the human lysozyme<br />

gene in mice was associated with a significant reduction of<br />

bacteria and reduced the frequency of mammary gland<br />

infections (59, 60). Lactoferrin has bactericidal and<br />

bacteriostatic effects, in addition to <strong>bei</strong>ng the main iron<br />

source in milk. These properties make an increase in<br />

lactoferrin levels in bovine transgenics a practical way to<br />

improve milk quality. Human lactoferrin has been expressed<br />

in the milk of transgenic mice and cattle at high levels (46,<br />

76) and is associated with an increased resistance against<br />

mammary gland diseases (93). Lycostaphin has been shown<br />

to confer specific resistance against mastitis caused by<br />

Staphylococcus aureus. A recent report indicates that<br />

transgenic technology has been used to produce cows that<br />

express a lycostaphin gene construct in the mammary gland,<br />

thus making them mastitis-resistant (95).<br />

Emerging transgenic<br />

technologies<br />

Lentiviral transfection of oocytes and zygotes<br />

Recent research has shown that lentiviruses can overcome<br />

previous limitations of viral-mediated gene transfer, which<br />

included the silencing of the transgenic locus and low<br />

expression levels (104). Injection of lentiviruses into the<br />

perivitelline space of porcine zygotes resulted in a very<br />

high proportion of piglets that carried and expressed the<br />

transgene. Stable transgenic lines have been established by<br />

this method (36). The generation of transgenic cattle by<br />

lentiviruses requires microinjection into the perivitelline<br />

space of oocytes and has a lower efficiency than that<br />

obtained in pigs (37). Lentiviral gene transfer in livestock<br />

promises unprecedented efficiency of transgenic animal<br />

production. Whether the multiple integration of<br />

lentiviruses into the genome is associated with unwanted<br />

side-effects like oncogene activation or insertional<br />

mutagenesis remains to be investigated.<br />

Chimera generation via injection of pluripotent<br />

cells into blastocysts<br />

Embryonic stem cells with pluripotent characteristics have<br />

the ability to participate in organ and germ cell<br />

development after injection into blastocysts or by<br />

aggregation with morulae (81). True ES cells (that is, those


Rev. sci. tech. Off. int. Epiz., 24 (1) 291<br />

able to contribute to the germ line) are currently only<br />

available from inbred mouse strains (48). In mouse<br />

genetics, ES cells have become an important tool for<br />

generating gene knockouts, gene knockins and large<br />

chromosomal rearrangements (25). Embryonic stem-like<br />

cells and primordial germ cell cultures have been reported<br />

for several farm animal species, and chimeric animals<br />

without germ line contribution have been reported in<br />

swine (1, 87, 99) and cattle (13). Recent data indicate that<br />

somatic stem cells may have a much greater potency than<br />

previously assumed (42, 49). Whether these cells will<br />

improve the efficiency of chimera generation or somatic<br />

nuclear transfer in farm animals has yet to be shown (48).<br />

Culture of spermatogonia and transplantation<br />

into recipient males<br />

Transplantation of genetically altered primordial germ cells<br />

into the testes of host male animals is an alternative<br />

approach to generating transgenic animals. Initial<br />

experiments in mice showed that the depletion of<br />

endogenous spermatogenesis by treatment with busulfan<br />

prior to transplantation is effective and compatible with recolonisation<br />

by donor cells. Recently, researchers<br />

succeeded in transmitting the donor haplotype to the next<br />

generation after germ-cell transplantation (38). The major<br />

obstacles to this strategy are the lack of efficient in vitro<br />

culture methods for primordial germ/prospermatogonial<br />

cells, and the limitations this imposes on gene transfer<br />

techniques for these cells.<br />

Ribonucleic acid interference<br />

Ribonucleic acid interference is a conserved posttranscriptional<br />

gene regulatory process in most biological<br />

systems, including fungi, plants and animals. Common<br />

mechanistic elements include siRNAs with 21 to<br />

23 nucleotides, which specifically bind complementary<br />

sequences on their target mRNAs and shut down<br />

expression. The target mRNAs are degraded by<br />

exonucleases and no protein is translated (75). The RNA<br />

interference seems to be involved in gene regulation by<br />

controlling/suppressing the translation of mRNAs from<br />

endogenous viral elements.<br />

The relative simplicity of active siRNAs has facilitated the<br />

adoption of this mechanism to generate transient or<br />

permanent knockdowns for specific genes. For transient<br />

gene knockdown, synthetic siRNAs can be transfected into<br />

cells or embryonic stages. For stable gene repression, the<br />

siRNA sequences must be incorporated into a gene<br />

construct (14). The conjunction of siRNA and lentiviral<br />

vector technology may soon provide a method with high<br />

gene transfer efficiency and highly specific gene<br />

knockdown for livestock.<br />

Safety aspects and outlook<br />

Biological products from any animal source are unique,<br />

and must be handled in a different manner from<br />

chemically synthesised drugs to assure their safety, purity<br />

and potency. Proteins derived from living systems are heat<br />

labile, subject to microbial contamination, can be damaged<br />

by shear forces and have the potential to be immunogenic<br />

and allergenic. In the USA, the FDA has developed<br />

guidelines to assure the safe commercial exploitation of<br />

recombinant biological products. A crucial consideration<br />

with animal-derived products is the prevention of<br />

transmission of pathogens from animals to humans (47).<br />

This requires sensitive and reliable diagnostic and<br />

screening methods for the various types of pathogenic<br />

organisms. Furthermore, it should be kept in mind that all<br />

transgenic applications of farm animals will require strict<br />

standards of ‘genetic security’ and reliable, sensitive<br />

methods for molecular characterisation of the products. A<br />

major contribution towards the goal of well-defined<br />

products will come from matrix-assisted laser<br />

desorption/ionisation time-of-flight spectometry (40, 91).<br />

Meanwhile improvements in RNA isolation and unbiased<br />

global amplification of picogram amounts of mRNA enable<br />

researchers to analyse RNA even from single embryos (7).<br />

This technology can offer insights into the entire<br />

transcriptome of a transgenic organism and thereby ensure<br />

the absence of unwanted side-effects (40, 91). Rigorous<br />

control is also required to maintain the highest possible<br />

levels of animal welfare in cases of genetic modification.<br />

Conclusion<br />

Throughout history, animal husbandry has made significant<br />

contributions to human health and well-<strong>bei</strong>ng. The<br />

convergence of recent advances in reproductive technologies<br />

(in vitro production of embryos, sperm sexing, somatic<br />

nuclear transfer) with the tools of molecular biology (gene<br />

targeting and array analysis of gene expression) adds a new<br />

dimension to animal breeding (68). Major prerequisites for<br />

success and safety will be the continuous refinement of<br />

reproductive biotechnologies and a rapid completion of<br />

genomic sequencing projects in livestock. The authors<br />

anticipate that within the next five to seven years genetically<br />

modified animals will play a significant role in the<br />

biomedical arena, in particular via the production of valuable<br />

pharmaceutical proteins and the supply of xenografts<br />

(Table II). Agricultural applications are already <strong>bei</strong>ng<br />

prepared (39), but general public acceptance may take as<br />

long as ten years to achieve. As the complete genomic<br />

sequences of all farm animals become available, it will be<br />

possible to refine targeted genetic modification in animal<br />

breeding and to develop strategies to cope with future<br />

challenges in global agricultural production.


292<br />

Glossary<br />

Artificial chromosome<br />

A construction composed of the basic elements of a normal<br />

chromosome into which one or more transgene sequences<br />

may be introduced. After introduction into the nucleus,<br />

this artificial chromosome is replicated and distributed to<br />

daughter cells just like a normal chromosome. The<br />

transgene expression is more predictable as it is not<br />

influenced by integration into an unknown location.<br />

Expression pattern<br />

A representation of relative levels of gene expression in the<br />

various cells and tissues of a transgenic animal. This is<br />

primarily determined by the promoter, but is also<br />

dependent on the position in the genome where the<br />

construct has become integrated. As cells differentiate,<br />

certain chromosomal regions are permanently turned off.<br />

Homologous recombination<br />

A transgene sequence is constructed which matches the<br />

sequence of the endogenous target gene. This biases<br />

insertion to the target gene. The design of the gene<br />

construct determines whether the target gene will be<br />

rendered inactive (knocked out) or simply modified.<br />

Inducible promoter<br />

A promoter that can be activated by an environmental<br />

signal such as the introduction of a heavy metal or an<br />

antibiotic.<br />

Knockout<br />

Removal of a gene or part of a gene to eliminate a particular<br />

protein product.<br />

Lentivirus<br />

One of several types of virus that have been evaluated for<br />

their ability to carry genes into early-stage embryos for the<br />

purpose of introducing a transgene. The transgene is<br />

placed within the viral genome and is carried into the<br />

target embryo by the viral infection mechanism.<br />

Microinjection<br />

A now routine procedure in which a glass micropipette is<br />

used to deliver a small amount of DNA solution (the<br />

transgene) into the nucleus of a fertilised egg. In a small<br />

percentage of these injected eggs, the transgene becomes<br />

incorporated into one of the chromosomes and so will be<br />

present in all subsequent cell divisions and in the germ<br />

cells, so that future offspring will also carry the transgene.<br />

Mosaicism<br />

The situation where an individual is composed of cells of<br />

more than one genetic background. Transgene integration<br />

takes place after the first round of cell division; thus, only<br />

a portion of the cells in the resulting individual will be<br />

transgenic. It is important that the germ cells carry the<br />

transgene so that subsequent offspring will also be<br />

transgenic.<br />

Multiple transgenic<br />

A transgenic organism carrying more than one transgene.<br />

This can be achieved either by introduction of several<br />

transgenes at the same time, or by obtaining cells or<br />

embryos from an existing transgenic animal and using<br />

these for a second round of transgenesis. This is most<br />

efficiently achieved by nuclear-transfer-based transgenesis<br />

because the transfected cells can be screened prior to the<br />

production of an embryo.<br />

Promoter<br />

The control sequence that determines in which tissue or<br />

cell type and at which point in time a gene will be<br />

expressed. Some promoters are universally active, and<br />

some are active only in specific cell types or developmental<br />

stages.<br />

Reporter gene<br />

A sequence that codes for a detectable product such as a<br />

fluorescent protein or an enzyme which produces a visible<br />

product.<br />

Somatic nuclear transfer<br />

In its most common form, this involves removal of the<br />

nucleus from a fertilised egg and replacement with the<br />

nucleus of a differentiated somatic cell such as a fibroblast.<br />

The result is a cloned embryo that has the genetic<br />

composition of the nuclear donor cell.<br />

Spermatogonia<br />

Rev. sci. tech. Off. int. Epiz., 24 (1)<br />

Early developmental precursors of sperm that can be<br />

removed from the testes, transfected with a transgene, and<br />

then re-implanted to develop into mature sperm that carry<br />

the transgene.


Rev. sci. tech. Off. int. Epiz., 24 (1) 293<br />

Targeted integration<br />

Sequences have been discovered (FLP, Cre/lox) that cause<br />

the insertion or removal of intervening gene sequences.<br />

When properly used, these systems can direct insertion to<br />

a specific location in the genome, or can at some point in<br />

the future remove portions of a transgene that are no<br />

longer needed (such as the antibiotic resistance sequences<br />

used to select transfected cells).<br />

Transfection<br />

A process mediated by an electrical pulse (electroporation),<br />

or by treatment with agents that make the cell membranes<br />

porous, such that DNA molecules in the culture medium<br />

are able to pass into the cell where they can be expressed<br />

or integrated into the genome.<br />

Animaux d’élevage transgéniques : le présent et l’avenir<br />

H. Niemann, W. Kues & J.W. Carnwath<br />

Résumé<br />

Jusqu’à une date récente, la microinjection pronucléaire d’acide<br />

désoxyribonucléique (ADN) était la méthode standard utilisée pour la production<br />

d’animaux transgéniques. Cette technique est actuellement remplacée par des<br />

protocoles plus efficaces basés sur le transfert nucléaire de cellules somatiques<br />

permettant également de réaliser des modifications génétiques ciblées. De<br />

même, les vecteurs lentiviraux et la technique de l’acide ribonucléique (ARN)<br />

interférent deviennent des outils importants de la transgenèse. Les animaux<br />

d’élevage transgéniques ont une importance en médecine humaine comme<br />

sources de protéines biologiquement actives, comme donneurs pour la<br />

xénotransplantation et à des fins de recherche en thérapie cellulaire et génique.<br />

Parmi les applications agricoles habituelles figurent l’amélioration de la<br />

composition des viandes, de la production laitière et lainière, de la résistance<br />

aux maladies, de même que la réduction de l’impact sur l’environnement. La<br />

sécurité sanitaire des produits peut être assurée par la standardisation des<br />

procédures et contrôlée par l’application de la réaction en chaîne par<br />

polymérase et de la technologie du criblage de filtres à ADN. Avec l’affinement<br />

des données relatives aux séquences et des cartes génomiques des animaux<br />

d’élevage, la suppression ou la modification de certains gènes devient de plus en<br />

plus réalisable. Cette méthode de reproduction d’animaux jouera un rôle<br />

déterminant dans la résolution des problèmes mondiaux qui se poseront demain<br />

en matière de production agricole.<br />

Mots-clés<br />

Animaux d’élevage – Application agricole – Avantage environnemental – « Gene<br />

pharming » – Microinjection – Séquence courte à interférence ARN – Transfert nucléaire<br />

– Transgénique – Vecteur lentiviral – Xénotransplantation.


294<br />

Presente y futuro del ganado transgénico<br />

References<br />

1. Anderson G.B. (1999). – Embryonic stem cells in agricultural<br />

species. In Transgenic animals in agriculture (J.D. Murray,<br />

G.B. Anderson, A.M. Oberbauer & M.M. McGloughlin, eds).<br />

CABI Publishing, New York, 57-66.<br />

2. Anon. (2004). – Chicken genome. Genome Sequencing<br />

Center at Washington University in St Louis. Website:<br />

genome.wustl.edu/projects/chicken/ (accessed on 17 May<br />

2005).<br />

3. Anon. (2005). – Bovine Genome Project. Human Genome<br />

Sequencing Center at Baylor College of Medicine. Website:<br />

www.hgsc.bcm.tmc.edu/projects/bovine/ (accessed on<br />

17 May 2005).<br />

H. Niemann, W. Kues & J.W. Carnwath<br />

Rev. sci. tech. Off. int. Epiz., 24 (1)<br />

Resumen<br />

Hasta hace poco tiempo, el procedimiento de rigor para obtener animales<br />

transgénicos era la microinyección pronuclear de ácido desoxirribonucleico<br />

(ADN). Ahora se empiezan a aplicar otros métodos más eficaces basados en la<br />

transferencia de núcleos de células somáticas, que además permiten inducir<br />

cambios genéticos de manera más específica. El uso de vectores lentivíricos y<br />

la técnica del ARN interferente pequeño son otros dos métodos de transgénesis<br />

cada vez más empleados. El ganado transgénico es importante en medicina<br />

humana porque de él se extraen proteínas biológicamente activas y tejidos u<br />

órganos para xenotrasplantes y porque proporciona material para investigar<br />

terapias celulares y génicas. En el terreno de la producción animal, sus<br />

aplicaciones van desde la mejora de la composición de carne, el rendimiento<br />

lechero y la producción de lana hasta el aumento de la resistencia a las<br />

enfermedades o la atenuación de los efectos ambientales de la ganadería. La<br />

estandarización de protocolos puede servir para garantizar la inocuidad de los<br />

productos obtenidos, y técnicas como la reacción en cadena de la polimerasa<br />

(PCR) o la de matrices (arrays) de ADN para efectuar los correspondientes<br />

controles. A medida que se perfeccionan los mapas genómicos del ganado y se<br />

conoce mejor su secuencia génica va resultando cada vez más fácil eliminar o<br />

modificar genes concretos. En el futuro, estos sistemas de selección animal<br />

serán básicos para resolver los problemas de producción agropecuaria en el<br />

plano mundial.<br />

Palabras clave<br />

Aplicación agropecuaria – ARN interferente pequeño – Beneficio ambiental – Ganado –<br />

“Gen pharming” – Microinyección – Transferencia nuclear – Transgénico – Vector<br />

lentivírico – Xenotrasplante.<br />

4. Bach F.H. (1998). – Xenotransplantation: problems and<br />

prospects. Annu. Rev. Med., 49, 301-310.<br />

5. Baguisi A., Behboodi E., Melican D., Pollock J.S.,<br />

Destrempes M.M., Cammuso C., Williams J.L., Nims S.D.,<br />

Porter C.A., Midura P., Palacios M.J., Ayres S.L.,<br />

Denniston R.S., Hayes M.L., Ziomek C.A., Meade H.M.,<br />

Godke R.A., Gavin W.G., Overström E.W. & Echelard Y.<br />

(1999). – Production of goats by somatic cell nuclear transfer.<br />

Nature Biotechnol., 17 (5), 456-461.<br />

6. Baise E., Pire G., Leroy M., Gerardin J., Goris N.,<br />

De Clercq K., Kerkhofs P. & Desmecht D. (2004). –<br />

Conditional expression of type I interferon-induced bovine<br />

Mx1 GTPase in a stable transgenic vero cell line interferes<br />

with replication of vesicular stomatitis virus. J. Interferon<br />

Cytokine Res., 24 (9), 512-521.


Rev. sci. tech. Off. int. Epiz., 24 (1) 295<br />

7. Brambrink T., Wabnitz P., Halter R., Klocke R.,<br />

Carnwath J.W., Kues W.A., Wrenzycki C., Paul D. &<br />

Niemann H. (2002). – Application of cDNA arrays to monitor<br />

mRNA profiles in single preimplantation mouse embryos.<br />

Biotechniques, 33 (2), 376-385.<br />

8. Brophy B., Smolenski G., Wheeler T., Wells D., L’Huillier P. &<br />

Laible G. (2003). – Cloned transgenic cattle produce milk<br />

with higher levels of β-casein and κ-casein. Nature Biotechnol.,<br />

21 (2), 157-162.<br />

9. Capecchi M.R. (1989). – The new mouse genetics: altering<br />

the genome by gene targeting. Trends Genet., 5 (3), 70-76.<br />

10. Castilla J., Pintado B., Sola I., Sánchez-Morgado J.M. &<br />

Enjuanes L. (1998). – Engineering passive immunity in<br />

transgenic mice secreting virus-neutralizing antibodies in<br />

milk. Nature Biotechnol., 16 (4), 349-354.<br />

11. Chan A.W., Homan E.J., Ballou L.U., Burns J.C. &<br />

Bremel R.D. (1998). – Transgenic cattle produced by reversetranscribed<br />

gene transfer in oocytes. Proc. natl Acad. Sci. USA,<br />

95 (24), 14028-14033.<br />

12. Chang K., Qian J., Jiang M., Liu Y.H., Wu M.C., Chen C.D.,<br />

Lai C.K., Lo H.L., Hsiao C.T., Brown L., Bolen J. Jr,<br />

Huang H.I., Ho P.Y., Shih P.Y., Yao C.W., Lin W.J., Chen C.H.,<br />

Wu F.Y., Lin Y.J., Xu J. & Wang K. (2002). – Effective<br />

generation of transgenic pigs and mice by linker based<br />

sperm-mediated gene transfer. BMC Biotechnol., 2 (1), 5.<br />

13. Cibelli J.B., Stice S.L., Golueke P.L., Kane J.J., Jerry J.,<br />

Blackwell C., Ponce de Leon F.A. & Robl J.M. (1998). –<br />

Transgenic bovine chimeric offspring produced from somatic<br />

cell-derived stem like cells. Nature Biotechnol., 16 (7), 642-<br />

646.<br />

14. Clark J. & Whitelaw B. (2003). – A future for transgenic<br />

livestock. Nat. Rev. Genet., 4 (10), 825-833.<br />

15. Clements J.E., Wall R.J., Narayan O., Hauer D., Schoborg R.,<br />

Sheffer D., Powell A., Carruth L.M., Zink M.C. &<br />

Rexroad C.E. (1994). – Development of transgenic sheep that<br />

express the Visna virus envelope gene. Virology, 200 (2),<br />

370-380.<br />

16. Copley M.S., Berstan R., Dudd S.N., Docherty G.,<br />

Mukherjee A.J., Straker V., Payne S. & Evershed R.P. (2003).<br />

– Direct chemical evidence for widespread dairying in<br />

prehistoric Britain. Proc. natl Acad. Sci. USA, 100 (4),<br />

1524-1529.<br />

17. Cowan P.J., Aminian A., Barlow H., Brown A.A., Chen C.G.,<br />

Fisicaro N., Francis D.M., Goodman D.J., Han W., Kurek M.,<br />

Nottle M.B., Pearse M.J., Salvaris E., Shinkel T.A.,<br />

Stainsby G.V., Stewart A.B. & d’Apice A.J. (2000). – Renal<br />

xenografts from triple-transgenic pigs are not hyperacutely<br />

rejected but cause coagulopathy in nonimmunosuppressed<br />

baboons. Transplantation, 69 (12), 2504-2515.<br />

18. Cozzi E. & White D.J.G. (1995). – The generation of<br />

transgenic pigs as potential organ donors for humans. Nature<br />

Med., 1 (9), 964-966.<br />

19. Cyranoski D. (2003). – Koreans rustle up madness-resistant<br />

cows. Nature, 426 (6968), 743.<br />

20. D’Agnillo F. & Chang T.M. (1998). – Polyhemoglobinsuperoxide<br />

dismutase-catalase as a blood substitute with<br />

antioxidant properties. Nature Biotechnol., 16 (7), 667-671.<br />

21. Dai Y., Vaught T.D., Boone J., Chen S.H., Phelps C.J., Ball S.,<br />

Monahan J.A., Jobst P.M., McCreath K.J., Lamborn A.E.,<br />

Cowell-Lucero J.L., Wells K.D., Colman A., Polejaeva I.A. &<br />

Ayares D.L. (2002). – Targeted disruption of the α1,3galactosyltransferase<br />

gene in cloned pigs. Nature Biotechnol.,<br />

20 (3), 251-255.<br />

22. Damak S., Jay N.P., Barrell G.K. & Bullock D.W. (1996). –<br />

Targeting gene expression to the wool follicle in transgenic<br />

sheep. Biotechnology, 14 (2), 181-184.<br />

23. Damak S., Su H., Jay N.P. & Bullock D.W. (1996). –<br />

Improved wool production in transgenic sheep expressing<br />

insulin-like growth factor 1. Biotechnology, 14 (2), 185-188.<br />

24. Denning C., Burl S., Ainslie A., Bracken J., Dinnyes A.,<br />

Fletcher J., King T., Ritchie M., Ritchie W.A., Rollo M.,<br />

de Sousa P., Travers A., Wilmut I. & Clark A.J. (2001). –<br />

Deletion of the alpha(1,3)galactosyltransferase (GGTA1) and<br />

the prion protein (PrP) gene in sheep. Nature Biotechnol., 19<br />

(6), 559-562.<br />

25. Downing G.J. & Battey J.F. Jr (2004). – Technical assessment<br />

of the first 20 years of research using mouse embryonic stem<br />

cell lines. Stem Cells, 22 (7), 1168-1180.<br />

26. Draghia-Akli R., Fiorotto M.L., Hill L.A., Malone P.B.,<br />

Deaver D.R. & Schwartz R.J. (1999). – Myogenic expression<br />

of an injectable protease-resistant growth hormone-releasing<br />

hormone augments long-term growth in pigs. Nature<br />

Biotechnol., 17 (12), 1179-1183.<br />

27. Dyck M.K., Lacroix D., Pothier F. & Sirard M.A. (2003). –<br />

Making recombinant proteins in animals: different systems,<br />

different applications. Trends Biotechnol., 21 (9), 394-399.<br />

28. Fändrich F., Lin X., Chai G.X., Schulze M., Ganten D.,<br />

Bader M., Holle J., Huang D.S., Parwaresch R., Zavazava N.<br />

& Binas B. (2002). – Preimplantation-stage stem cells induce<br />

long-term allogeneic graft acceptance without supplementary<br />

host conditioning. Nature Med., 8 (2), 171-178.<br />

29. Forsberg E.J. (2005). – Commercial applications of nuclear<br />

transfer cloning: three examples. Reprod. Fertil. Dev., 17 (2),<br />

59-68.<br />

30. Golovan S.P., Meidinger R.G., Ajakaiye A., Cottrill M.,<br />

Wiederkehr M.Z., Barney D.J., Plante C., Pollard J.W.,<br />

Fan M.Z., Hayes M.A., Laursen J., Hjorth J.P., Hacker R.R.,<br />

Phillips J.P. & Forsberg C.W. (2001). – Pigs expressing<br />

salivary phytase produce low-phosphorus manure. Nature<br />

Biotechnol., 19 (8), 741-745.<br />

31. Grosse-Hovest L., Muller S., Minoia R., Wolf E.,<br />

Zakhartchenko V., Wenigerkind H., Lassnig C.,<br />

Besenfelder U., Müller M., Lytton S.D., Jung G. & Brem G.<br />

(2004). – Cloned transgenic farm animals produce a<br />

bispecific antibody for T cell-mediated tumor cell killing.<br />

Proc. natl Acad. Sci. USA, 101 (18), 6858-6863.


296<br />

32. Grunwald K.A., Schueler K., Uelmen P.J., Lipton B.A.,<br />

Kaiser M., Buhman K. & Attie A.D. (1999). – Identification of<br />

a novel Arg-->Cys mutation in the LDL receptor that<br />

contributes to spontaneous hypercholesterolemia in pigs.<br />

J. Lipid. Res., 40 (3), 475-485.<br />

33. Hammer R.E., Pursel V.G., Rexroad C.E. Jr, Wall R.J.,<br />

Bolt D.J., Ebert K.M., Palmiter R.D. & Brinster R.L. (1985). –<br />

Production of transgenic rabbits, sheep and pigs by<br />

microinjection. Nature, 315 (6021), 680-683.<br />

34. Hansen K. & Khanna C. (2004). – Spontaneous and<br />

genetically engineered animal models: use in preclinical<br />

cancer drug development. Eur. J. Cancer, 40 (6), 858-880.<br />

35. Haskell R.E. & Bowen R.A. (1995). – Efficient production of<br />

transgenic cattle by retroviral infection of early embryos.<br />

Molec. Reprod. Dev., 40 (3), 386-390.<br />

36. Hofmann A., Kessler B., Ewerling S., Weppert M., Vogg B.,<br />

Ludwig H., Stojkovic M., Boelhauve M., Brem G., Wolf E. &<br />

Pfeifer A. (2003). – Efficient transgenesis in farm animals by<br />

lentiviral vectors. EMBO Rep., 4 (11), 1054-1060.<br />

37. Hofmann A., Zakhartchenko V., Weppert M., Sebald H.,<br />

Wenigerkind H., Brem G., Wolf E. & Pfeifer A. (2004). –<br />

Generation of transgenic cattle by lentiviral gene transfer into<br />

oocytes. Biol. Reprod., 71 (2), 405-409.<br />

38. Honaramooz A., Behboodi E., Megee S.O., Overton S.A.,<br />

Galantino-Homer H., Echelard Y. & Dobrinski I. (2003). –<br />

Fertility and germline transmission of donor haplotype<br />

following germ cell transplantation in immunocompetent<br />

goats. Biol. Reprod., 69 (4), 1260-1264.<br />

39. Houdebine L.M. (2002). – Transgenesis to improve animal<br />

production. Livest. Prod. Sci., 74 (3), 255-268.<br />

40. Hughes T.R., Roberts C.J., Dai H., Jones A.R., Meyer M.R.,<br />

Slade D., Burchard J., Dow S., Ward T.R., Kidd M.J.,<br />

Friend S.H. & Marton M.J. (2000). – Widespread aneuploidy<br />

revealed by DNA microarray expression profiling.<br />

Nature Genet., 25 (3), 333-337.<br />

41. Hyttinen J.M., Peura T., Tolvanen M., Aalto J., Alhonen L.,<br />

Sinervirta R., Halmekyto M., Myohanen S. & Janne J. (1994).<br />

– Generation of transgenic dairy cattle from transgeneanalyzed<br />

and sexed embryos produced in vitro. Biotechnology,<br />

12 (6), 606-608.<br />

42. Jiang Y., Jahagirdar B.N., Reinhardt R.L., Schwartz R.E.,<br />

Keene C.D., Ortiz-Gonzalez X.R., Reyes M., Lenvik T.,<br />

Lund T., Blackstad M., Du J., Aldrich S., Lisberg A.,<br />

Low W.C., Largaespada D.A. & Verfaillie C.M. (2002). –<br />

Pluripotency of mesenchymal stem cells derived from adult<br />

marrow. Nature, 418 (6893), 41-49.<br />

43. Jost B., Vilotte J.L., Duluc I., Rodeau J.L. & Freund J.N.<br />

(1999). – Production of low-lactose milk by ectopic<br />

expression of intestinal lactase in the mouse mammary gland.<br />

Nature Biotechnol., 17 (2), 160-164.<br />

44. Karatzas C.N. & Turner J.D. (1997). – Toward altering milk<br />

composition by genetic manipulation: current status and<br />

challenges. J. Dairy Sci., 80 (9), 2225-2232.<br />

Rev. sci. tech. Off. int. Epiz., 24 (1)<br />

45. Kilby N.J., Snaith M.R. & Murray J.A.H. (1993). – Sitespecific<br />

recombinases: tools for genome engineering. Trends<br />

Genet., 9 (12), 413-421.<br />

46. Krimpenfort P., Rademakers A., Eyestone W., van der Schans A.,<br />

van den Broek S., Kooiman P., Kootwijk E., Platenburg G.,<br />

Pieper F., Strijker R. & de Boer H. (1991). – Generation of<br />

transgenic dairy cattle using in vitro embryo production.<br />

Biotechnology, 9 (9), 844-847.<br />

47. Kues W.A. & Niemann H. (2004). – The contribution of farm<br />

animals to human health. Trends Biotechnol., 22 (6), 286-294.<br />

48. Kues W.A., Carnwath J.W. & Niemann H. (2005). – From<br />

fibroblasts and stem cells: implications for cell therapies and<br />

somatic cloning. Reprod. Fertil. Dev., 17 (2), 125-134.<br />

49. Kues W.A., Petersen B., Mysegades W., Carnwath J.W. &<br />

Niemann H. (2005). – Isolation of murine and porcine fetal<br />

stem cells from somatic tissue. Biol. Reprod., 72 (4), 1020-<br />

1028. Website: www.biolreprod.com/cgi/content/abstract/<br />

biolreprod.104.031229v1 (accessed on 23 March 2005).<br />

50. Kumar S., Clarke A.R., Hooper M.L., Horne D.S., Law A.J.,<br />

Leaver J., Springbett A., Stevenson E. & Simons J.P. (1994). –<br />

Milk composition and lactation of β-casein deficient mice.<br />

Proc. natl Acad. Sci. USA, 91 (13), 6138-6142.<br />

51. Kuroiwa Y., Kasinathan P., Choi Y.J., Naeem R., Tomizuka K.,<br />

Sullivan E.J., Knott J.G., Duteau A., Goldsby R.A.,<br />

Osborne B.A., Ishida I. & Robl J.M. (2002). –<br />

Cloned transchromosomic calves producing human<br />

immunoglobulin. Nature Biotechnol., 20 (9), 889-894.<br />

52. Kuwaki K., Tseng Y.L., Dor F.J., Shimizu A., Houser S.L.,<br />

Sanderson T.M., Lancos C.J., Prabharasuth D.D., Cheng J.,<br />

Moran K., Hisashi Y., Mueller N., Yamada K., Greenstein J.L.,<br />

Hawley R.J., Patience C., Awwad M., Fishman J.A.,<br />

Robson S.C., Schuurman H.J., Sachs D.H. & Cooper D.K.<br />

(2005). – Heart transplantation in baboons using<br />

α1,3-galactosyltransferase gene-knockout pigs as donors:<br />

initial experience. Nature Med., 11 (1), 29-31.<br />

53. Lai L., Kolber-Simonds D., Park K.W., Cheong H.T.,<br />

Greenstein J.L., Im G.S., Samuel M., Bonk A., Rieke A.,<br />

Day B.N., Murphy C.N., Carter D.B., Hawley R.J.<br />

& Prather R.S. (2002). – Production of<br />

α1,3-galactosyltransferase knockout pigs by nuclear transfer<br />

cloning. Science, 295 (5557), 1089-1092.<br />

54. Lavitrano M., Camaioni A., Fazio V.M., Dolci S., Farace M.G.<br />

& Spadafora C. (1989). – Sperm cells as vectors for<br />

introducing foreign DNA into eggs: genetic transformation of<br />

mice. Cell, 57 (5), 717-723.<br />

55. Lavitrano M., Bacci M.L., Forni M., Lazzereschi D.,<br />

Di Stefano C., Fioretti D., Giancotti P., Marfe G., Pucci L.,<br />

Renzi L., Wang H., Stoppacciaro A., Stassi G., Sargiacomo M.,<br />

Sinibaldi P., Turchi V., Giovannoni R., Della Casa G., Seren E.<br />

& Rossi G. (2002). – Efficient production by sperm-mediated<br />

gene transfer of human decay accelerating factor (hDAF)<br />

transgenic pigs for xenotransplantation. Proc. natl Acad. Sci.<br />

USA, 99 (22), 14230-14235.<br />

56. Li Z. & Engelhardt J.F. (2003). – Progress toward generating<br />

a ferret model of cystic fibrosis by somatic cell nuclear<br />

transfer. Reprod. Biol. Endocrinol., 1 (1), 83.


Rev. sci. tech. Off. int. Epiz., 24 (1) 297<br />

57. Lo D., Pursel V., Linton P.J., Sandgren E., Behringer R.,<br />

Rexroad C., Palmiter R.D. & Brinster R.L. (1991). –<br />

Expression of mouse IgA by transgenic mice, pigs and sheep.<br />

Eur. J. Immunol., 21 (4), 1001-1006.<br />

58. Ma J.K., Drake P.M. & Christou P. (2003). – The production<br />

of recombinant pharmaceutical proteins in plants. Nat. Rev.<br />

Genet., 4 (10), 794-805.<br />

59. Maga E.A., Anderson G.B. & Murray J.D. (1995). – The<br />

effects of mammary gland expression of human lysozyme on<br />

the properties of milk from transgenic mice. J. Dairy Sci.,<br />

78 (12), 2645-2652.<br />

60. Maga E.A. & Murray J.D. (1995). – Mammary gland<br />

expression of transgenes and the potential for altering the<br />

properties of milk. Biotechnology, 13 (13), 1452-1457.<br />

61. Mahmoud T.H., McCuen B.W., Hao Y., Moon S.J.,<br />

Tatebayashi M., Stinnett S., Petters R.M. & Wong F. (2003). –<br />

Lensectomy and vitrectomy decrease the rate of<br />

photoreceptor loss in rhodopsin P347L transgenic pigs.<br />

Graefes Arch. Clin. Exp. Ophtalmol., 241 (4), 298-308.<br />

62. Massoud M., Attal J., Thepot D., Pointu H., Stinnakre M.G.,<br />

Theron M.C., Lopez C. & Houdebine L.M. (1996). – The<br />

deleterious effects of human erythropoietin gene driven by<br />

the rabbit whey acidic protein gene promoter in transgenic<br />

rabbits. Reprod. Nutr. Dev., 36 (5), 555-563.<br />

63. Meade H.M., Echelard Y., Ziomek C.A., Young M.W.,<br />

Harvey M., Cole E.S., Groet S. & Curling J.M. (1999). –<br />

Expression of recombinant proteins in the milk of transgenic<br />

animals. In Gene expression systems: using nature for the art<br />

of expression (J.M. Fernandez & J.P. Hoeffler, eds). Academic<br />

Press, San Diego, 399-427.<br />

64. Müller M. & Brem G. (1991). – Disease resistance in farm<br />

animals. Experientia, 47 (9), 923-934.<br />

65. Müller M., Brenig B., Winnacker E.L. & Brem G. (1992). –<br />

Transgenic pigs carrying cDNA copies encoding the murine<br />

Mx1 protein which confers resistance to influenza virus<br />

infection. Gene, 121 (2), 263-270.<br />

66. Niemann H. (2004). – Transgenic pigs expressing plant<br />

genes. Proc. natl Acad. Sci. USA, 101 (19), 7211-7212.<br />

67. Niemann H., Halter R., Carnwath J.W., Herrmann D.,<br />

Lemme E. & Paul D. (1999). – Expression of human blood<br />

clotting factor VIII in the mammary gland of transgenic<br />

sheep. Transgenic Res., 8 (3), 237-247.<br />

68. Niemann H. & Kues W.A. (2003). – Application of<br />

transgenesis in livestock for agriculture and biomedicine.<br />

Anim. Reprod. Sci., 79 (3-4), 291-317.<br />

69. Nottle M.B., Nagashima H., Verma P.J., Du Z.T., Grupen C.G.,<br />

MacIlfatrick S.M., Ashman R.J., Harding M.P., Giannakis C.,<br />

Wigley P.L., Lyons I.G., Harrison D.T., Luxford B.G.,<br />

Campbell R.G., Crawford R.J. & Robins A.J. (1999). –<br />

Production and analysis of transgenic pigs containing a<br />

metallothionein porcine growth hormone gene construct.<br />

In Transgenic animals in agriculture (J.D. Murray,<br />

G.B. Anderson, A.M. Oberbauer & M.M. McGloughlin, eds).<br />

CABI Publishing, New York, 145-156.<br />

70. Palmarini M. & Fan H. (2001). – Retrovirus-induced ovine<br />

pulmonary adenocarcinoma: an animal model for lung<br />

cancer. J. natl Cancer Inst., 93 (21), 1603-1614.<br />

71. Perry A.C., Wakayama T., Kishikawa H., Kasai T., Okabe M.,<br />

Toyoda Y. & Yanagimachi R. (1999). – Mammalian<br />

transgenesis by intracytoplasmic sperm injection. Science,<br />

284 (5417), 1180-1183.<br />

72. Perry A.C., Rothman A., de las Heras J.I., Feinstein P.,<br />

Mombaerts P., Cooke H.J. & Wakayama T. (2001). – Efficient<br />

metaphase II transgenesis with different transgene archetypes.<br />

Nature Biotechnol., 19 (11), 1071-1073.<br />

73. Petters R.M., Alexander C.A., Wells K.D., Collins E.B.,<br />

Sommer J.R., Blanton M.R., Rojas G., Hao Y., Flowers W.L.,<br />

Banin E., Cideciyan A.V., Jacobson S.G. & Wong F. (1997). –<br />

Genetically engineered large animal model for studying cone<br />

photoreceptor survival and degeneration in retinitis<br />

pigmentosa. Nature Biotechnol., 15 (10), 965-970.<br />

74. Phelps C.J., Koike C., Vaught T.D., Boone J., Wells K.D.,<br />

Chen S.H., Ball S., Specht S.M., Polejaeva I.A., Monahan J.A.,<br />

Jobst P.M., Sharma S.B., Lamborn A.E., Garst A.S., Moore M.,<br />

Demetris A.J., Rudert W.A., Bottino R., Bertera S., Trucco M.,<br />

Starzl T.E., Dai Y. & Ayares D.L. (2003). – Production of<br />

α1,3-galactosyltransferase deficient pigs. Science, 299 (5605),<br />

411-414.<br />

75. Plasterk R.H. (2002). – RNA silencing: the genome’s immune<br />

system. Science, 296 (5571), 1263-1265.<br />

76. Platenburg G.J., Kootwijk E.A.P., Kooiman P.M.,<br />

Woloshuk S.L., Nuijens J.H., Krimpenfort P.J.A., Pieper F.R.,<br />

de Boer H.A. & Strijker R. (1994). – Expression of human<br />

lactoferrin in milk of transgenic mice. Transgenic Res., 3 (2),<br />

99-108.<br />

77. Platt J.L. & Lin S.S. (1998). – The future promises of<br />

xenotransplantation. Ann. N.Y. Acad. Sci., 862, 5-18.<br />

78. Pursel V.G., Pinkert C.A., Miller K.F., Bolt D.J.,<br />

Campbell R.G., Palmiter R.D., Brinster R.L. & Hammer R.E.<br />

(1989). – Genetic engineering of livestock. Science,<br />

244 (4910), 1281-1288.<br />

79. Pursel V.G., Wall R.J., Mitchell A.D., Elsasser T.H.,<br />

Solomon M.B., Coleman M.E., Mayo F. & Schwartz R.J.<br />

(1999). – Expression of insulin-like growth factor-I in skeletal<br />

muscle of transgenic pigs. In Transgenic animals in<br />

agriculture (J.D. Murray, G.B. Anderson, A.M. Oberbauer &<br />

M.M. McGloughlin, eds). CABI Publishing, New York,<br />

131-144.<br />

80. Rapacz J. & Hasler-Rapacz J. (1989). – Animal models: the<br />

pig. In Genetic factors in atherosclerosis: approaches and<br />

model systems (R.S. Sparkes & A.J. Lusis, eds). Karger, Basle,<br />

139-169.<br />

81. Rossant J. (2001). – Stem cells from the mammalian<br />

blastocyst. Stem Cells, 19 (6), 477-482.<br />

82. Rudolph N.S. (1999). – Biopharmaceutical production in<br />

transgenic livestock. Trends Biotechnol., 17 (9), 367-374.


298<br />

83. Saeki K., Matsumoto K., Kinoshita M., Suzuki I., Tasaka Y.,<br />

Kano K., Taguchi Y., Mikami K., Hirabayashi M.,<br />

Kashiwazaki N., Hosoi Y., Murata N. & Iritani A. (2004). –<br />

Functional expression of a Delta12 fatty acid desaturase gene<br />

from spinach in transgenic pigs. Proc. natl Acad. Sci. USA,<br />

101 (17), 6361-6366.<br />

84. Schätzlein S., Lucas-Hahn A., Lemme E., Kues W.A.,<br />

Dorsch M., Manns M.P., Niemann H. & Rudolph K.L. (2004).<br />

– Telomere length is reset during early mammalian<br />

embryogenesis. Proc. natl Acad. Sci. USA, 101 (21), 8034-<br />

8038.<br />

85. Schätzlein S. & Rudolph K.L. (2005). – Telomere length<br />

regulation during cloning, embryogenesis and aging. Reprod.<br />

Fertil. Dev., 17 (1-2), 85-96.<br />

86. Schnieke A.E., Kind A.J., Ritchie W.A., Mycock K., Scott A.R.,<br />

Ritchie M., Wilmut I., Colman A. & Campbell K.H. (1997).<br />

– Human factor IX transgenic sheep produced by transfer of<br />

nuclei from transfected fetal fibroblasts. Science, 278 (5346),<br />

2130-2133.<br />

87. Shim H., Gutierrez-Adan A., Chen L.R., BonDurant R.H.,<br />

Behboodi E. & Anderson G.B. (1997). – Isolation of<br />

pluripotent stem cells from cultured porcine primordial germ<br />

cells. Biol. Reprod., 57 (5), 1089-1095.<br />

88. Staeheli P. (1991). – Intracellular immunization: a new<br />

strategy for producing disease-resistant transgenic livestock?<br />

Trends Biotechnol., 9 (3), 71-72.<br />

89. Stinnakre M.G., Vilotte J.L., Soulier S. & Mercier J.C. (1994).<br />

– Creation and phenotypic analysis of α-lactalbumindeficient<br />

mice. Proc. natl Acad. Sci. USA, 91 (14), 6544-6548.<br />

90. Swanson M.E., Martin M.J., O’Donnell J.K., Hoover K.,<br />

Lago W., Huntress V., Parsons C.T., Pinkert C.A., Pilder S. &<br />

Logan J.S. (1992). – Production of functional human<br />

hemoglobin in transgenic swine. Biotechnology, 10 (5), 557-<br />

559.<br />

91. Templin M.F., Stoll D., Schrenk M., Traub P.C., Vohringer C.F.<br />

& Joos T.O. (2002). – Protein microarray technology. Trends<br />

Biotechnol., 20 (4), 160-166.<br />

92. Theuring F., Thunecke M., Kosciessa U. & Turner J.D.<br />

(1997). – Transgenic animals as models of neurodegenerative<br />

disease in humans. Trends Biotechnol., 15 (8), 320-325.<br />

93. Van Berkel P.H., Welling M.M., Geerts M., van Veen H.A.,<br />

Ravensbergen B., Salaheddine M., Pauwels E.K., Pieper F.,<br />

Nuijens J.H. & Nibbering P.H. (2002). – Large scale<br />

production of recombinant human lactoferrin in the mik of<br />

trangenic cows. Nature Biotechnol., 20 (5), 484-487.<br />

94. Van den Hout J.M., Reuser A.J., de Klerk J.B., Arts W.F.,<br />

Smeitink J.A. & Van der Ploeg A.T. (2001). – Enzyme therapy<br />

for Pompe disease with recombinant human α-glucosidase<br />

from rabbit milk. J. inherited metabol. Dis., 24 (2), 266-274.<br />

95. Wall R.J., Powell A., Paape M.J., Kerr D.E.,<br />

Bannermann D.D., Pursel V.G., Wells K.D., Talbot N.<br />

& Hawk H. (2005). Genetically enhanced cows resist<br />

intramammary Staphylococcus aureus infection.<br />

Nat. Biotechnol. 23 (4), 445-451.<br />

Rev. sci. tech. Off. int. Epiz., 24 (1)<br />

96. Ward K.A. (2000). – Transgene-mediated modifications to<br />

animal biochemistry. Trends Biotechnol., 18 (3), 99-102.<br />

97. Weidle U.H., Lenz H. & Brem G. (1991). – Genes encoding<br />

a mouse monoclonal antibody are expressed in transgenic<br />

mice, rabbits and pigs. Gene, 98 (2), 185-191.<br />

98. Weissmann C., Enari M., Klohn P.C., Rossi D. & Flechsig E.<br />

(2002). – Transmission of prions. Proc. natl Acad. Sci. USA,<br />

99 (Suppl. 4), 16378-16383.<br />

99. Wheeler M.B. (1994). – Development and validation of<br />

swine embryonic stem cells: a review. Reprod. Fertil. Dev.,<br />

6 (5), 563-568.<br />

100. Wheeler M.B., Bleck G.T. & Donovan S.M. (2001). –<br />

Transgenic alteration of sow milk to improve piglet growth<br />

and health. Reproduction, 58 (Suppl.), 313-324.<br />

101. White D. (1996). – Alteration of complement activity: a<br />

strategy for xenotransplantation. Trends Biotechnol.,<br />

14 (1), 3-5.<br />

102. Wigdorovitz A., Mozgovoj M., Santos M.J., Parreno V.,<br />

Gomez C., Perez-Filgueira D.M., Trono K.G., Rios R.D.,<br />

Franzone P.M., Fernandez F., Carrillo C., Babiuk L.A.,<br />

Escribano J.M. & Borca M.V. (2004). – Protective lactogenic<br />

immunity conferred by an edible peptide vaccine to bovine<br />

rotavirus produced in transgenic plants. J. gen. Virol.,<br />

85 (Pt 7), 1825-1832.<br />

103. Wilmut I., Beaujean N., De Sousa P.A., Dinnyes A., King<br />

T.J., Paterson L.A., Wells D.N. & Young L.E. (2002). –<br />

Somatic cell nuclear transfer. Nature, 419 (6907), 583-586.<br />

104. Wiznerowicz M. & Trono D. (2005). – Harnessing HIV for<br />

therapy, basic research and biotechnology.<br />

Trends Biotechnol., 23 (1), 42-47.<br />

105. Yamada K., Yazawa K., Shimizu A., Iwanaga T.,<br />

Hisashi Y., Nuhn M., O’Malley P., Nobori S., Vagefi P.A.,<br />

Patience C., Fishman J., Cooper D.K., Hawley R.J.,<br />

Greenstein J., Schuurman H.J., Awwad M., Sykes M. &<br />

Sachs D.H. (2005). – Marked prolongation of porcine renal<br />

xenograft survival in baboons through the use of<br />

α1,3-galactosyltransferase gene-knockout donors and the<br />

cotransplantation of vascularized thymic tissue. Nature<br />

Med., 11 (1), 32-34.<br />

106. Yom H.C. & Bremel R.D. (1993). – Genetic engineering of<br />

milk composition: modification of milk components in<br />

lactating transgenic animals. Am. J. clin. Nutr., 58 (2 Suppl.),<br />

299S-306S.<br />

107. Ziomek C.A. (1998). – Commercialization of proteins<br />

produced in the mammary gland. Theriogenology,<br />

49 (1), 139-144.


The FASEB Journal Express article fj.05-5415fje. Published online May 9, 2006.<br />

The FASEB Journal FJ Express Full-Length Article<br />

Epigenetic silencing and tissue independent expression<br />

of a novel tetracycline inducible system in doubletransgenic<br />

pigs<br />

Wilfried A. Kues,* Reinhard Schwinzer, † Dagmar Wirth, ‡ Els Verhoeyen, ‡,1<br />

Erika Lemme,* Doris Herrmann,* Brigitte Barg-Kues,* Hansjörg Hauser, ‡<br />

Kurt Wonigeit, †,2 and Heiner Niemann* ,2<br />

*Department of Biotechnology, Institute for Animal Breeding (FAL), Mariensee, Germany; † Klinik<br />

für Viszeral- und Transplantationschirurgie, Medizinische Hochschule Hannover, Hannover,<br />

Germany; and ‡ Molecular Biotechnology, Gesellschaft für Biotechnologische Forschung mbH,<br />

Braunschweig, Germany<br />

ABSTRACT The applicability of tightly regulated<br />

transgenesis in domesticated animals is severely hampered<br />

by the present lack of knowledge of regulatory<br />

mechanisms and the long generation intervals. To<br />

capitalize on the tightly controlled expression of mammalian<br />

genes made possible by using prokaryotic control<br />

elements, we have used a single-step transduction<br />

to introduce an autoregulative tetracycline-responsive<br />

bicistronic expression cassette (NTA) into transgenic<br />

pigs. Transgenic pigs carrying one NTA cassette showed<br />

a mosaic transgene expression restricted to single muscle<br />

fibers. In contrast, crossbred animals carrying two<br />

NTA cassettes with different transgenes, revealed a<br />

broad tissue-independent and tightly regulated expression<br />

of one cassette, but not of the other one. The<br />

expression pattern correlated inversely with the methylation<br />

status of the NTA transcription start sites indicating<br />

epigenetic silencing of one NTA cassette. This<br />

first approach on tetracycline regulated transgene expression<br />

in farm animals will be valuable for developing<br />

precisely controlled expression systems for transgenes<br />

in large animals relevant for biomedical and<br />

agricultural biotechnology.—Kues, W.A., Schwinzer,<br />

R., Wirth, D., Verhoeyen, E., Lemme, E., Hermann, D.,<br />

Barg-Kues, B., Hauser, H., Wonigeit, K., and Niemann,<br />

H. Epigenetic silencing and tissue independent expression<br />

of a novel tetracycline inducible system in doubletransgenic<br />

pigs. FASEB J. 20, E000–E000 (2006)<br />

Key Words: autoregulative cassette bicistronic transgenic<br />

livestock tissue independent expression DNA methylation<br />

Transgenic mice and farm animals harboring the first<br />

generation of conditional promoter elements, which<br />

were responsive to heavy metals or steroid hormones,<br />

suffered from high basal expression levels and pleiotropic<br />

effects (1–4). In contrast, binary expression systems<br />

based on prokaryotic control elements, which are responsive<br />

to exogenous IPTG, RU-486, ecdysone, or<br />

tetracycline (tet) derivatives are compatible with a<br />

0892-6638/06/0020-0001 © FASEB<br />

tightly controlled expression (5, 6). Several transgenic<br />

mouse lines expressing modified tet-system components<br />

in various tissues have been generated and a<br />

rapid and reversible transgene regulation has been<br />

shown (7–13). Factors affecting the efficiency of tetracycline-dependent<br />

gene regulation include the strain of<br />

mice and the site of integration (14), as well as clearance<br />

of tetracycline derivatives (15). A high degree of<br />

variability in activator-dependent target gene expression<br />

is observed and high background expression<br />

makes it necessary to screen for appropriately responding<br />

lines (16).<br />

The original tetracycline system requires two DNA<br />

constructs for expression of the transactivator and<br />

transactivator dependent expression of the target gene,<br />

respectively. These DNA-constructs are usually integrated<br />

in two different lines of transgenic mice. On<br />

breeding, offspring are obtained in which the target<br />

gene can be regulated by addition of doxycycline (7).<br />

The long-generation intervals make this approach unfeasible<br />

in other mammalian species, including livestock<br />

(17). One approach to overcome this limitation is<br />

to combine both transactivator and target genes, including<br />

individual promoters in a single construct. This<br />

approach was compatible with efficient control of gene<br />

expression in mammalian cells and mice (18, 19).<br />

However, the near physical distance of two promoter<br />

sequences makes this system prone to elevated background<br />

expression. A tetracycline-controlled transcriptional<br />

silencer has been used to suppress background<br />

expression (9). Alternatively, autoregulatory expression<br />

1 Present address: Envelope retrovirale et ingeniérerie retrovirale,<br />

INSERM 758, ENS de Lyon, France.<br />

2 Correspondence: K. W., Klinik f. Viszeral-und Transplantations<br />

Clinique, Medizinische Hochschule Hannover, Carl-<br />

Neuberg Str. 1, Hannover D-30623, Germany. E-mail:<br />

wonigeit.kurt@mh-hannover.de H. N., Institut for Animal<br />

Breeding (FAL) Department of Biotechnology Höltystr. 10,<br />

Neustadt D-31535, Germany. E-mail: niemann@tzv.fal.de<br />

doi: 10.1096/fj.05-5415fje<br />

E1


systems have been established that overcome the problem<br />

of promoter interference. In these systems the<br />

transactivator is either expressed from a tet-responsive<br />

bicistronic expression cassette (20–23) or from a tetresponsive<br />

bidirectional promoter (24, 25). It has been<br />

shown that expression of biologically active proteins<br />

can be tightly controlled with these systems, al<strong>bei</strong>t<br />

minimum basal expression levels of the transactivator<br />

are required (26). Another advantage of autoregulative<br />

cassettes is that transgene transcription is linked to<br />

ubiquitously present cellular cofactors, thereby circumventing<br />

cell-/tissue-specific expression as observed for<br />

classical promoters.<br />

Here, we describe tightly controlled transgene overexpression<br />

in the domestic pig by an autoregulative<br />

bicistronic tetracycline-responsive cassette (NTA). The<br />

NTA-cassette was designed to ensure high and ubiquitous<br />

expression levels of human regulators of complement<br />

activation (RCA) relevant for xenotransplantation<br />

(17), (i.e., decay accelerating factor (DAF)) and CD59.<br />

The rationale for this approach was twofold: (i) As<br />

proof of principle we wanted to demonstrate that<br />

single-construct cassettes can be successfully expressed<br />

in large animals; (ii) we intended to improve xenotransplantation<br />

by producing tet-off regulated human<br />

complement regulator transgenic animals. Unexpectedly,<br />

nine independent transgenic pig lines showed<br />

expression restricted to muscle fibers. However, in<br />

crossbred double-transgenic animals a broad tissue<br />

independent transgene expression of the CD59-NTA<br />

cassette, but not of the DAF-NTA cassette was discovered.<br />

This asymmetric expression pattern correlated<br />

with different methylation patterns of the NTA cassettes<br />

indicating that epigenetic mechanisms are critical<br />

for the function of the NTA system. Expression of<br />

human complement regulators and tet transactivator<br />

did not compromise animal health and fertility.<br />

MATERIALS AND METHODS<br />

Construction of the bicistronic NTA-cassettes<br />

CD59 was amplified from pWTCD59 (27). DAF was amplified<br />

from pWTDAF (27). The transactivator tTA was amplified<br />

from pUHD15–1 (28), thereby flanking the tTA cDNA with<br />

NotI and PstI sites. The cDNAs were independently integrated<br />

into pTBC-1 (29) and both plasmids were then fused via the<br />

unique BamHI and VspI sites. All polymerase chain reaction<br />

(PCR)-derived sequences were verified by DNA sequencing.<br />

Vector sequences are available on request.<br />

Cell culture and transfection<br />

NIH3T3 cells were cultivated in Dulbecco’s modified Eagle’s<br />

medium (DMEM). The RCA-NTA cassettes were cotransfected<br />

with a pAG60 by DNA-calcium phosphate coprecipitation<br />

(27). Cells were selected for resistance to G418 (1<br />

mg/ml). Pools and cell clones were generated and cultivated<br />

in the presence and absence of doxycycline (2 g/ml).<br />

Transfection of swine testis epitheloid (STE) cell line (kindly<br />

provided by A. Saalmüller, Wien, Austria) was carried out<br />

using essentially the same protocol as for NIH3T3 cells. Flow<br />

cytometry was performed as described earlier (27) using<br />

specific monoclonal antibodies (PharMingen, Hamburg, Germany)<br />

to detect CD59 and DAF.<br />

Determination of doxycyline plasma levels in pigs<br />

A mouse-derived cell line NIH-TAluc with a stably integrated<br />

doxycycline-responsive luciferase cassette was used to determine<br />

doxycycline levels in plasma and fodder extracts.<br />

Generation of transgenic pigs<br />

Transgenic pigs were generated by pronuclear DNA microinjection<br />

as described (30). In brief, zygotes were collected<br />

from hormonally pretreated prepubertal German landrace<br />

donor gilts 15–18 h after the second mating. Animals were<br />

sacrificed, and zygotes were flushed from the excised oviducts.<br />

The NTA constructs were released from vector backbones<br />

by digesting the plasmids with BsrBI and Tth111I (New<br />

England Biolabs (NEB), Frankfurt, Germany) and isolated by<br />

gel electrophoresis. Approximately 10 pl of purified DNAs<br />

(4–10 ng/l) were injected into one pronucleus of zygotes<br />

after centrifugation for 6 min at 14,500 g to visualize the<br />

pronuclei by concentrating the cytoplasmic lipids at one pole.<br />

A total of 653 microinjected zygotes were surgically transferred<br />

into the oviducts of 20 synchronized prepubertal gilts,<br />

resulting in 80 piglets from which 11 could be identified as<br />

transgenic by Southern blotting of the genomic DNA. For<br />

Southern blotting, the genomic DNA was digested with PstI<br />

(NEB) and blotted to Hybond membranes. The PstI digest<br />

released internal fragments from the NTA-cassettes, which<br />

were 2049 and 1658 bp for the CD59 and DAF-constructs,<br />

respectively. The complete constructs were isolated from<br />

vector backbones and used as probes after digoxigenin labeling.<br />

Hybridization products were visualized by chemiluminescence.<br />

Founder animals were bred with nontransgenic animals<br />

to test for germline transmission. Transgenic F1 and<br />

F2-animals were used for detailed analysis of the expression<br />

pattern and doxycycline responsiveness. Animals were treated<br />

according to institutional guidelines and experiments were<br />

approved by an external animal welfare committee.<br />

Detection of CD59 and DAF expression<br />

Transgenic offspring and wild-type control animals were<br />

sacrificed and tissue samples were obtained from the following<br />

organs: heart, lung, liver, kidney, spleen, thymus, skeletal<br />

muscle, pancreas, skin, brain, and aorta. Samples were used<br />

for primary cell cultures or snap-frozen in liquid nitrogen and<br />

used for RT–PCR, Northern blotting and immunohistological<br />

analysis. For RT–PCR, total RNA was isolated using the<br />

guanidinium thiocyanate-phenol-chloroform extraction. PCR<br />

detection and Northern blotting of CD59 and DAF transcripts<br />

were done as described recently (30). In brief, digoxigenin<br />

labeled 605 and 613 bp antisense cRNAs for CD59 and DAF<br />

were used for specific detection of the transgenic transcripts.<br />

The calculated sizes for the bicistronic mRNAs were 2.3 and<br />

3.0 kb for the CD59 and DAF encoding transcripts, respectively.<br />

In addition to the 3.0 kb band a slightly faster migrating<br />

fragment was specifically detected in the DAF Northern.<br />

Apparently, the smaller fragment represented a degradation<br />

product of the 3.0 kb mRNA. In some experiments tissues<br />

collected from transgenic animals carrying a cytomegalovirus<br />

(CMV) promoter-CD59 construct were used as controls (30).<br />

A -actin (Actb) probe (31) served as an internal control.<br />

E2 Vol. 20 June 2006 The FASEB Journal<br />

KUES ET AL.


To detect CD59 and DAF, immunohistological staining was<br />

performed on cryostat sections using biotinylated anti-human<br />

mAB H19 (CD59) and mAB33572X (DAF/CD55), respectively<br />

(both antibodies were from BD PharMingen). Human<br />

heart tissue served as positive control. Specific antibody (Ab)<br />

binding was detected by incubation with streptavidine-peroxidase<br />

and 3-amino-9-ethylcarbazole (AEC) as substrate, cell<br />

nuclei were counterstained with hematoxylin. Tissue samples<br />

were evaluated at 100–200 magnifications.<br />

Primary cell cultures of porcine fibroblasts were established<br />

from subdermal tissues as described (30, 32) and employed<br />

for expression analysis by Northern blotting or flow cytometry<br />

(FACS).<br />

Genomic DNA was isolated from porcine tissues by proteinase<br />

K lysis. The following primers spanning the complete<br />

promoter regions of DAF-NTA and CD59-NTA constructs were<br />

used to amplify these regions by PCR: CMV-0, 5-AAA ATA<br />

GGC GTA CAC GAG G; hDAF-1, 5-ATC ACT GAG TCC TTC<br />

TCG CC and hCD59–1, 5-CCC TCA AGC GGG TTG TGA<br />

CG. The PCR products were directly sequenced, employing<br />

the CMV-O primer.<br />

Application of doxycycline<br />

Doxycycline was fed in the form of tablets (doxy 200®, CT<br />

Arzneimittel GmbH, Berlin, Germany; one tablet contains<br />

200 mg doxycycline). A daily dose of 3.3 mg/kg body weight<br />

was applied for periods from 6 to 45 d. In one feeding group,<br />

blood samples were taken during the course of the experiment<br />

to determine doxycycline levels in plasma. Muscle<br />

biopsies from the M. longissimus were obtained by a special<br />

biopsy device routinely used in meat quality testing program<br />

(33). Biopsies were taken from the longissimus muscle prior<br />

to doxycycline treatment, and at different time points thereafter<br />

and were analyzed by Northern blotting and immunohistology.<br />

In some experiments, subdermal tissue pieces from<br />

the biopsies were used to establish primary cell cultures.<br />

Bioinformatic analysis of CpG islands<br />

For CpG island prediction the MethPrimer program (http://<br />

www.urogene.org/methprimer) was employed. The NTA-cassette<br />

sequences were analyzed using the following criteria for<br />

CpG island prediction: i) a CG content greater than 60%, ii)<br />

a presence of CpG dinucleotides of greater than 0.6, and iii)<br />

a DNA region of greater than 300 bp.<br />

Detection of methylated CpG dinucleotides by bisulfite<br />

sequencing<br />

Bisulfite sequencing was done as described by Hajkova et al.<br />

(34). In brief, 20 ng genomic DNA was digested with EcoRI<br />

(NEB). Denatured DNA was embedded in 7 l of 20 mg/ml<br />

low melting agarose (Biozym, Hess. Oldendorf, Germany),<br />

cooled to form agarose beads, and incubated in 2.5 M<br />

bisulfite-hydroquinone solution pH 5 (Roth, Hamburg, Germany)<br />

for4hat50°C. PCR was carried out in a final vol of 100<br />

l with individual agarose beads containing embedded<br />

genomic DNA. PCR conditions were as follows: 1 PCR<br />

reaction buffer (20 mM Tris-HCl pH 8.4, 50 mM KCl 2), 1.5<br />

mM MgCl 2, 0.2 M dNTPs (Amersham Biosciences Europe,<br />

Freiburg, Germany), and 0.6 M of each primer for 40 cycles.<br />

The PCR fragments were separated by gel electrophoresis,<br />

isolated and directly sequenced. Only sequences with 95%<br />

cytosine conversion of non-CpGs were analyzed. Methylation<br />

status of CpG dinucleotides was determined based on the<br />

ratio of cytosines to converted cytosines. The following prim-<br />

INDUCIBLE TRANSGENIC EXPRESSION IN A LARGE ANIMAL MODEL<br />

ers were used to amplify the CMV minimal promoter of the<br />

CD59-NTA construct, CMV-2bi, 5- GAA AGT TGA GTT CGG<br />

TAT TT, CD59–2bi, 5- AAA AAT ATC CCA CCT TTT TC; for<br />

the DAF-NTA construct the CMV-2bi primer in combination<br />

with DAF-2bi, 5- CCA TTA ACT ACC CTT AAA AC was used.<br />

RESULTS<br />

Evaluation of autoregulatory RCA–NTA expression<br />

cassettes in cell lines<br />

We constructed an autoregulatory bicistronic tet-off<br />

based cassette that supports single step transduction<br />

(Fig. 1A) and in which the mRNA is driven by the tet<br />

responsive promoter (P tTA). In two constructs carrying<br />

either the human CD59 or DAF cDNAs (designated<br />

CD59-NTA and DAF-NTA) the RCA reading frames are<br />

followed by an internal ribosomal entry site (IRES) that<br />

provides translation of the recombinant transactivator<br />

(tTA). Except for the human cDNAs, both cassettes<br />

were completely identical.<br />

Mouse NIH 3T3 cells stably expressing the cassettes<br />

showed autoactivation and expression of the respective<br />

RCA (Fig. 1B). Expression was efficiently suppressed by<br />

addition of doxycycline to the culture medium (Fig.<br />

1B). In addition, swine testis epitheloid (STE) cells<br />

were transfected with the RCA-NTA cassettes, and again<br />

autoactivation and conditional transgene regulation<br />

were found (not shown). Thus, autoregulatory bicistronic<br />

expression cassettes support tight regulation of<br />

human RCAs in xenogenic cells, suggesting that ubiquitous<br />

activation could be achieved in vivo.<br />

Generation of RCA-NTA transgenic pigs<br />

Five hCD59-NTA and six DAF-NTA transgenic founder<br />

animals were generated by pronuclear microinjection<br />

of the respective cassette into porcine zygotes, from<br />

which ten transgenic lines were established. Transgenic<br />

juvenile and adult F1 and F2-offspring were used for<br />

expression profiling in the absence of any tetracycline.<br />

Remarkably, an intensive immunostaining indicative<br />

for human RCA protein was found in single fibers of<br />

striated muscle in 9 of the 10 lines (Table 1, Fig. 1C; 2–3<br />

animals investigated per line). Depending on the line,<br />

5–30% of the muscle fibers stained positive. Expression<br />

in muscle was confirmed by Northern blotting with DAF<br />

and CD59 specific probes, respectively. Contrary to the<br />

tissue-independent design of the NTA cassette, Northern<br />

blotting, immunohistology, and FACS revealed that<br />

the animals did not express human RCAs in brain,<br />

heart, kidney, tongue, liver, skin, lymph nodes, blood<br />

and pancreas (summarized in Table 1). Apparently, the<br />

(different) integration sites of the nine expressing lines<br />

had no or only a marginal effect on the muscle fiberspecific<br />

expression.<br />

Integration of the entire bicistronic cassettes into the<br />

porcine genome was confirmed by Southern blotting of<br />

genomic DNA digested with restriction enzymes, cut-<br />

E3


Figure 1. Design of the autoregulative NTA<br />

expression cassette and conditional expression<br />

of transgenes in cell lines and in singletransgenic<br />

pigs. A) The P tTA promoter (28)<br />

drives a bicistronic mRNA encoding a human<br />

regulator of complement activation (RCA)<br />

linked with a tet-off transactivator (tTA) via a<br />

poliovirus IRES. Two constructs carrying the<br />

human cDNAs encoding either CD59 (CD59-<br />

NTA) orDAF(DAF-NTA) were used. Minimal<br />

expression of the transactivator results in tTA<br />

binding to the tet-operator sequences and<br />

initiates an autoregulative loop. In the presence<br />

of exogenous tetracycline the transactivator<br />

is inactivated and transcription silenced.<br />

B) Conditional expression of hDAF in cell<br />

culture. NIH3T3 cells were stably transfected<br />

with DAF-NTA. A cell pool and a single clone<br />

(clone 12) were cultivated in presence (green<br />

line) or absence (black line) of doxycycline.<br />

Cells were harvested and DAF expression was<br />

determined by indirect immunofluorescence using<br />

flow cytometry. An overlay of the histograms<br />

is presented with nontransfected NIH3T3 cells<br />

(red line) as a control. C) Muscle fiber confined<br />

expression of DAF and CD59 in transgenic pigs<br />

bearing a single NTA-cassette. Longitudinal (upper<br />

row) and cross sections (lower row) of<br />

transgenic and wild-type pig muscles after specific<br />

Ab staining are shown, nuclei are counterstained<br />

with hematoxylin. Note that only individual<br />

fibers displayed intense Ab staining,<br />

whereas neighboring fibers are negative. No expression of the transgenes was detected in several other tissues (for details see<br />

Table 1).<br />

ting the NTA cassette internally near the ends (not<br />

shown). To rule out the possibility, that transgenic<br />

animals carried defective promoter sequences, the respective<br />

elements were amplified from porcine<br />

genomic DNA by PCR. Sequencing confirmed that<br />

heptamerized tet-operator and CMV basal promoter<br />

sequences were present in both, CD59-NTA and DAF-<br />

NTA cassettes. To further rule out putative contamination<br />

of the fodder with tetracycline and/or its derivatives,<br />

fodder and plasma samples of transgenic pigs<br />

were analyzed via a doxycycline sensitive reporter cell<br />

line. All samples proved to be negative for tetracycline<br />

activity (data not shown).<br />

Enhanced tissue-independent expression of CD59 in<br />

crossbred animals carrying two NTA-RCA cassettes<br />

To investigate whether the unexpected mosaic expression<br />

pattern in striated muscle was due to suboptimal<br />

transactivator expression levels, tTA was expressed in<br />

porcine primary fibroblasts from transgenic animals.<br />

Although cotransfected tTA-responsive reporter genes<br />

were activated, the chromosomal RCA-NTA constructs<br />

remained unresponsive (data not shown) indicating<br />

stable silencing of the chromosomal cassettes.<br />

To test if the typical mosaic expression pattern in<br />

transgenic pigs could be altered by increased threshold<br />

levels of tTA in vivo, heterozygous transgenic animals<br />

carrying either CD59-NTA or DAF-NTA were crossbred<br />

to obtain double-transgenic animals. From three litters<br />

with a total of 24 piglets the expected mendelian<br />

distribution of transgene combinations was obtained;<br />

i.e., 5 piglets (21%) were double-transgenic (CD59-NTA<br />

DAF-NTA), 12 (50%) carried either the CD59 or DAF<br />

cassette and 7 (29%) were nontransgenic. In addition,<br />

a female and male double-transgenic sibling (CD59-<br />

NTA DAF-NTA) were mated and produced nine<br />

piglets, i.e., four piglets carrying both cassettes, four<br />

piglets had one cassette and one nontransgenic offspring<br />

was obtained. From the nine double-transgenic<br />

animals, three were sacrificed at the age of 6 wk and 6<br />

mo, respectively, and organs were analyzed for expression<br />

of the transgenes. Muscle biopsies and blood<br />

samples were analyzed from the other double-transgenic<br />

animals. Interestingly, a selective up-regulation of<br />

CD59 expression was found in double-transgenic animals<br />

derived from crossbreeding L13 and L20 (Table<br />

1). The CD59 mRNA levels were significantly (50–100<br />

fold) elevated in three different muscles, i.e., Musculus<br />

(M.) longissimus, front and hind leg muscles, in double-transgenic<br />

animals over transgenic siblings carrying<br />

only one CD59-NTA cassette (Fig. 2A). Immunohistological<br />

staining revealed that virtually all fibers in<br />

striated muscle stained positive for CD59 (Fig. 3A). In<br />

addition, other organs, including liver, pancreas, kidney<br />

and lung expressed the transgene at the mRNA<br />

E4 Vol. 20 June 2006 The FASEB Journal<br />

KUES ET AL.


TABLE 1. Expression patterns in animals carrying one and two NTA cassettes<br />

Transgenic lines bearing one NTA cassette<br />

DAF-NTA CD59-NTA<br />

(Fig. 2B) and protein levels (Fig. 3A). RNA and protein<br />

expression pattern of CD59 correlated well, e.g., muscle<br />

and liver, which showed the highest mRNA levels<br />

(Fig. 2A) also displayed the highest proportion of<br />

anti-CD59 Ab stained cells (Fig. 3A). Tissues with lower<br />

mRNAs levels, such as lung, showed only few positive<br />

cells in the immunohistology. Importantly, in all examined<br />

tissues the positive cells were strongly stained for<br />

CD59, suggesting that the autoregulative up-regulation<br />

worked well in this population of cells. However, DAF<br />

mRNA and protein levels were not increased in the<br />

same double-transgenic animals and immunoreactivity<br />

remained confined to single muscle fibers (Fig. 3A).<br />

Seven out of nine double-transgenic pigs, derived from<br />

crossing transgenic lines L13 with L20 showed this<br />

asymmetric over-expression of CD59.<br />

The remaining two double-transgenic pigs, derived<br />

from crossing L15 with L20, mimicked expression<br />

patterns of their parental lines showing mosaic expression<br />

in single muscle fibers. Individual muscle fibers<br />

coexpressed both CD59 and DAF (Fig. 3B), while some<br />

fibers expressed only CD59 (Fig. 3B, arrows), but not<br />

DAF (Fig. 3B, circles), highlighting that in these fibers<br />

high transactivator levels alone were not sufficient to<br />

activate the DAF-NTA gene. Fibers with exclusive expression<br />

of DAF were never detected.<br />

Since none of the single-transgenic animals showed<br />

this enhanced transgene expression, we argue that the<br />

presence of two cassettes is necessary, but not always<br />

sufficient for tissue independent expression. This finding<br />

was confirmed by the widespread expression of<br />

CD59 in a CD59-NTA homozygous animal (CD59-NTA<br />

CD59-NTA) (data not shown).<br />

Health status, development, and fertility of animals<br />

bearing either one or two NTA-cassettes were not<br />

compromised. All animals were not distinguishable<br />

from their nontransgenic counterparts.<br />

Transgenic animals with two cassettes<br />

DAF-NTA CD59-NTA<br />

L13 L20<br />

(n7)<br />

L15 L20<br />

(n2)<br />

L12 L13 L14 L15 L16 L17 L19 L20 L21 L22 DAF CD59 DAF CD59<br />

Heart - - - - - - - - - - - - -<br />

Liver - - - - - - - - - - - - -<br />

Lung - - - - - - - - - - - - -<br />

Kidney - - - - - - - - - - - - -<br />

Pancreas - - - - - - - - - - - - -<br />

Skeletal mus. - <br />

Thymus - - - - - - - - - - n.d. n.d. n.d. n.d.<br />

Lymphocytes - - n.d. n.d. n.d. n.d. n.d. - n.d. n.d * * - -<br />

Fibroblasts n.d - n.d. n.d. n.d. n.d. n.d. - n.d. n.d. - n.d. n.d.<br />

Summarized expression data obtained by Northern blotting, immunohistology and FACS; plus signs indicate low () to high ()<br />

levels of expression, –: below detection limit, n.d.: not determined; *after mitogen stimulation.<br />

INDUCIBLE TRANSGENIC EXPRESSION IN A LARGE ANIMAL MODEL<br />

Figure 2. Asymmetric over-expression in pigs carrying two<br />

cassettes. A) Overexpression of CD59-NTA transcripts in double-transgenic<br />

siblings. Northern blot of total RNAs isolated<br />

from fore limb (lane 1), hind-limb (lane 2) and M. longissimus<br />

(lane 3) of a double-transgenic animal. Lanes 4–6 were<br />

loaded with RNA from fore-limb, hind-limb, and M. longissimus<br />

from a CD59-NTA single-transgenic sibling. Lanes 7 and<br />

8 were loaded with samples from a wild-type animal and lanes<br />

9 and 10 with RNA isolated from a CMV-CD59 transgenic<br />

animal described previously (30). The CD59 transcripts in<br />

lanes 1–3 are connected via an IRES element with the tet<br />

transactivator resulting in a 2.3 kb transcript, whereas the<br />

conventional CMV-hCD59 construct (lanes 9, 10) produces a<br />

1.6 kb transcript. B) Tissue specific expression of CD59 and<br />

DAF in double-transgenic animals. The gel was loaded with<br />

(1) a RNA size marker and total RNA isolated from heart (2),<br />

lung (3), liver (4), kidney (5) and skeletal muscle (6) and<br />

probed with CD59 and DAF specific probes.<br />

E5


Figure 3. Asymmetric expression of CD59 and DAF in doubletransgenic<br />

animals. A) Asymmetric over expression of CD59<br />

in different organs of double-transgenic animal #335. This<br />

animal was produced by breeding L13 with L20. Note the<br />

widespread expression of CD59 in muscle, liver and pancreas<br />

as indicated by a red-brown precipitate. In contrast, DAF<br />

expression is still confined to single muscle fibers (arrows). B)<br />

Coexpression of CD59 and DAF in single muscle fibers of<br />

double-transgenic animal #364 detected by immunostaining<br />

of serial cross sections. This animal was generated by breeding<br />

L15 with L20. Note that some fibers expressed exclusively<br />

CD59 (arrows), but not DAF (dashed circles).<br />

Conditional RCA-NTA regulation in single- and<br />

double-transgenic animals<br />

To evaluate the exogenous control of the autoregulated<br />

tet-off cassettes, 12 single-transgenic and four<br />

double-transgenic animals were fed with doxycycline<br />

and muscle biopsies were taken before, during and<br />

after antibiotic treatment. A daily dose of 3.3 mg<br />

doxycycline/kg body weight for 6 d was found to be<br />

effective to shut down transgene expression (Fig. 4).<br />

This dose is well below the recommended effective dose<br />

of 12 mg doxycycline/kg/day for antibiotic treatment<br />

of pigs. RCA-NTA transcripts were virtually eliminated<br />

by doxycycline feeding within 2–6 d. Thus, the expression<br />

cassettes located at different chromosomal sites<br />

could be readily switched off. Re-expression of RCA-<br />

NTAs took unexpectedly long, and did not resume until<br />

eight weeks after termination of the doxycycline application<br />

(Fig. 4).<br />

Tightly controlled expression of human RCA in white<br />

blood cells<br />

To analyze the regulatory capacity of the NTA cassettes,<br />

we studied the expression patterns of DAF and CD59 by<br />

flow cytometry on white blood cells collected from<br />

seven double-transgenic animals (L13L20), and 2<br />

single-transgenic (1 DAF, 1 CD59) pigs. No significant<br />

expression of DAF or CD59 was found in lymphocytes<br />

(resting and concanavalin A (Con A)-activated cells)<br />

from the two single-transgenic pigs. (Fig. 5A). Absence<br />

of DAF and CD59 was also typical for resting cells from<br />

double-transgenic animals (Fig. 5B), although in one<br />

animal a small number of CD59 expressing cells were<br />

detected (data not shown). However, expression of<br />

DAF and CD59 could be induced in cells from all<br />

double-transgenic animals by activation with ConA.<br />

The intensity of ConA-mediated up-regulation of the<br />

two molecules was asymmetric showing strong expression<br />

of CD59 and weak but significant up-regulation of<br />

DAF. Expression of both molecules could be terminated<br />

by culturing the cells with tetracycline for 48 h<br />

(Fig. 5B), again providing evidence for a tight control<br />

of gene expression from these cassettes.<br />

Identification of CpG islands and determination of<br />

the methylation status<br />

Epigenetic mechanisms were hypothesized to be causatively<br />

involved in the asymmetric expression of the two<br />

RCA-NTA cassettes in double-transgenic animals and<br />

the muscle fiber-specific expression in single-transgenic<br />

Figure 4. Conditional transgene expression by doxycycline<br />

feeding. Two double-transgenic pigs (#368, #369) were fed<br />

with a daily dose of 3.3 mg/kg doxycycline for 6 d. As control,<br />

a third double-transgenic animal (#366) was left untreated.<br />

Muscle biopsies were taken at several time points before (day<br />

0), during (day 2), and after (days 13, 19, 75, and 180) the<br />

feeding regime, 5 g total RNA was loaded per lane and<br />

analyzed by Northern blotting with RCA and -actin (ACTB)<br />

specific probes. Note that the RCA-NTA transcripts disappeared<br />

already after2dofdoxycycline feeding. Repression of<br />

the RCAs resumed 8 wk after the end of doxycycline feeding.<br />

The calculated size of the DAF-IRES-tTA mRNA was 3.0 kb.<br />

E6 Vol. 20 June 2006 The FASEB Journal<br />

KUES ET AL.


Figure 5. Expression patterns of hDAF and hCD59 on lymphocytes<br />

from transgenic pigs. The cells were stained by<br />

indirect immunofluorescence and analyzed on a FACSCalibur<br />

flow cytometer. Closed histograms (thin lines) represent<br />

fluorescence intensity obtained after incubation of the cells<br />

with mouse anti-human DAF (IA10, IgG2a) or CD59 (H19,<br />

IgG2a) monoclonal antibody followed by a second incubation<br />

step with FITC-conjugated goat antimouse immunoglobulin<br />

(FITC-immunoglobulin). Open histograms (bold lines) were<br />

obtained after staining of the cells with FITC-immunoglobulin<br />

only. A) Analysis of lymphocytes obtained from a single<br />

hCD59 transgenic pig. Resting cells and cells, which had<br />

been activated by culturing with Con A plus interleukin (IL)-2<br />

for 48 h were studied. Similar results (absence of hDAF<br />

and hCD59) were obtained when cells from a single DAF<br />

INDUCIBLE TRANSGENIC EXPRESSION IN A LARGE ANIMAL MODEL<br />

pigs. Bioinformatic analysis of the DAF-NTA cassette<br />

revealed that the minimal CMV promoter within P tTA<br />

and the 5coding region of the DAF cDNA together<br />

contained a putative CpG island (Fig. 6A). In contrast,<br />

no CpG island was predicted for the P tTA and the<br />

5coding region of the CD59 gene (Fig. 6A).<br />

Genomic DNA from muscle biopsies of single and<br />

double-transgenic pigs was isolated, and analyzed by<br />

bisulfite sequencing to assess the methylation status of<br />

the NTA-promoter regions in the pig genome.<br />

The promoter region of the CD59-NTA cassette was<br />

completely methylated in single-transgenic animals,<br />

while in double-transgenic animals only 50% of the<br />

CpG dinucleotides were fully methylated (Fig. 6B). In<br />

contrast, this region was essentially nonmethylated in<br />

CD59-NTA cassettes from stably transfected STE cells.<br />

An almost complete methylation of the CpG dinucleotides<br />

around the transcription initiation site of the<br />

DAF-NTA cassette was discovered either in single or<br />

double-transgenic animals (Fig. 6C). Primary fibroblasts<br />

isolated from biopsies also maintained this methylation<br />

pattern (Fig. 6C). For comparison, the methylation<br />

status of the DAF-NTA cassette was determined in<br />

a pool of stably transfected NIH 3T3 cells and in two<br />

cell clones, which expressed high levels of the DAF-NTA<br />

construct. These cells possessed a relative low concentration<br />

of methylation (Fig. 6C). Similarly, stably transfected<br />

porcine STE cells showed weak methylation of<br />

the promoter regions (Fig. 6C). Thus an inverse correlation<br />

between the observed gene expression and the<br />

methylation status of these cassettes is obvious. These<br />

data indicate that the P tTA is prone to methylation in<br />

transgenic pigs, in particular in combination with the<br />

5coding region of the human DAF gene.<br />

DISCUSSION<br />

Here, we report the first conditional and broad transgene<br />

expression in a domestic animal species using an<br />

autoregulative tet-off system. The methylation of transgene<br />

promoters correlated well with expression levels,<br />

suggesting de novo methylation as a crucial mechanism<br />

for expression of RCA-NTA constructs in these pigs.<br />

The bicistronic, autoregulative cassettes were designed<br />

transgenic pig or cells from wild-type nontransgenic control<br />

animals were analyzed. B) Analysis of lymphocytes from a<br />

double- transgenic pig. Expression of DAF and CD59 was first<br />

examined on resting cells. The cells were then activated for<br />

48 h with Con A plus IL-2 and expression of the two<br />

molecules was reanalyzed (activated). Note, the expression of<br />

both CD59 and DAF (arrows) in ConA activated cells. Cultivation<br />

was extended for additional 48h in the absence or<br />

presence of tetracycline (Tet, 1 g/ml). The histograms<br />

shown in the bottom panels (Tet) indicated termination of<br />

human RCA expression in tetracycline treated cells. Cells<br />

which had been cultured for 96 h in the absence of tetracycline<br />

showed the same asymmetric expression pattern (weak<br />

DAF, strong CD59) as the 48 h activated cell population.<br />

E7


Figure 6. CpG rich regions in NTA promoter regions and<br />

methylation status. A) CpG island analysis of the DAF- and<br />

CD59-NTA cassettes with the MethPrimer program. The plots<br />

depict tet operator region, CMV basal promoter and the RCA<br />

coding regions; CpG dinucleotides are indicated by red bars<br />

below the plots; putative CpG islands are indicated as solid<br />

blue areas. Note that the DAF-NTA construct carries a putative<br />

CpG island, which covers the transcription start and the<br />

5coding region. B) Methylation status of the CD59-NTA<br />

transcription start region in transgenic pigs and stably transfected<br />

STE cells. CpG dinucleotides are depicted as circles:<br />

black, fully methylated; gray, half-methylated; white, nonmethylated.<br />

C) Methylation status of the DAF-NTA transcription<br />

start region in transgenic pigs, primary fibroblasts and<br />

stably transfected NIH 3T3 and STE cells.<br />

to be compatible with stable and widespread expression<br />

and carried either CD59 or DAF, followed by an IRES<br />

and the tTA coding sequence. The target genes were<br />

chosen, since human CD59 and DAF are critically<br />

involved in the suppression of the hyperacute rejection<br />

of xenotransplanted porcine organs (35). For autoactivation,<br />

a minimum basal expression of the minimal<br />

promoter is required. The original axiome of the tet<br />

system predicts that in the absence of tetracycline, tTA<br />

binds to the tet operator sequences and activates transcription<br />

(5, 6, 36). The presumed autoregulated function<br />

of the NTA cassettes was found to be fulfilled in<br />

NIH3T3 and porcine STE cells.<br />

Unexpectedly, transgenic pigs carrying one NTA<br />

cassette displayed confined expression in single striated<br />

muscle fibers, irrespective of the transgene or the<br />

integration site. Mosaic expression patterns have also<br />

been described in tet-transgenic mice and, in some<br />

cases, the highest expression levels were found in<br />

skeletal muscle (7, 16, 18). In the present study, confinement<br />

to the muscle compartment could be overcome<br />

by crossbreeding selected double-transgenic pigs<br />

carrying both the DAF-NTA and CD59-NTA cassettes.<br />

This resulted in a widespread expression of CD59 in<br />

muscle, liver, pancreas, lung and other tissues, whereas<br />

DAF expression remained restricted to individual muscle<br />

fibers.<br />

The lack of a widespread expression of DAF/NTA<br />

correlated with the methylation of critical parts of the<br />

tTA dependent promoter and neighboring sequences<br />

and is thus epigenetic in nature. Methylation is usually<br />

associated with additional changes in histone conformation<br />

and chromatin structure, rendering the DNA<br />

region into a compressed state and leading to gene<br />

silencing (37–39).<br />

Although the CD59-NTA transgene did not possess a<br />

predicted CpG island, bisulfite sequencing revealed<br />

that the CpGs around the transcription start site were<br />

fully methylated in single-transgenic animals and that<br />

approximately half of the CpG dinucleotides were<br />

methylated in double-transgenic animals. Studies with<br />

Epstein-Barr virus have shown that even methylation of<br />

few sites is sufficient to suppress transcription of a<br />

promoter, and experiments with episomes have demonstrated<br />

that methylation of as few as 7% of the CpG<br />

sites can induce quiescence of a gene locus (40, 41).<br />

Why individual skeletal muscle fibers showed a relaxed<br />

transgene expression control of the NTA cassettes in<br />

single-transgenic animals remains to be investigated.<br />

We cannot rule out that simple gene dose effects led to<br />

the widespread expression in double-transgenic animals.<br />

However, we assume that unknown molecular<br />

mechanisms during critical time windows have resulted<br />

in the observed widespread expression in double-transgenic<br />

pigs. It is well known that genomic methylation<br />

and histone modifications undergo major changes during<br />

development (42).<br />

Our data indicate that de novo methylation may<br />

interfere with proper expression of tetracycline sensitive<br />

promoters with significant implications for the<br />

generation of tet-transgenic animals as most of the<br />

employed promoter sequences are based on the CMV<br />

minimal promoter. Investigations of DMNT1 over-expressing<br />

cell lines favor the concept that the sequence<br />

context might play a crucial role in the methylation<br />

sensitivity of CpG islands (43). This might explain the<br />

high degree of variability in activator-dependent target<br />

gene expression requiring screening for appropriately<br />

responding mouse lines, as well as the here observed<br />

asymmetric expression in pigs. Along this line it is<br />

interesting that CpG depleted CMV promoter constructs<br />

gave rise to long term expression in vivo (44).<br />

Effective conditional transgene expression could be<br />

E8 Vol. 20 June 2006 The FASEB Journal<br />

KUES ET AL.


demonstrated in transgenic pigs and in white blood<br />

cells. Feeding of transgenic pigs with a daily dose of 3.3<br />

mg doxycycline per kg body weight terminated transgene<br />

transcription in vivo. Re-expression of human<br />

RCAs after termination of doxycycline application was<br />

resumed by autoactivation after 8 wk. Whether this<br />

delayed re-expression is due to a sluggish clearance of<br />

doxycycline (15) or the architecture of the autoregulative<br />

expression cassette remains to be determined. At<br />

present we do not have evidence for specific metabolic<br />

features or a tetracycline reservoir in the pig that could<br />

explain the long interval from termination to re-expression<br />

of the cassette. Measurements of doxycyline levels<br />

in porcine plasma indicated a rapid decline after the<br />

application. Analysis of the fodder for doxycycline<br />

contaminations did not reveal any doxycycline residues.<br />

Overall, these data show that the regulation of the<br />

RCA-NTA cassettes is tight, once the expression is turned<br />

off by doxycycline. Probably, time consuming stochastic<br />

events re-express RCA-NTA cassettes and induce autoactivation.<br />

However, in cell culture RCAs were re-expressed<br />

within 6–10 d after withdrawal of doxycycline.<br />

In conclusion, we have generated transgenic pigs<br />

bearing a bicistronic tetracycline-responsive cassette<br />

and demonstrated tightly controlled expression in a<br />

large animal model for the first time. The usage of the<br />

autoregulative bicistronic cassette supports efficient<br />

single step transduction, which is of utmost importance<br />

in view of the long generation intervals in large domestic<br />

animals. Expression of RCAs and transactivator did<br />

not compromise health status and fertility of transgenic<br />

pigs. Design of the next generation of expression<br />

cassettes will take into account potential methylation<br />

prone sequences and should thus be compatible with<br />

true ubiquitous transgene expression already in the<br />

F0-generation. The use of optimized autoregulative<br />

constructs in combination with improved gene transfer<br />

techniques, such as somatic nuclear transfer or lentiviral<br />

vectors will pave the way toward precisely controlled<br />

transgene expression in farm animals for biomedical<br />

and agricultural application (45).<br />

The critical discussions and comments by J.W. Carnwath are<br />

gratefully acknowledged. The authors thank K-G. Hadeler, H.H.<br />

Döpke (Mariensee) and L. Schöberlein (Leipzig) for expert<br />

help with the muscle biopsy device and A. Saalmüller (Wien) for<br />

gift of the STE cell line. The excellent technical assistance of A.<br />

Kanwischer (Hannover) and W. Mysegades, the introduction<br />

into bisulfite sequencing by C. Gebert, as well as the expert<br />

animal caretaking by H. Hornbostel and T. Peker are kindly<br />

acknowledged. This work was supported by grants of the DFG<br />

(SFB 265, Forschergruppe 535) and BMBF to H.N. Support of<br />

the Kuratorium für Dialyze und Nierentransplantation to K.W. is<br />

acknowledged.<br />

Competing financial interests: The authors declare that they<br />

have no competing financial interests in relation to this paper.<br />

REFERENCES<br />

1. Mayo, K.E., Warren, R., and Palmiter, R. D. (1982) The mouse<br />

metallothionein-I gene is transcriptionally regulated by cad-<br />

INDUCIBLE TRANSGENIC EXPRESSION IN A LARGE ANIMAL MODEL<br />

mium following transfection into human or mouse cells. Cell 29,<br />

99–108<br />

2. Lee, F., Mulligan, R., Berg, P., and Ringold, G. (1981) Glucocorticoids<br />

regulate expression of dihydrofolate reductase cDNA in<br />

mouse mammary tumour virus chimaeric plasmids. Nature 294,<br />

228–232<br />

3. Miller, K.F., Bolt, D.J., Pursel, V.G., Hammer, R.E., Pinkert, C.A.,<br />

Palmiter, R. D., and Brinster, R. L. (1989) Expression of human<br />

or bovine growth hormone gene with a mouse metallothionein-1<br />

promoter in transgenic swine alters the secretion of<br />

porcine growth hormone and insulin-like growth factor-I. J.<br />

Endocrinol. 120, 481–488<br />

4. Pursel, V.G., Pinkert, C.A., Miller, K.F., Bolt, D.J., Campbell,<br />

R.G., Palmiter, R.D., Brinster, R. L., and Hammer, R. E. (1989)<br />

Genetic engineering of livestock. Science 244, 1281–1288<br />

5. Lewandowski, M. (2001) Conditional control of gene expression<br />

in the mouse. Nat. Rev. 2, 743–755<br />

6. Corbel, S. C., and Rossi, F. M. (2002) Latest developments and<br />

in vivo use of the Tet system: ex vivo and in vivo delivery of<br />

tetracycline-regulated genes. Curr. Opin. Biotechnol. 13, 448–452<br />

7. Furth, P.A., St Onge, L., Boger, H., Gruss, P., Gossen, M.,<br />

Kistner, A., Bujard, H., and Hennighausen, L. (1994) Temporal<br />

control of gene expression in transgenic mice by a<br />

tetracycline-responsive promoter. Proc. Natl. Acad. Sci. U. S. A.<br />

91, 9302–9306<br />

8. Kistner, A., Gossen, M., Zimmermann, F., Jerecic, J., Ullmer, C.,<br />

Lubbert, H., and Bujard, H. (1996) Doxycycline-mediated quantitative<br />

and tissue-specific control of gene expression in transgenic<br />

mice. Proc. Natl. Acad. Sci. U. S. A. 93, 10933–10938<br />

9. Zhu, Z., Ma, B., Homer, R.J., Zheng, T., and Elias, J. A. (2001)<br />

Use of the tetracycline controlled transcriptional silencer (tTS)<br />

to eliminate transgene leak in inducible overexpression of<br />

transgenic mice. J. Biol. Chem. 276, 25222–25229<br />

10. Sakai, N., Thome, J., Newton, S.S., Chen, J., Kelz, M.B., Steffen,<br />

C., Nestler, E. J., and Duman, R. S. (2002) Inducible and brain<br />

region-specific CREB transgenic mice. Mol. Pharmacol. 61, 1453–<br />

1464<br />

11. Lindeberg, J., Mattsson, R., and Ebendal, T. (2002) Timing the<br />

doxycycline yields different patterns of genomic recombination<br />

in brain neurons with a new inducible Cre transgene. J. Neurosci.<br />

Res 68, 248–253<br />

12. Smith-Arica, J.R., Morelli, A.E., Larregina, A.T., Smith, J., Lowenstein,<br />

P. R., and Castro, M. G. (2000) Cell-type-specific and<br />

regulatable transgenesis in the adult brain: adenovirus-encoded<br />

combined transcriptional targeting and inducible transgene<br />

expression. Mol. Ther. 2, 579–587<br />

13. Gunther, E.J., Belka, G.K., Wertheim, G.B., Wang, J., Hartman,<br />

J.L., Boxer, R. B., and Chodosh, L. A. (2002) A novel doxycycline-inducible<br />

system for the transgenic analysis of mammary<br />

gland biology. FASEB J. 16, 283–292<br />

14. Robertson, A., Perea, J., Tolmachova, T., Thomas, P. K., and<br />

Huxley, C. (2002) Effects of mouse strain, position of integration<br />

and tetracycline analogue on the tetracycline conditional<br />

system in transgenic mice. Gene 282, 65–74<br />

15. Stepensky, D., Kleinberg, L., and Hoffman, A. (2003) Bone as<br />

an effect compartment. Clin. Pharmacokinet. 42, 863–881<br />

16. Böger, H., and Gruss, P. (1999) Functional determinants for the<br />

tetracycline- dependent transactivator tTA in transgenic mouse<br />

embryos. Mech. Dev. 83, 141–153<br />

17. Niemann, H., and Kues, W. A. (2003) Application of transgenesis<br />

in livestock for agriculture and biomedicine. Anim. Reprod.<br />

Sci. 79, 291–317<br />

18. Schultze, N., Burki, Y., Lang, Y., Certa, U., and Bluethmann, H.<br />

(1996) Efficient control of gene expression by single step<br />

integration of the tetracycline system in transgenic mice. Nat.<br />

Biotechnol. 14, 499–503<br />

19. Kringstein, A.M., Rossi, F.M., Hofmann, A., and Blau, H. M.<br />

(1998) Graded transcriptional response to different concentrations<br />

of a single transactivator. Proc. Natl. Acad. Sci. U. S. A. 95,<br />

13670–13675<br />

20. Shockett, P., Difilippantonio, M., Hellman, N., and Schatz, D. G.<br />

(1995) A modified tetracycline-regulated system provides autoregulatory,<br />

inducible gene expression in cultured cells and<br />

transgenic animals. Proc. Natl. Acad. Sci. U. S. A. 92, 6522–6526<br />

21. Hofmann, A., Nolan, G. P., and Blau, H. M. (1996). Rapid<br />

retroviral delivery of tetracycline-inducible genes in a single<br />

E9


autoregulatory cassette. Proc. Natl. Acad. Sci. U. S. A. 93, 5185–<br />

5190<br />

22. Kühnel, F., Fritsch, C., Krause, S., Mundt, B., Wirth, T., Paul, Y.,<br />

Malek, N.P., Zender, L., Manns, M. P., and Kubicka, S. (2004)<br />

Doxycycline regulation in a single retroviral vector by an autoregulatory<br />

loop facilitates controlled gene expression in liver<br />

cells. Nucleic Acids Res. 32, e30<br />

23. Markusic, D., Oude-Elferink, R., Das, A.T., Berkhout, B., and<br />

Seppen, J. (2005) Comparison of single regulated lentiviral<br />

vectors with rtTA expression driven by an auto regulatory loop<br />

or a constitutive promoter. Nucleic Acids Res. 33, e63<br />

24. Unsinger, J., Kröger, A., Hauser, H., and Wirth, D. (2001)<br />

Retroviral vectors for the transduction of autoregulated bidirectional<br />

expression cassettes. Mol. Ther. 4, 484–489<br />

25. Unsinger, J., Lindenmaier, W., May, T., Hauser, H., and Wirth,<br />

D. (2004) Stable and strictly controlled expression of LTRflanked<br />

autoregulated expression cassettes upon adenoviral<br />

transfer. Biochem. Biophys. Res. Commun. 319, 379–387<br />

26. May, T., Hauser, H., and Wirth, D. (2004) Transcriptional<br />

control of SV40 T-antigen expression allows a complete reversion<br />

of immortalization. Nucleic Acids Res. 32, 5529–5538<br />

27. Spitzer, D., Hauser, H., and Wirth, D. (1999) Complementprotected<br />

amphotrophic retroviruses from murine packaging<br />

cells. Hum. Gene Ther. 10, 1893–1902<br />

28. Gossen, M., and Bujard, H. (1992) Tight control of gene<br />

expression in mammalian cells by tetracycline-responsive promoters.<br />

Proc. Natl. Acad. Sci. U. S. A. 89, 5547–5551<br />

29. Dirks, W., Schaper, F., Kirchhoff, S., Morelle, C., and Hauser, H.<br />

(1994) A multifunctional vector family for gene expression in<br />

mammalian cells. Gene 149, 387–388<br />

30. Niemann, H., Verhoeyen, E., Wonigeit, K., Lorenz, R., Hecker, J.,<br />

Schwinzer, R., Hauser, H., Kues, W.A., Halter, R., Lemme, E., et al.<br />

(2001) CMV early promoter induced expression of hCD59 in<br />

porcine organs provides protection against hyperacute rejection.<br />

Transplantation 72, 1898–1906<br />

31. Kues, W.A., Brenner, H.R., Sakmann, B., and Witzemann, V.<br />

(1995) Local neurotrophic repression of gene transcripts encoding<br />

fetal AChRs at rat neuromusclular synapses. J. Cell Biol.<br />

130, 949–957<br />

32. Kues, W.A., Petersen, B., Mysegades, W., Carnwath, J.W., and<br />

Niemann. H. (2005) Isolation of stem cell-like cells from murine<br />

and porcine somatic tissue. Biol. Reprod. 72, 1020–1028<br />

33. Schöberlein, L. (1976) Die Schussbiopsie - eine neue Methode <strong>zur</strong><br />

Entnahme von Muskelproben. Monatsh. Vet -Med. Jena 31, 457–460<br />

34. Hajkova, P., el-Maarri, O., Engemann, S., Oswald, J., Olek, A.,<br />

and Walter, J. (2002) DNA-methylation analysis by the bisulfiteassisted<br />

genomic sequencing method. Methods Mol. Biol. 200,<br />

143–154<br />

35. White, D. (1996) Alteration of complement activity: a strategy<br />

for xenotransplantation. Trends Biotechnol. 14, 3–5<br />

36. Baron, U., Freundlieb, S., Gossen, M., and Bujard, H. (1995)<br />

Coregulation of two gene activities by tetracycline via a bidirectional<br />

promoter. Nuc. Acid Res. 17, 3605–3606<br />

37. Jones, P. A. (1999) The DNA methylation paradox. Trends Genet.<br />

15, 34–37<br />

38. Antequera, F. (2003) Structure, function and evolution of CpG<br />

island promoters. Cell Mol. Life Sci. 60, 1647–1658<br />

39. Jones, P. A., and Baylin, S. B. (2002) The fundamental role of<br />

epigenetic events in cancer. Nat. Rev. Genet. 3, 415–428<br />

40. Robertson, K.D., Hayward, S.D., Ling, P.D., Samid, D., and<br />

Ambinder, R. F. (1995) Transcriptional activation of the Epstein-Barr<br />

virus latency C promoter after 5-azacytidine treatment:<br />

evidence that demethylation of a single CpG site is<br />

crucial. Mol. Cell. Biol. 15, 6150–6159<br />

41. Hsieh, C. L. (1994) Dependence of transcriptional repression<br />

on CpG methylation density. Mol. Cell. Biol. 14, 5487–5494<br />

42. Reik, W., and Walter, J. (2001) Genomic imprinting: Parental<br />

influence on the genome. Nat. Rev. Genet. 2, 21–32<br />

43. Feltus, F.A., Lee, E.K., Costello, J.F., Plass, C., and Vertino, P. M.<br />

(2003) Predicting aberrant CpG island methylation. Proc. Natl.<br />

Acad. Sci. U. S. A. 100, 12253–12258<br />

44. Yew, N.S., Zhao, H., Przybylska, M., Wu, I.H., Tousignant, J.D.,<br />

Scheule, R. K., and Cheng, S. H. (2002) CpG depleted plasmid<br />

DNA vectors with enhanced safety and long-term gene expression<br />

in vivo. Mol. Ther. 5, 731–738<br />

45. Kues, W. A., and Niemann, H. (2004) The contribution of farm<br />

animals to human health. Trends Biotechnol. 22, 286–294<br />

Received for publication November 15, 2005.<br />

Accepted for publication February 14, 2006.<br />

E10 Vol. 20 June 2006 The FASEB Journal<br />

KUES ET AL.


Xenotransplantation 2006: 13: 345–356<br />

Printed in Singapore. All rights reserved<br />

doi: 10.1111/j.1399-3089.2006.00317.x<br />

Health status of transgenic pigs expressing<br />

the human complement regulatory protein<br />

CD59<br />

Deppenmeier S, Bock O, Mengel M, Niemann H, Kues W, Lemme E,<br />

Wirth D, Wonigeit K, Kreipe H. Health status of transgenic pigs<br />

expressing the human complement regulatory protein CD59.<br />

Xenotransplantation 2006; 13: 345–356. Ó Blackwell Munksgaard, 2006<br />

Abstract: Background: Microinjection of foreign DNA into pronuclei<br />

of zygotes has been the method of choice for the production of transgenic<br />

domestic animals. Following microinjection the transgene is randomly<br />

integrated into the host genome which can be associated with<br />

insertional mutagenesis and unwanted pathological side effects.<br />

Methods: Here, we evaluated the health status of pigs transgenic for the<br />

human regulator of complement activation (RCA) CD59 and conducted<br />

a complete pathomorphological examination on 19 RCA transgenic pigs<br />

at 1 to 32 months of age from nine transgenic lines. Nine wild-type<br />

animals served as controls. Expression levels of human complement<br />

regulator CD59 (hCD59) mRNA were measured by RT-PCR and distribution<br />

of hCD59 protein was determined by immunohistochemistry.<br />

Results: Al<strong>bei</strong>t variable transgene expression levels, no specific pathomorphologic<br />

phenotype associated with the presence of the transgene in<br />

all analyzed pig lines could be detected.<br />

Conclusions: Transgenic expression of this human RCA gene construct<br />

is not correlated with a specific pathological phenotype in pigs. This is<br />

crucial for the application of the technology and the use of transgenic<br />

pigs for biomedical and agricultural applications.<br />

Introduction<br />

The critical shortage of human donor organs has<br />

prompted considerable efforts in establishing<br />

porcine to human xenotransplantation. The first<br />

Copyright Ó Blackwell Munksgaard 2006<br />

Stefanie Deppenmeier, 1,2 Oliver<br />

Bock, 2 Michael Mengel, 2 Heiner<br />

Niemann, 3 Wilfried Kues, 3 Erika<br />

Lemme, 3 Dagmar Wirth, 4 Kurt<br />

Wonigeit 5 and Hans Kreipe 2<br />

1 Department of Pathology, School of Veterinary<br />

Medicine Hannover, Hannover, 2 Department of<br />

Pathology, Hannover Medical School, Hannover,<br />

3 Department of Biotechnology, Institute of Animal<br />

Breeding, Mariensee (FAL), Neustadt, 4 Department<br />

of Gene Regulation and Differentiation, German<br />

Research Center for Biotechnology, Braunschweig,<br />

5 Klinik für Viszeral- und Transplantationschirurgie,<br />

Hannover Medical School, Hannover, Germany<br />

Key words: expression pattern – hCD59-transgenic<br />

pigs – health status – phenotype –<br />

xenotransplantation<br />

Abbreviations: CMV: cytomegalovirus; DAF: decay<br />

accelerating factor; DEPC: diethylpyrocarbonate;<br />

EF1a: elongation factor 1 alpha; F1, F2: generation<br />

number of transgenic offspring; FFPE: formalinefixed<br />

paraffine-embedded; hCD59: human<br />

complement regulator CD59; Ln./Lnn.: lymph node<br />

(lymphonodus)/lymph nodes (lymphonodi); MCP:<br />

membrane cofactor protein; PCR: polymerase chain<br />

reaction; PCV: porcine circovirus; PRRSV: porcine<br />

respiratory and reproduction syndrome virus;<br />

RT-PCR: reverse transcriptase-polymerase chain<br />

reaction; SPF: specific pathogen free; TNF: tumor<br />

necrosis factor.<br />

Address reprint requests to Professor Oliver Bock<br />

MD, Institute of Pathology, Medizinische Hochschule<br />

Hannover, Carl-Neuberg-Strasse 1, 30625 Hannover,<br />

Germany (E-mail: bock.oliver@mh-hannover.de)<br />

or Professor Heiner Niemann, Dept. Biotechnology,<br />

Institute of Animal Breeding, (FAL) Mariensee, 31535<br />

Neustadt, Germany (E-mail: niemann@tzv.fal.de)<br />

Received 25 September 2005;<br />

Accepted 9 April 2006<br />

XENOTRANSPLANTATION<br />

immunological barrier in xenotransplantation is the<br />

hyperacute rejection response (HAR), which is<br />

mediated by preformed xenoreactive antibodies<br />

that are directed against the a-galactosyl-epitope<br />

[1,2]. These antibodies activate the complement<br />

345


Deppenmeier et al.<br />

cascade of the organ recipient which results in a<br />

rapid and irreversible destruction of the xenogenic<br />

donor organ [3,4]. Transgenic overexpression of<br />

human complement regulators such as CD59, CD55<br />

[decay accelerating factor (DAF)] or CD46 [membrane<br />

cofactor protein (MCP)] has been shown to<br />

inhibit the complement cascade and reliably prevents<br />

HAR [see 5–8]. The complement regulatory<br />

protein human complement regulator CD59<br />

(hCD59) inhibits formation of the membrane attack<br />

complex at the end of the complement cascade [6–8].<br />

Microinjection of foreign DNA into pronuclei of<br />

zygotes has been the method of choice for the<br />

production of transgenic domestic animals. However,<br />

the technology is associated with a random<br />

integration of the transgene and frequently an<br />

unpredictable expression pattern in the transgenic<br />

animals [9]. Due to the random integration and<br />

variable expression patterns, insertional mutagenesis<br />

and unwanted pathological side effects cannot<br />

be ruled out and concerns have been raised about<br />

animal welfare of animals resulting from this<br />

approach [10,11]. Ectopic transgene expression<br />

has been detected in some cases, but without any<br />

obvious pathology [12,13]. Pathomorphological<br />

phenotypes have been shown in transgenic and<br />

non-transgenic offspring derived from microinjection,<br />

in vitro fertilization and/or somatic cloning.<br />

Overexpression of various growth hormone (GH)<br />

constructs in transgenic pigs or sheep was associated<br />

with a specific pathological phenotype, which<br />

was avoided by improved GH constructs [14,15].<br />

The large offspring syndrome [16], and specific<br />

cardiopulmonary and placental lesions [17–19] and<br />

lesions in the skeletal and muscle system [20,21]<br />

have been observed as long-term consequences of<br />

in vitro handling of bovine and ovine embryos.<br />

However, a systematic analysis of potential pathological<br />

side effects putatively associated with the<br />

random integration and expression of a specific<br />

transgene in transgenic domestic animals has not<br />

yet been reported.<br />

Here, we evaluated the health status of pigs<br />

transgenic for the human complement regulator<br />

protein CD59 and conducted a complete pathomorphological<br />

examination on transgenic and wildtype<br />

control animals. The transgenic pigs used in the<br />

present study were hemizygous F1–F2 offspring<br />

from nine transgenic lines. The founder animals had<br />

been produced by pronuclear microinjection of<br />

human regulator of complement activation (RCA)<br />

constructs into porcine zygotes [22].<br />

The main goals of the present study were to<br />

determine (i) the health status of hCD59-transgenic<br />

pigs by a detailed pathomorphological analysis,<br />

and (ii) if transgenesis is correlated with specific<br />

346<br />

pathological changes in F1–F2 transgenic offspring.<br />

In addition, the expression pattern of both<br />

hCD59 mRNA and hCD59 protein in a large panel<br />

of organs and tissues was determined.<br />

Materials and methods<br />

Animals<br />

Nineteen German Landrace F1 and F2 hCD59transgenic<br />

pigs (TG) from nine transgenic lines and<br />

nine non-transgenic control pigs (WT), between 1<br />

and 32 months of age were slaughtered at the<br />

Institute for Animal Breeding, Mariensee. Both<br />

groups were kept under identical specific pathogen<br />

free (SPF) conditions. All animals were treated<br />

according to institutional guidelines. The sex, age<br />

and breeding line of the animals examined are<br />

shown in Table 1. We selected a representative<br />

number of animals from a total of nine different<br />

transgenic lines with a rather broad range in age<br />

for this study. This should render the results<br />

representative for other pig populations derived<br />

from this transgenic approach.<br />

The vast majority of the animals analyzed in this<br />

study were from cytomegalovirus promoter (CMV)hCD59<br />

transgenic lines; in addition one animal with<br />

a CMV-DAF transgene (from line 13) was analyzed.<br />

The transgene expression patterns for line 5, carrying<br />

a construct in which the CMV promoter (PCMV) drives expression of human CD59, have been<br />

characterized previously [22]. Recently, transgenic<br />

expression patterns of five lines (13, 17, 19, 20 and<br />

21), carrying a tetracycline responsive promoter<br />

(PtTA) have been described [23]. The expression<br />

pattern from two lines (8 and 10) with a CMV driven<br />

construct had not been published earlier. All<br />

animals did not show any obvious health problems.<br />

One animal (TG26) was from a double transgenic<br />

background (hCD59 · hCD55).<br />

Pathomorphological analysis<br />

All pigs were killed and organs and tissues were<br />

evaluated macroscopically for pathomorphological<br />

changes. Lung and Lymphonodus pulmonalis were<br />

tested by PCR for porcine circovirus type 2 (PCV<br />

2) and for porcine respiratory and reproductive<br />

syndrome virus (PRRSV) by a certified laboratory<br />

(Laboklin, Bad Kissingen, Germany). These viruses<br />

were included as they are emerging pathogens<br />

that are now frequently found in domestic pig<br />

populations. Histopathological examination<br />

showed evidence (intracytoplasmic, basophilic<br />

inclusion bodies in lymph nodes) for a PCV 2<br />

infection in the animals.


Table 1. Age, sex and transgenic construct in pigs analyzed for pathologies<br />

The following organs and tissues were fixed in<br />

10% buffered formalin: skin (inner tight), skeletal<br />

muscle (M. gracilis), heart (left ventricle, right<br />

ventricle, septum), lung (Lobus cranialis, Lobus<br />

caudalis), spleen, liver, kidney, urinary bladder,<br />

pancreas, brain (cerebellum, cerebrum), thyroid<br />

gland, lymphatic nodes (Ln. pulmonalis, Ln. cervicalis<br />

superficialis, Lnn. ileofemorales, Lnn. mesenteriales),<br />

esophagus, stomach, duodenum,<br />

jejeunum, ileum, colon, and aorta. If the organs<br />

and tissues did not show macroscopic lesions,<br />

samples were taken from representative sites for<br />

histopathological examination. If macroscopic lesions<br />

were visible, the samples were taken from the<br />

site of the lesion. After 24 h of fixation, samples<br />

were paraffin-embedded, stained with hematoxylin–eosin<br />

and histopathologically evaluated.<br />

Determination of integration of the hCD59 construct<br />

DNA extraction was performed as described previously<br />

[24]. In brief, eight to 10 formalin-fixed,<br />

paraffin-embedded (FFPE) samples derived from<br />

heart, lung, spleen, liver, kidney, pancreas, brain,<br />

skeletal muscle, and esophagus were employed for<br />

genomic DNA isolation. The isolated DNA was<br />

then adjusted to a final concentration of 50 ng/ll.<br />

The human complement regulatory protein hCD59<br />

Animal no. Age (months) Sex Construct Breeding line Litter size Stillborn G TG m TG f<br />

WT1 8 Male – –<br />

WT2 12 Male – –<br />

WT3 15 Male – –<br />

WT4 4 Male – –<br />

WT5 18 Female – –<br />

WT6 17 Female – –<br />

WT7 21 Male – –<br />

WT8 28 Male – –<br />

WT9 1 Male – –<br />

TG6 18 Male P CMV CD59 5 8 (4m, 4f) – F1 4 4<br />

TG7 4 Female P CMV CD59 5 11 (3m, 8f) 1 F2 2 6<br />

TG8 4 Female P CMV CD59 5 Sib. of TG7 1 F2 2 6<br />

TG1 21 Female P CMV CD59-GFP 8 10 (6m, 4f) 1 F1 1 2<br />

TG4 7 Male P CMV CD59-GFP 8 13 (7m, 6f) 2 F1 2 2<br />

TG15 32 Female P CMV CD59-GFP 8 10 (4m, 6f) – F1 3 2<br />

TG20 25 Male P CMV CD59-GFP 8 Sib. of TG4 2 F1 2 2<br />

TG2 17 Female P CMV CD59-GFP 10 10 (3m, 7f) – F1 2 2<br />

TG5 8 Female PCMV CD59-GFP 10 11 (5m, 6f) – F1 1 4<br />

TG23 24 Male P tTA DAF-TA 13<br />

TG13 7 Male P tTA CD59-TA 17<br />

TG14 7 Male P tTA CD59-TA 19<br />

TG19 15 Male P tTA CD59-TA 19<br />

TG18 11 Male P tTA CD59-TA 20<br />

TG21 16 Female P tTA CD59-TA 20<br />

TG25 12 Female P tTA CD59-TA 20<br />

TG27 1 Male P tTA CD59-TA 20<br />

TG12 7 Male P tTA CD59-TA 21<br />

TG26 1 Female P tTA CD59-TA 15 · 20<br />

P tTA DAF-TA<br />

GFP, enhanced green fluorescent protein; Sib, Sibling; G, generation; TG m, male transgenic piglets; TG f, female transgenic piglets; m, male; f, female.<br />

Primers for amplification of hCD59 DNA were<br />

based on the sequence of the hCD59 construct as<br />

follows:<br />

hCD59 forward: 5¢-GGTCATAGCCTGCAGT-<br />

GCTACA-3¢,<br />

hCD59 reverse: 5¢-AAATCAGATGAACAAT-<br />

TGACGGC-3¢.<br />

Primers for hCD59 generated a 77 bp product.<br />

The sequence of tumor necrosis factor (TNF)<br />

was chosen as control gene to prove successful<br />

extraction of genomic DNA. The following TNFspecific<br />

primers were employed (GenBank X54001):<br />

TNF forward: 5¢-ATAGTATGAGGAGGAC-<br />

TCGCTGAGC-3¢<br />

TNF reverse: 5¢-ACATCCCATCTGTCCATG-<br />

AGGTT-3¢.<br />

The TNF primers generated a 94 bp product.<br />

PCR amplification was performed in a Gene-<br />

Amp Ò PCR System 2700 (Applied Biosystems,<br />

Darmstadt, Germany) with a final reaction mixture<br />

of 25 ll containing 250 nm of each primer, 0.5 units<br />

of Platinum Taq Polymerase (Invitrogen, Karlsruhe,<br />

Germany), 200 lm each of dATP, dCTP,<br />

dTTP, and dGTP in 1x Platinum Taq reaction<br />

buffer with 1.5 mm magnesium chloride and 5.0 ll<br />

347


Deppenmeier et al.<br />

DNA. The reaction mixture was preheated at 95°C<br />

for 5 min, followed by 40 cycles at 95, 60, and 72°C,<br />

30 s each. Final extension was performed for 5 min<br />

at 72°C.<br />

DNA derived from fibroblasts transfected with<br />

the hCD59 construct served as positive controls.<br />

DNA derived from organs and tissues of nontransgenic<br />

control pigs and mock PCR (without<br />

template) served as negative controls.<br />

Analysis of hCD59 mRNA expression<br />

Formalin-fixed, paraffin-embedded samples derived<br />

from heart, lung, spleen, liver, kidney, pancreas,<br />

brain, skeletal muscle, and esophagus of the<br />

19 RCA-transgenic pigs were analyzed. For RNA<br />

extraction from FFPE samples, a one-step protocol<br />

was used as described previously [25]. Briefly, eight<br />

to 10 sections ( 15 lm each) were cut from the<br />

paraffin block with a microtome and placed directly<br />

into a 1.5 ml screw cap-tube; 750 ll digestion<br />

solution (4 m guanidinium isothiocyanate/0.75 m<br />

sodium citrate/10% sarcosyl), 5.25 ll 0.1 b-mercaptoethanol<br />

(Merck, Darmstadt, Germany) and<br />

300 ll proteinase K solution (20 mg/ml in H2O;<br />

Merck) were added, and the tubes were placed in a<br />

thermoshaker for overnight incubation (55°C,<br />

1400 g). The following day, the samples were<br />

centrifuged at 4°C for 5 min at 14 000 g, and 1 ml<br />

of each digested sample was transferred into a<br />

2.0 ml tube. After addition of 100 ll 3m sodium<br />

acetate (pH 5.2), 630 ll water-saturated phenol (pH<br />

4.5 to 5.5), and 270 ll chloroform, the sample was<br />

mixed thoroughly and stored on ice for 15 min.<br />

After centrifugation at 4°C for 20 to 25 min at<br />

14 000 g, the upper aqueous phase (approximately<br />

1 ml) was pipetted into a fresh 2.0 ml tube; 1 ll<br />

glycogen (20 mg/ml; Roche Molecular Diagnostics,<br />

Mannheim, Germany) as an RNA carrier and 1 ml<br />

2-propanol were added to the ribonucleic acidcontaining<br />

aqueous phase. After mixing, the samples<br />

were stored for at least 24 h at )20°C. After<br />

precipitation, the samples were centrifuged at 4°C<br />

for 20 min at 14 000 g. The RNA pellet was washed<br />

with 70% ice-cold ethanol, air-dried, and dissolved<br />

in 35 ll precooled DEPC-treated H 2O. The RNA<br />

concentration was then measured spectrophotometrically.<br />

Total RNA (1 lg), pretreated with RNase )<br />

DNase (1 U/lg RNA; RQ1, Promega, Madison,<br />

WI, USA), was then transcribed into complementary<br />

DNA using 250 ng random hexamers (Amersham<br />

Pharmacia, Freiburg, Germany) and 200 U<br />

of SuperScript II Rnase ) Reverse Transcriptase<br />

(Invitrogen) in a volume of 20 ll following the<br />

manufacturer’s protocol. Negative controls were<br />

348<br />

performed by adding water instead of Reverse<br />

Transcriptase.<br />

Primers used for the hCD59 RT-PCR were<br />

identical to the hCD59 DNA PCR (see above). The<br />

housekeeping gene elongation factor 1 (EF1) a was<br />

chosen as a positive control [26]. The following<br />

sequences were used (GenBank XM203909):<br />

EF1 a forward: 5¢-CAAAAAY(C/T)GAC-<br />

CCACCAATGG-3¢<br />

EF1 a reverse: 5¢-GGCCTGGATGGTTCAGG-<br />

ATA-3¢.<br />

The EF1 a RT-PCR generated a 68 bp product.<br />

RT-PCR conditions were similar to the hCD59<br />

PCR as described above.<br />

RNA extracted from human kidney samples<br />

served as positive controls. RNA derived from<br />

various organs and tissues of non-transgenic control<br />

pigs, of a hCD55-transgenic pig, and samples<br />

without reverse transcriptase ()RT) served as<br />

negative controls.<br />

Immunohistochemistry for determination of hCD59 protein<br />

expression<br />

Immunohistochemistry was performed using a<br />

tissue microarray technology as previously described<br />

[27] employing the same FFPE samples as<br />

for molecular analysis. For immunohistochemical<br />

staining the tyramine amplification technique was<br />

used as described [28]. Slides were deparaffinized,<br />

dehydrated and rinsed in deionized water. Antigen<br />

retrieval was carried out by incubation with<br />

protease XIV (21.4 mg/50 ml aqua dest; Sigma-<br />

Aldrich, Mu¨nchen, Germany) for 10 min. Endogenous<br />

peroxidase blocking was performed with 3%<br />

H 2O 2 (pH 7.5). Blocking of unspecific binding was<br />

achieved by incubating with 0.5% casein for<br />

10 min. To avoid unspecific binding to endogenous<br />

biotin an incubation with avidin (Avidin/Biotin<br />

blocking kit; Vector, Burlingame, CA, USA) for<br />

10 min each was performed. To detect expression<br />

of hCD59, a monoclonal mouse-anti-human<br />

CD59-antibody [clone p282 (H19); BD PharMingen,<br />

Heidelberg, Germany], diluted 1 : 200 was<br />

applied and incubated overnight at 4°C. As secondary<br />

antibody, biotinylated goat-anti-mouse-antibody<br />

(Zytomed, Berlin, Germany) was applied in<br />

dilution 1 : 900 for 30 min, followed by horseradish<br />

peroxidase complex (Zytomed) diluted 1 : 500<br />

was applied for 30 min. To amplify the reaction,<br />

biotinylated tyramine with 0.003% H 2O 2 was<br />

applied for 10 min. Biotinylation of tyramine was<br />

performed as reported previously [27]. Incubation<br />

with avidin-biotinylated alkaline phosphatase (Zytomed)<br />

diluted 1 : 300 was performed for 30 min.


For color development, slides were incubated with<br />

the chromogen fast red (1 mg fast red salt/1 ml fast<br />

red buffer; Sigma-Aldrich) for 8 min, counterstained<br />

with hemalum (dilution 1 : 3; Merck) and<br />

thoroughly washed with buffer (0.1 m Tris–HCl,<br />

pH 7.6/0.15 m NaCl/0.05% Tween 20) after each<br />

incubation step. The staining pattern was evaluated<br />

and the percentage of positively labeled cells<br />

and their distribution pattern was assessed<br />

() ¼ negative, + ¼ weak staining in up to 30%<br />

of the cells, ++ ¼ moderate staining of 30% to<br />

60% of the cells, +++ ¼ strong staining in<br />

>60% of the cells). Human tonsillar tissue served<br />

as positive control, whereas tissue derived from<br />

non-transgenic pigs served as negative controls.<br />

To ensure that the hCD59 staining pattern as<br />

determined by screening of tissue arrays represented<br />

the true overall staining pattern in transgenic<br />

animals, whole tissue sections were stained and<br />

evaluated accordingly in positive cases.<br />

Results<br />

Pathomorphological evaluation of transgenic pigs lines<br />

expressing hCD59<br />

Transgenic animals as well as non-transgenic<br />

control pigs only rarely showed any pathological<br />

lesions. Only in individual animals non-specific<br />

pathologies were detected (Fig. 1). One non-transgenic<br />

animal (WT3) had a bilateral, abdominal<br />

cryptorchism. Another non-transgenic animal<br />

(WT7) showed a moderate, focal, chronic, purulent<br />

bronchopneumonia with multifocal minerali-<br />

Fig. 1. Examples of histopathological<br />

lesions in transgenic and non-transgenic<br />

animals. (A) Lnn. ileofemorales (WT6):<br />

Deposition of siderin-containing macrophages<br />

(arrow) in the lymph node; (B)<br />

Skin (WT7): mild, multifocal, perivascular<br />

infiltration with lymphocytes, plasma cells<br />

and single neutrophils (arrow); (C) Lung<br />

(WT7): moderate, focal, chronic, purulent<br />

bronchopneumonia with multifocal mineralization<br />

(arrow); (D) Colon (TG7):<br />

Balantidium coli (arrow) in the lumen (bar<br />

100 lm).<br />

The human complement regulatory protein hCD59<br />

zation in the right frontal lobe of the lung. One<br />

transgenic animal (TG4) had a severe, focal,<br />

chronic, ulcerative gastritis in the pars proventricularis<br />

and a severe, multifocal, follicular hyperplasia<br />

of peribronchiolar follicles. One transgenic<br />

animal (TG7) showed a severe, diffuse, chronic,<br />

fibrous epi- and pericarditis and a mild, multifocal,<br />

chronic, fibrous pleuritis in the lung. One transgenic<br />

animal (TG15) had a moderate, focal,<br />

chronic fibrosis of the liver capsule, and transgenic<br />

animal TG20 showed histologically multifocal,<br />

acute, single cell necroses of hepatocytes. One<br />

transgenic animal (TG23) had multifocal, chronic<br />

ossification in the mesenterium. In addition to<br />

these specific lesions in individual animals, several<br />

non-transgenic and transgenic animals showed<br />

mild lesions in a variety of organs (Table 2,<br />

Fig. 1) including mild, perivascular infiltration in<br />

the skin, mild infiltration in the liver, mild, chronic<br />

interstitial nephritis, mild infection with Balantidium<br />

coli and deposition of siderin containing<br />

macrophages in the Lnn. ileofemorales. Nonspecific<br />

pathological signs were found in nontransgenic<br />

control animals and transgenic animals<br />

at the same ratio and to the same extent, and no<br />

specific transgene related pathomorphology was<br />

apparent.<br />

PCV 2 could be detected in 65% of the transgenic<br />

animals and 75% of the non-transgenic<br />

animals. However, despite the existence of intracytoplasmic,<br />

basophilic inclusion bodies in some<br />

lymph nodes, no pathomorphological phenotype<br />

typical for PCV 2-associated diseases, such as postweaning<br />

multisystemic wastening syndrome<br />

349


Deppenmeier et al.<br />

Table 2. Lesions found in several animals<br />

Organ Histopathological lesion Affected animals<br />

(PMWS) or porcine dermatitis and nephropathy<br />

syndrome (PDNS) were detected. PRRSV could<br />

not be detected in any of the transgenic or control<br />

animals.<br />

Expression patterns of hCD59 mRNA in organs and tissues of<br />

transgenic pigs<br />

In general, the previously reported mosaic expression<br />

patterns were confirmed [22,23]. Three animals<br />

(TG18, TG19, TG25) expressed hCD59<br />

mRNA in all organs examined. Twelve animals<br />

(TG1, TG2, TG7, TG8, TG12, TG13, TG14,<br />

TG15, TG20, TG21, TG26, T27) expressed<br />

hCD59 mRNA only in a subset of organs and<br />

tissues. The level of hCD59 mRNA expression<br />

varied in organs and tissues within a single<br />

transgenic animal as well as between different<br />

animals (Table 4, Fig. 2). In four animals (TG4,<br />

TG5, TG6, TG23) no hCD59 mRNA expression<br />

could be detected. The hCD59 mRNA was<br />

consistently demonstrated in positive controls<br />

whereas in negative controls and samples without<br />

reverse transcriptase ()RT) hCD59 mRNA<br />

expression was never determined. EF1 a transcripts<br />

could be detected in all samples indicating<br />

Affected<br />

animals (%)<br />

Skin Mild, multifocal, perivascular infiltration with WT1, WT2, WT6, WT7<br />

44% WT<br />

lymphocytes, plasma cells and single neutrophils TG1, TG2, TG5, TG15, TG18, TG19, TG20, TG21<br />

47% TG<br />

Liver Mild, multifocal, infiltration with lymphocytes, WT3, WT4, WT6, WT7, WT8<br />

56% WT<br />

macrophages and single neutrophils<br />

TG7, TG8, TG12, TG13, TG14, TG15, TG18, TG19, TG20, TG21, TG25 58% TG<br />

Kidney Mild, multifocal, chronic, non-suppurative,<br />

WT7<br />

11% WT<br />

interstitial nephritis<br />

TG13, TG20, TG21<br />

16% TG<br />

Colon Balantidium coli in the lumen without<br />

WT2<br />

11% WT<br />

inflammatory reaction<br />

TG5, TG6, TG7, TG18<br />

21% TG<br />

Spleen Mild to moderate, multifocal, follicular hyperplasia WT3 11% WT<br />

Ln. pulmonalis Multifocal, acute hemorrhages WT6, WT7<br />

TG18, TG19, TG22, TG23, TG26<br />

Follicular hyperplasia WT1, WT2, WT8<br />

TG1, TG2, TG4, TG5, TG13, TG14, TG23, TG26<br />

Edema and sinusoidal histiocytosis WT4, WT6, WT7<br />

TG5, TG7, TG8, TG13, TG15, TG18, TG19, TG21<br />

Lnn. mesenteriales Follicular hyperplasia WT2, WT3, WT4, WT5, WT6, WT8<br />

TG4, TG5, TG7, TG12, TG13, TG15, TG18, TG19, TG20, TG21, TG23, TG25<br />

Edema and sinusoidal histiocytosis WT4, WT7, WT8, WT9<br />

TG4, TG8, TG13, TG14, TG18, TG19, TG21, TG25, TG26<br />

Lnn. ileofemorales Follicular hyperplasia WT2, WT4<br />

TG4, TG12, TG13, TG14, TG20<br />

Edema and sinusoidal histiocytosis WT2, WT3, WT5, WT6, WT7, WT9<br />

TG4, TG5, TG6, TG7, TG8, TG13, TG14, TG26, TG27<br />

Siderin containing macrophages in the sinus WT3, WT5, WT6, WT7, WT9<br />

TG7, TG8, TG15, TG19, TG20, TG23, TG26, TG27<br />

Ln. cervicalis Follicular hyperplasia WT2<br />

cran. superf.<br />

TG4<br />

Edema and sinusoidal histiocytosis WT2, WT4<br />

TG5, TG7, TG8<br />

350<br />

22% WT<br />

26% TG<br />

33% WT<br />

42% TG<br />

33% WT<br />

42% TG<br />

67% WT<br />

63% TG<br />

44% WT<br />

47% TG<br />

22% WT<br />

26% TG<br />

66% WT<br />

47% TG<br />

56% WT<br />

42% TG<br />

11% WT<br />

11% TG<br />

22% WT<br />

16% TG<br />

successful RNA extraction, cDNA synthesis, and<br />

RT-PCR. Samples without )RT were predominantly<br />

negative, but some cases showed weak false<br />

positive amplification products, probably due to an<br />

incomplete DNase digestion.<br />

As expected, hCD59 DNA could be detected in<br />

all animals derived from the group of hCD59transgenic<br />

animals (data not shown), confirming<br />

transmission of the transgene over several generations.<br />

Non-transgenic animals were negative for<br />

hCD59-DNA.<br />

Organ- and tissue-specific expression patterns of hCD59 protein<br />

in transgenic animals<br />

Thirteen of the 19 transgenic pigs showed cellular<br />

expression of the hCD59 protein in various organs<br />

and tissues (Table 3, Fig. 3). From these 13<br />

animals, none showed a consistent expression<br />

pattern of the transgene in all organs. Cellular<br />

hCD59 protein expression could be demonstrated<br />

only in a subset of organs and tissues, and with<br />

remarkable variations between different animals.<br />

In some animals a ‘‘patchy’’ expression pattern was<br />

detected which was characterized by distinct positive<br />

labeling of single cells in heart and kidney,


Table 3. Pattern of hCD59 protein expression in transgenic pigs<br />

Skeletal<br />

muscle Esophagus Duodenum Jejunum Ileum Colon<br />

Urinary<br />

bladder Pancreas Brain<br />

L. Skin Heart Lung Spleen Liver Kidney<br />

TG 6 (hCD59) 5 ++ End. ++ End. ) + End. + End. ) + End. +++ Acin. En. ++ End. ) ++ End. ) +End. ) ++ End.<br />

TG 7 (hCD59) + End. + Card. ++ Alv., End. ) ) ) ) ) ) ) ) ) ) ) )<br />

TG 8 (hCD59) ) ++ ++ Alv., End. ) + End. ) ) + Acin. ++ End. ) ++ End. + End. ) ++ End. ++ End.<br />

TG1 (hCD59) 8 + End. ) ++ Alv., End. ) ) ++ End. ) ) ) ) ) ) ) ) )<br />

TG 4 (hCD59) ) + End. ++ Alv., End ) ) ) ++ End. ) ) ) + End. ) ) ) ++ End.<br />

TG 15 (hCD59) ) ++ End. ++ Alv., End. ) ) ++ End. ) + Acin. +++ End. ++ End. ) ) ) ) )<br />

TG 20 (hCD59) +++ End. + End. ++ Alv., End. ) ) + End. ) ) ) ) ) ) ) ) )<br />

TG2 (hCD59) 10 ) ) ) ) ) ) ) ) ) ) ) ) ) ) )<br />

TG5 (hCD59) ) ) ) ) ) ) ) ) ) ) ) ) ) ) )<br />

TG23 (hCD55) 13 ) ) ) ) ) ) ) ) ) ) ) ) ) ) )<br />

TG13 (hCD59) 17 ) ) ) ) ) ) ) ) ) ) ) ) ) ) )<br />

TG14 (hCD59) 19 ) ) ) ) ) ) ) ) ) ) ) ) ) ) )<br />

TG19 (hCD59) ++ End. ++ End. +++ Alv., End. ) ++ End. + End. ) ) ) + Myoc. ) ) ) ) )<br />

TG18 (hCD59) 20 +++ End. + End. +++ Alv., End. ) ) + End. ) ) ) ++ Myoc. ++ Myoc. ) ) ) ++ End.<br />

TG21 (hCD59) ) ) ) ) ) ) ) ) ) ++ Myoc. ++ Myoc. ) ) ) )<br />

TG25 (hCD59) ) ) ) ) ) ) ) ) ) + Myoc. + Myoc. ) ) ) )<br />

TG27 (hCD59) ) ) ) ) ) ) ) ) ) + Myoc. + Myoc. ) ) ) )<br />

TG12 (hCD59) 21 ) ) ) ) ) ) ) ) ) ) ) ) ) ) )<br />

TG26 (hCD59+hCD55) 15 · 20 ) ) ) ) ) ) ) ) ) + Myoc. + Myoc. ) ) ) )<br />

The human complement regulatory protein hCD59<br />

Number and intensity of positively staining cells (referring to the tissue in general): +++ strong, ++ moderate, + weak, ) negative.<br />

End., endothelial cells; Card., cardiomyocytes; Alv., alveolar epithelial cells; Ac., acinic cells; En., endocrine pancreas (islet cells); Myoc., myocytes.<br />

351


Deppenmeier et al.<br />

Table 4. Comparative analysis of hCD59 protein and mRNA expression in transgenic pigs<br />

Heart Lung Spleen Liver Kidney Pancreas Brain<br />

with neighboring cells remaining negative for the<br />

hCD59 protein (Fig. 3). The following organs<br />

showed a positive, intracytoplasmic staining for<br />

Skeletal<br />

muscle Esophagus<br />

Prot. RNA Prot. RNA Prot. RNA Prot. RNA Prot. RNA Prot. RNA Prot. RNA Prot. RNA Prot. RNA<br />

TG6 5 ++ ) ) ) + ) + ) ) ) +++ ) ++ ) ) ) ++ )<br />

TG7 5 + ++(+) ++ ++ (+) ) ) ) ++ ) ) ) n.d. ) n.d. ) ) ) )<br />

TG8 5 ++ ) ++ + ) ) + ++ ) ) + ) ++ ++ ) ) ++ )<br />

TG1 8 ) ) ++ ) ) + ) ) ++ +++ ) ) ) ) ) ) ) )<br />

TG4 8 + ) ++ ) ) ) ) ) ) ) ) ) ) ) ) ) + )<br />

TG15 8 ++ ++ ++ ++ ) ++ ) ) ++ ) + ) +++ + ) ) ) ++<br />

TG20 8 + ++ ++ + ) +++ ) ) + +++ ) ) ) + ) + ) ++<br />

TG2 10 ) ) ) ) ) ++ ) ) ) ) ) ) ) ) ) ) ) )<br />

TG5 10 ) ) ) ) ) ) ) ) ) ) ) ) ) ) ) ) ) )<br />

TG23 13 ) ) ) ) ) ) ) ) ) ) ) ) ) ) ) ) ) )<br />

TG13 17 ) ) ) ) ) ) ) ) ) + ) + ) ) ) + ) +<br />

TG14 19 ) ) ) ) ) ++ ) ++ ) ++ ) + ) + ) ) ) )<br />

TG19 19 ++ ++(+) +++ +++ ) +++ ++ +++ + +++ ) +(+) ) +++ + +++ ) +++<br />

TG18 20 + +++ +++ ++ ) ++ ) ++ + ++ ) ++ ) +(+) ++ +++ ++ +++<br />

TG21 20 ) + ) ++ ) ++(+) ) (+) ) ++(+) ) ) ) ) ++ +++ ++ ++(+)<br />

TG25 20 ) ++ ) ++ ) ++ ) +++ ) ) ++ ) n.d. + +++ + +++<br />

TG27 20 ) ) ) +(+) ) ) ) ) ) ++ ) ) ) ) + ++ + +<br />

TG12 21 ) ) ) + ) +(+) ) ) ) + ) ) ) ) ) ) ) )<br />

TG26 15<br />

20<br />

) ) ) ) ) ) ) ) ) ) ) ) ) ) + +(+) + +<br />

n.d., not done.<br />

+++ strong, ++(+) strong to moderate, ++ moderate, +(+) weak to moderate, + weak, ) negative.<br />

352<br />

Fig. 2. Variable hCD59 mRNA expression<br />

as demonstrated for a hCD59 transgenic<br />

animal. (A) hCD59 mRNA could be<br />

detected by RT-PCR in heart, lung, spleen,<br />

kidney, brain, muscle and esophagus<br />

whereas liver and pancreas virtually<br />

showed no expression of the transgene.<br />

hCD59-transfected cells served as the<br />

positive control. Non-transgenic pig<br />

without expression. (B) RT-PCR detection<br />

of EF1 a-transcripts.<br />

hCD59 protein: skin, heart, lung, spleen, liver,<br />

kidney, urinary bladder, skeletal muscle, esophagus,<br />

duodenum, jejunum, ileum and colon. Endo-


thelial cells of small and medium blood vessels<br />

expressed hCD59 protein. In the lung, not only<br />

endothelial cells, but also alveolarepithelial cells<br />

The human complement regulatory protein hCD59<br />

Fig. 3. Representative staining pattern of hCD59 protein in a variety of organs derived from different transgenic animals. (A) Kidney<br />

(TG1): expression in glomerular endothelial cells (arrow) and in the endothelial cells of an afferent arteriole (arrowhead); (B) Lung<br />

(TG8): expression in alveolar epithelial cells type I (arrowhead) and type II (arrow); (C) Heart (TG7): expression in the cytoplasm of<br />

a cardiomyocyte (arrow) without staining of the neighboring cells, ‘‘patchy pattern’’; (D) Pancreas (TG6): expression in acinary cells<br />

of the exocrine pancreas and in single islet cells (arrows); (E) Pancreas (TG15): expression in acinary cells of the exocrine pancreas<br />

(arrow), ‘‘patchy pattern’’, no staining of endothelial cells; (F) Heart (TG6): expression in endothelial cells (arrow); (G) Brain<br />

(TG15): expression in endothelial cells (arrow); (H) Skin (TG6): expression in endothelial cells (arrow); (I) human tonsillar tissue used<br />

as positive control: expression in endothelial cells; (J–O) negative controls of a non-transgenic pig (WT4): (J) kidney; (K) lung; (L)<br />

heart; (M) pancreas; (N) brain; (O) skin. A, D, I, J, M, O: ·250; B, C, E, F, G, H, K, L, N: ·400 (bar 100 lm).<br />

type I and II showed hCD59 labeling. In the<br />

pancreas, acinic cells of the exocrine pancreas, and<br />

islet cells of the endocrine pancreas showed hCD59<br />

353


Deppenmeier et al.<br />

protein expression. The acinic cells of the exocrine<br />

pancreas of one animal (TG15) showed a patchy<br />

expression pattern. The cytoplasm, especially the<br />

outer membrane, of some cardiomyocytes in animal<br />

TG7 stained patchy for hCD59. The myocytes<br />

in the skeletal muscle and the muscular layer of the<br />

esophagus in six animals (TG18, TG19, TG21,<br />

TG25, TG26, TG27) showed a patchy expression<br />

pattern of hCD59, with only a part of the skeletal<br />

muscle cells showing a positive reaction. In the<br />

kidney of one animal (TG1), endothelial cells of<br />

the capillaries in the glomerula and afferent arteries<br />

showed a positive staining, with only single glomerula<br />

showing the positive reaction. Non-transgenic<br />

animals never showed expression of hCD59protein.<br />

Discussion<br />

The study reported here is the first systemic<br />

analysis to detect pathological side effects potentially<br />

associated with expression of a specific<br />

transgene, using hCD59 as a model. Results show<br />

that in transgenic pigs produced by microinjection<br />

of foreign DNA into pronuclei the health status is<br />

not different from that of the wild-type control<br />

animals. A total of 19 transgenic pigs from nine<br />

lines were analyzed. As the production and breeding<br />

of transgenic pigs is a time consuming, complex<br />

and costly process, we restricted the number of<br />

transgenic animals to be analyzed to one to four<br />

animals per transgenic line and covered a rather<br />

broad range of different ages. This approach<br />

ensures representative results also for other transgenic<br />

programs. Wild-type animals from the same<br />

facility served as controls.<br />

It is well known that random integration of a<br />

foreign DNA could be associated with insertional<br />

mutagenesis [9]. A major prerequisite for a potential<br />

usage of transgenic porcine organs in xenotransplantation<br />

is that animal health and welfare<br />

are not compromised. Here, we show that microinjection<br />

of foreign DNA is not associated with a<br />

specific pathomorphological phenotype in pigs.<br />

Some of the animals had chronic interstitial<br />

nephritis, pneumonia or epi- and pericarditis.<br />

These lesions are common pathological features<br />

especially in older animals in modern pig populations<br />

and were probably caused by previous viral<br />

or bacterial infections [29]. Because of the chronic<br />

nature of these lesions, the causative agent could<br />

not be detected any more. Lesions in organs like<br />

kidney or lung show that it is of crucial importance<br />

to check the organs for transferable diseases before<br />

xenotransplantation. The activation of spleen and<br />

lymph nodes as shown by follicular hyperplasia,<br />

354<br />

edema and sinusoidal histiocytosis, can be attributed<br />

to a previous challenge with unspecific pathogens.<br />

As the animals were not maintained under<br />

gnotobiotic conditions, but kept under SPF conditions,<br />

a challenge with ubiquitous pathogens is<br />

likely to occur. The deposition of siderin-containing<br />

macrophages in the sinus of the ileofemoral<br />

lymph nodes are likely remnants from ferric<br />

injections which are commonly performed in pig<br />

populations. Non-transgenic control animals and<br />

transgenic animals were affected in approximately<br />

the same ratio and to the same extent by the lesions<br />

(Table 2). Prior to the use of porcine organs in<br />

xenotransplantation, the donor animals must be<br />

screened thoroughly for possible infections with<br />

pathogens, because of the risk of transmitting these<br />

pathogens to the host [5].<br />

In the majority of the transgenic and nontransgenic<br />

animals PCV 2 was detected by PCR.<br />

PCV 2 is found widely within modern pig populations<br />

and in most cases a PCV 2 infection is not<br />

associated with clinical or pathomorphological<br />

changes. In some cases the infection can be<br />

associated with certain diseases, including PMWS<br />

or PDNS [30]. In this study, none of the PCV IIpositive<br />

transgenic and non-transgenic pigs showed<br />

any signs of such PCV II-associated diseases.<br />

In accordance with the original report a varying<br />

expression pattern for hCD59 was detected in lines<br />

5, 17, 19, 20 and 21. Lines 17, 19, 20 and 21 showed<br />

a muscle fiber-confined expression of hCD59. In<br />

some cases the mRNA, but not the protein could<br />

be detected. This might be due to a higher<br />

sensitivity of the PCR assay compared with the<br />

immunohistology.<br />

CD59 mRNA was variably expressed in the<br />

tested animals. In few animals mRNA expression<br />

of the transgene was not detected (Table 4). As<br />

hCD59 protein was detected in heart, lung, and<br />

skin of these animals (Table 3), this most likely<br />

reflects RNA degradation in these samples.<br />

Similar to the hCD59 mRNA, the protein<br />

expression pattern was variable as well. Despite<br />

the use of an ubiquitously active promoter (CMV)<br />

no animal expressed the protein in all organs,<br />

instead most animals expressed the transgene in a<br />

subset of organs. In addition, a patchy expression<br />

pattern could be detected in heart and kidney. A<br />

likely explanation for the variable expression<br />

pattern could be the promoter and/or the impact<br />

of chromosomal flanking regions. The CMV promoter<br />

had been selected because it is known to<br />

induce a strong and ubiquitous transgenic expression<br />

in cell lines and transgenic mice as well [31,32].<br />

However, reports from different transgenic animal<br />

models confirm our finding that CMV and


CMV-derived promoter expression can vary both<br />

within different lines and even between tissues of a<br />

single line, and can thus result in mosaic patterns<br />

even within one tissue [33–36]. Recent studies and<br />

own results point to de novo DNA methylation in<br />

transgenic pigs as a mechanism to silence expression<br />

[23,37]. Crossbreeding experiments of lines 13<br />

and 20 resulted in a broad tissue-unspecific expression<br />

of hCD59 [23] highlighting that breeding of<br />

the existing lines might yield pigs with an ubiquitous<br />

expression of human complement regulators.<br />

In summary, no evidence for pathomorphological<br />

changes was found in a representative number<br />

of transgenic pigs from nine transgenic pig lines of<br />

the F1 and F2 generation, including animals with<br />

high levels of the transgenic expression in several<br />

tissues. Results of this study provide compelling<br />

evidence that stable integration of constructs<br />

encoding the human complement regulatory protein<br />

hCD59 into the pig genome and its expression<br />

do not affect the health status of pigs. Results of<br />

the present study are crucial for the application of<br />

the technology and the use of transgenic pigs for<br />

biomedical and agricultural applications.<br />

Acknowledgments<br />

The expert technical support by KG. Hadeler, H.<br />

Hornbostel, W. Hasselbring and T. Peker are<br />

gratefully acknowledged. H.N., K.W. and H.K<br />

received support via SFB 265 from the DFG.<br />

References<br />

1. Galili U, Rachmilewitz EA, Peleg A, Flechner I. A<br />

unique natural human IgG antibody with anti-a-galactosyl<br />

specificity. J Exp Med 1984; 160: 1519.<br />

2. Galili U, Clark MR, Shohet SB, Buehler J, Macher<br />

BA. Evolutionary relationship between the natural anti-<br />

Gal antibody and the Gala1–3Gal epitopes in primates.<br />

Proc Natl Acad Sci USA 1987; 84: 1369.<br />

3. Platt JL, Fischel RJ, Matas AJ et al. Immunopathology<br />

of hyperacute xenograft rejection in a swine-to-primate<br />

model. Transplantation 1991; 52: 214.<br />

4. Rose AG, Cooper DKC. Venular thrombosis is the key<br />

event in the pathogenesis of antibody-mediated cardiac<br />

rejection. Xenotransplantation 2000; 7: 31.<br />

5. Kues WA, Niemann H. The contribution of farm animals<br />

to human health. Trends Biotechnol 2004; 22: 286.<br />

6. Davies A, Simmons DL, Hale G et al. CD59, an<br />

LY-6-like protein expressed in human lymphoid cells,<br />

regulates the action of the complement membrane attack<br />

complex on homologous cells. J Exp Med 1989; 170: 637.<br />

7. Meri S, Morgan BP, Davies AR et al. Human protectin<br />

(CD59), an 18,000–20,000 MW complement lysis restricting<br />

factor, inhibits C5b-8 catalysed insertion of C9 into<br />

lipid bilayers. Immunology 1990; 71: 1.<br />

8. Rollins SA, Sims PJ. The complement-inhibitory activity<br />

of CD59 resides in its capacity to block incorporation of<br />

C9 into membrane C5b-9. J Immunol 1990; 144: 3478.<br />

The human complement regulatory protein hCD59<br />

9. Niemann H, Kues WA. Transgenic livestock: premises<br />

and promises. Anim Reprod Sci 2000; 60–61: 277.<br />

10. Fox MW. Genetic engineering biotechnology: animal<br />

welfare and environmental concerns. Appl Anim Behav<br />

Sci 1988; 20: 83.<br />

11. Hughes BO, Hughes GS, Waddington D, Appleby<br />

MC. Behavioural comparison of transgenic and control<br />

sheep: movement order, behaviour on pasture and in<br />

covered pens. J Anim Sci 1996; 63: 91.<br />

12. Niemann H, Halter R, Carnwath JW et al. Expression<br />

of human blood clotting factor VIII in the mammary<br />

gland of transgenic sheep. Transgenic Res 1999; 8: 237.<br />

13. Van Reenen CG, Meuwissen THE, Hopster H et al.<br />

Transgenesis may affect farm animal welfare: a case for<br />

systemic risk assessment. J Anim Sci 2001; 79: 1763.<br />

14. Pursel VG, Pinkert CA, Miller KF et al. Genetic<br />

engineering of livestock. Science 1989; 244: 1281.<br />

15. Nottle MB, Nagashima H, Verma PJ et al. Production<br />

and analysis of transgenic pigs containing a metallothionein<br />

porcine growth hormone gene construct. In:<br />

Murray JD, Anderson GB, Oberbauer AM, McGloughlin<br />

MM, eds. Transgenic Animals in Agriculture.<br />

New York, NY, USA: CABI Publ., 1999. 145–156.<br />

16. Niemann H, Wrenzycki C. Alterations of expression of<br />

developmentally important genes in preimplantation bovine<br />

embryos by in vitro culture conditions: implications<br />

for subsequent development. Theriogenology 2000; 53: 21.<br />

17. Cibelli JB, Stice SL, Golueke PJ et al. Cloned transgenic<br />

calves produced from nonquiescent fetal fibroblasts.<br />

Science 1998; 280: 1256.<br />

18. Hill JR, Roussel AJ, Cibelli JB et al. Clinical and<br />

pathologic features of cloned transgenic calves and features<br />

(13 case studies). Theriogenology 1999; 51: 1451.<br />

19. Carter DB, Lai L, Park KW et al. Phenotyping of<br />

transgenic cloned piglets. Cloning Stem Cells 2002; 4: 131.<br />

20. Park KW, Cheong HT, Lai L et al. Production of nuclear<br />

transfer-derived swine that express the enhanced green<br />

fluorescent protein. Anim Biotechnol 2001; 12: 173.<br />

21. Lai L, Kolber-Simonds D, Park KW et al. Production<br />

of a-1,3-galactosyltransferase knockout pigs by nuclear<br />

transfer cloning. Science 2002; 295: 1089.<br />

22. Niemann H, Verhoeyen E, Wonigeit K et al. Cytomegalovirus<br />

early promoter induced expression of hCD59<br />

in porcine organs provides protection against hyperacute<br />

rejection. Transplantation 2001; 72: 1898.<br />

23. Kues WA, Schwinzer R, Wirth D et al. Epigenetic<br />

silencing and tissue independent expression of a novel<br />

tetracycline inducible system in double transgenic pigs.<br />

FASEB J 2006; Express fj. 05-4515 fjev1.<br />

24. Lehmann U, Kreipe H. Real time PCR analysis of DNA<br />

and RNA extracted from formalin-fixed and paraffinembedded<br />

biopsies. Methods 2001; 25: 409.<br />

25. Bock O, Kreipe H, Lehmann U. One-step extraction of<br />

RNA from archival biopsies. Anal Biochem 2001; 295: 116.<br />

26. Gruber AD, Levine RA. In situ assessment of mRNA<br />

accessibility in heterogenous tissue samples using elongation<br />

factor-1a (EF-1a). Histochem Cell Biol 1997; 107: 411.<br />

27. Von Wasielewski R, Mengel M, Wiese B et al. Tissue<br />

array technology for testing interlaboratory and interobserver<br />

reproducibility of immunohistochemical estrogen<br />

receptor analysis in a large multicenter trial. Am J Clin<br />

Pathol 2002; 118: 675.<br />

28. Mengel M, Werner M, Von Wasielewski R. Concentration<br />

dependent and adverse effects in immunohistochemistry<br />

using the tyramine amplification technique.<br />

Biochem J 1999; 31: 195.<br />

355


Deppenmeier et al.<br />

29. Jubb KVF, Kennedy PC, Palmer N. Pathology of<br />

Domestic Animals, 4th edn, Vol. 2. San Diego, CA, USA:<br />

Academic Press, 1993.<br />

30. Krakowka S, Ellis JA, McNeilly F et al. Activation of<br />

the immune system is the pivotal event in the production<br />

of wasting disease in pigs infected with porcine circovirus-<br />

2 (PCV-2). Vet Pathol 2001; 38: 31.<br />

31. Boshart M, Weber F, Jahn G et al. A very strong<br />

enhancer is located upstream of an immediate early gene<br />

of human cytomegalovirus. Cell 1985; 41: 521.<br />

32. Fodor WL, Williams BL, Matis LA et al. Expression<br />

of a functional human complement inhibitor in a transgenic<br />

pig as a model for the prevention xenogeneic<br />

hyperacute organ reaction. Proc Natl Acad Sci USA 1994;<br />

91: 11153.<br />

33. Furth PA, Hennighausen L, Baker C, Beatty B,<br />

Woychick R. The variability in activity of the universally<br />

356<br />

expressed human cytomegalovirus immediate early gene 1<br />

enhancer/promotor in transgenic mice. Nucleic Acids Res<br />

1991; 19: 6205.<br />

34. McGrew MJ, Sherman A, Ellard FM et al. Efficient<br />

production of germline transgenic chickens using lentiviral<br />

vectors. EMBO Rep 2004; 5: 1.<br />

35. Baskar JF, Smith PP, Nilaver G et al. The enhancer<br />

domain of the human cytomegalovirus major immediateearly<br />

Promotor determines cell type-specific expression in<br />

transgenic mice. J Virol 1996; 70: 3207.<br />

36. Furth PA, St Onge L, Boger H et al. Temporal control<br />

of gene expression in transgenic mice by a tertracyclineresponsive<br />

Promotor. Proc Natl Acad Sci USA 1994; 91:<br />

9302.<br />

37. Hofmann A, Kessler B, Ewerling S et al. Epigenetic<br />

regulation of lentiviral transgene vectors in a large animal<br />

model. Mol Ther 2006; 13: 59.


Arch Virol (2006)<br />

DOI 10.1007/s00705-006-0868-y<br />

Printed in The Netherlands<br />

Brief Report<br />

Inhibition of porcine endogenous retroviruses (PERVs) in primary<br />

porcine cells by RNA interference using lentiviral vectors<br />

B. Dieckhoff 1; , A. Karlas 1; , A. Hofmann 2 , W. A. Kues 3 , B. Petersen 3 , A. Pfeifer 2 , H. Niemann 3 ,<br />

R. Kurth 1 , and J. Denner 1<br />

1 Robert Koch-Institute, Berlin, Germany<br />

2 Department of Pharmacy, Ludwig-Maximilians-University, M€unchen, Germany<br />

3 Department of Biotechnology, Institute for Animal Science, Neustadt, Germany<br />

Received July 14, 2006; accepted September 13, 2006; published online November 16, 2006<br />

# Springer-Verlag 2006<br />

Summary<br />

A potential risk in pig-to-human xenotransplantation<br />

is the transmission of PERVs to human recipients.<br />

Here we show for the first time the inhibition of<br />

PERV expression in primary porcine cells by RNA<br />

interference using lentiviral vectors. Cells were<br />

transduced with lentiviral vectors coding for short<br />

hairpin (sh) RNAs directed against PERV. In all<br />

primary porcine cells studied and in the porcine<br />

kidney cell line PK-15, expression of PERV-mRNA<br />

was significantly reduced as measured by real-time<br />

PCR. Most importantly, expression of PERV proteins<br />

was almost completely suppressed, as shown<br />

by Western blot analysis. Thus, lentiviral shRNA<br />

vectors could be used to knockdown PERV expression<br />

and create transgenic pigs with a reduced risk of<br />

PERV transmission during xenotransplantation.<br />

The transplantation of porcine xenografts may offer<br />

a potential way to overcome the shortage of<br />

Authors contributed equally.<br />

Author’s address: Dr. Joachim Denner, Robert<br />

Koch-Institute, Nordufer 20, 13353 Berlin, Germany.<br />

e-mail: DennerJ@rki.de<br />

allogeneic organs. However, xenotransplantation<br />

may be associated with the risk of transmission of<br />

PERVs, which are integrated in the porcine genome<br />

[3, 23] and have been found to be released as infectious<br />

particles from normal pig cells [15, 17, 26].<br />

Three subtypes of PERVs (A, B and C) are known<br />

which mainly differ in the env region encoding for<br />

the receptor-binding site [12, 23, 29]. PERV-A and<br />

PERV-B are polytropic viruses and able to infect<br />

human cells in vitro [16, 22, 25, 29, 30], whereas<br />

PERV-C is an ecotropic virus. Even though PERVs<br />

are able to infect human cells in vitro, no virus<br />

transmission has been observed in previous experimental<br />

and clinical xenotransplantations [6, 8, 21,<br />

27]. However, in most of these approaches, the time<br />

of contact was short and the immunosuppression<br />

was weak. In contrast, the intended conditions for<br />

future xenotransplantations include long-term transplantation<br />

of porcine cells or organs, direct contact<br />

of human and porcine cells, and a strong immunosuppression<br />

to avoid rejection. PERV transmission<br />

has also not been observed in pig-to-nonhuman primate<br />

transplantations and infection experiments<br />

[2, 14, 31]. However, the risk of PERV transmission<br />

associated with tumours or immunodeficiencies<br />

is still existent. Therefore, different strategies to


prevent PERV transmission have been developed,<br />

such as the selection of low-producer animals [20,<br />

28], the generation of an antiviral vaccine [4], and<br />

the inhibition of PERV expression by RNA interference<br />

[9, 18]. In these first experiments, inhibition<br />

of PERV expression by small interfering RNAs<br />

(siRNAs) was studied in PERV-infected human<br />

cells.<br />

In this study, we describe for the first time inhibition<br />

of PERV expression in primary porcine cells<br />

transduced by lentiviral vectors expressing anti-<br />

PERV shRNA. We studied the genetic distribution,<br />

the expression pattern of different PERV subtypes,<br />

and the inhibitory effect of two different lentiviral<br />

vectors on PERV expression.<br />

In order to analyze the presence of PERV-A,<br />

PERV-B and PERV-C proviruses, DNA was isolated<br />

from three foetal fibroblast cultures (SE101, SE105<br />

and P1F10) obtained from individual D25 (p.c.)<br />

German landrace foetuses [11] as well as from<br />

PK-15 cells, using the QIAamp DNA Blood Mini<br />

Kit (Qiagen, Hilden, Germany), and a PCR was<br />

Fig. 1. Analysis of PERV provirus integration and expression<br />

in porcine primary fibroblasts and PK-15 cells. A<br />

Detection of PERV proviruses in the genome of two independent<br />

primary fibroblast cultures (SE101, P1 F10) and<br />

PK-15 cells using PCR. B PERV expression in foetal fibroblasts<br />

(SE101 and P1 F10) and in PK-15 cells using onestep<br />

RT-PCR. In addition to specific primers for env-A,<br />

env-B, env-C, primers specifically detecting full-length<br />

and spliced env mRNA were used. Integrity of RNA was<br />

determined by amplifying GAPDH<br />

B. Dieckhoff et al.<br />

performed using primers specific for env of subtypes<br />

A, B [12] and C [29]. All proviruses were<br />

found integrated in the genome of the primary<br />

fibroblasts P1 F10 and SE101, whereas PK-15 cells<br />

did not contain PERV C (Fig. 1A).<br />

In order to study PERV expression, total RNA<br />

was isolated (RNeasy Kit, Qiagen), and a onestep<br />

reverse transcription (RT) PCR (Invitrogen,<br />

Karlsruhe, Germany) was performed using the<br />

same primers as described above as well as primers<br />

allowing the discrimination between viral fulllength<br />

RNA (forward: 5 0 TGCTGTTTGCATCAA<br />

GACCGC, reverse: 5 0 ACAGACACTCAGAA<br />

CAGAGAC) and spliced env-mRNA (forward: 5 0<br />

TGCTGTTTGCATCAAGACCGC, reverse: 5 0<br />

ATGGAGGCGAAGCTTAAGGGGA). The integrity<br />

of the RNA was determined by amplifying the<br />

housekeeping gene glyceraldehyde-3-phosphate<br />

dehydrogenase (GAPDH, for: 5 0 CTGCCCCTT<br />

CTG CTGATGC; rev: 5 0 TCC ACGATGCCGAA<br />

GTTGTC). In all primary cells tested, full-length<br />

mRNAs corresponding to the PERV subtypes A and<br />

BandCaswellassplicedenv mRNA were detected<br />

(Fig. 1B). Since no PERV-C proviruses were integrated<br />

in PK-15 cells, no PERV-C-specific mRNA<br />

was found (Fig. 1B). The presence of spliced env<br />

mRNA indicates that Env protein may be generated<br />

and consequently virus particles may be produced.<br />

In our previous publications, different synthetic<br />

siRNAs corresponding to several parts of the sequence<br />

of PERV were designed and screened for<br />

their ability to inhibit expression of PERV in infected<br />

human 293 cells [9, 10]. The most effective<br />

siRNA corresponded to a highly conserved sequence<br />

in the pol gene (pol2); the target sequence<br />

is identical in all functional PERV strains. Like<br />

other siRNA used, the pol2 siRNA was able to inhibit<br />

PERV-A and PERV-B [10]. Inserted into the<br />

vector pSuper, permanent expression of this polspecific<br />

shRNA was achieved, and efficient inhibition<br />

of PERV expression was observed [9, 10]. This<br />

pSuper-pol2 construct was chosen as a basis for<br />

establishing a lentiviral expression system.<br />

Two different lentiviral vectors, RRL-pGK-GFP<br />

[5] (kindly provided by L. Naldini) and pLVTHM<br />

[32] (kindly provided by D. Trono), which allow<br />

expression of the corresponding short hairpin


Inhibition of PERV expression by RNA interference<br />

Fig. 2. Schematic representation of the lentiviral vectors<br />

RRL-pGK-GFP and pLVTHM. LTR long terminal repeat,<br />

H1 polymerase III H1-RNA gene promotor, shRNA short<br />

hairpin RNA, SIN self-inactivating element, cPPT central<br />

polypurine tract, GFP green fluorescent protein, WPRE<br />

post-transcriptional element of woodchuck hepatitis virus,<br />

EF-1alfa elongation factor-1a promoter, teto tetracycline<br />

operator, hPGK phosphoglycerate kinase promotor, loxP<br />

locus of X-over of P1, recombination site of Cre recombinase<br />

for bacteriophage P1<br />

RNAs, were used (Fig. 2). They differed in the<br />

promoters inducing the expression of the GFP reporter<br />

(pLVTHM used an EF1a promoter, whereas<br />

RRL-pGK-GFP used a PGK promoter) and in the<br />

arrangement of shRNA and GFP expression cassettes.<br />

In the case of the pLVTHM, the shRNAexpressing<br />

cassette is located within the 3 0 -LTR<br />

(long terminal repeat). The shRNA expression cassette<br />

of the vector pSuper-pol2 was inserted into<br />

pLVTHM and RRL-pGK-GFP using the restriction<br />

enzymes EcoRI and ClaI, and EcoRI and HindIII<br />

(New England Biolab, Frankfurt, Germany), respectively,<br />

resulting in pLVTHM-pol2 and RRLpGK-GFP-pol2.<br />

Lentiviral particles were obtained<br />

by co-transfecting the packaging cell line 293T with<br />

the envelope plasmid pCL-VSV-G, the packaging<br />

plasmid psPAX2, the transfer vector pLVTHM,<br />

and the transfer vector pLVTHM-pol2. Titration of<br />

concentrated lentiviral particles was performed on<br />

Hela cells as described (lentiweb.com). The titres<br />

were 1.73 10 8 IU=ml for pLVTHM and 1.78<br />

10 7 IU=ml for pLVTHM-pol2. RRL-pGK-GFPpol2<br />

and RRL-pGK-GFP were produced as described<br />

recently [7, 24]. The titres were 2.25<br />

10 10 and 2.0 10 10 IU=ml, respectively.<br />

For transduction of foetal fibroblasts and PK-15<br />

cells, 1 10 5 cells were seeded in a 6-well culture<br />

plate in the presence of 6 mg=ml polybrene in<br />

DMEM and incubated with concentrated lentiviral<br />

particles at a MOI of 2.5 for 5 days. By fluorescence<br />

microscopy and flow cytometry, transduction<br />

efficiencies of 90% were measured. In the case of<br />

lower efficiencies, transduced cells were enriched<br />

by cell sorting (MoFlo, Dako, Glostrup, Denmark)<br />

or repeated addition of lentiviral particles. As a result,<br />

transduction efficiencies of around 95% were<br />

achieved.<br />

Total RNA and protein were isolated from transduced<br />

and non-transduced foetal fibroblasts and<br />

PK-15 cells using TRI Reagent (Sigma, Steinheim,<br />

Germany) and the RNeasy Kit (Qiagen). To measure<br />

PERVexpression quantitatively, a SYBR Green<br />

real-time RT-PCR was performed as described<br />

previously [9]. An inhibition of PERV mRNA expression<br />

of 82.5 and 78% was observed in foetal<br />

fibroblasts SE101 and SE105 using the vector RRL-<br />

PGK-GFP-pol2 (Fig. 3A). In addition, a one-step<br />

RT real-time PCR (Invitrogen) was performed<br />

using a FAM-labelled probe (5 0 FAM-AGAAGG<br />

GACCTTGGCAGACTTTCT [3BQ1], Sigma),<br />

primers specific for PERV gag (forward: 5 0<br />

TCCAGGGCTCATAATTTGTC, reverse: 5 0 TGA<br />

TGGCCATCCAACATCGA), and the Mx4000<br />

multiplex quantitative PCR system (Stratagene,<br />

La Jolla, USA). The assay revealed a coefficient<br />

of correlation of 0.980 with an almost optimal efficiency<br />

of amplification. For each reaction, 50 ng<br />

total RNA was used. The amount of full-length<br />

RNA was calculated relative to the constituted<br />

amount of total RNA using a modification of the<br />

CT<br />

2 method [13], and the expression of PERV<br />

in cells transduced with shRNA lentiviral particles<br />

was compared with the expression in cells transduced<br />

with the control vector as well as non-transduced<br />

cells. An inhibition of PERV full-length<br />

mRNA expression of 81.8% was observed in foetal<br />

fibroblasts P1 F10 transduced with pLVTHM-pol2<br />

(Fig. 3A). When P1 F10 cells that had been sorted<br />

for GFP-positive cells in order to increase the number<br />

of transduced cells (nearly 99%) were analysed,<br />

the inhibition of PERV expression increased up<br />

to 94.8%. Inhibition of PERV expression in trans-


B. Dieckhoff et al.<br />

duced cells was stable and remained at 95% over<br />

months (Fig. 3A). Finally, it is important to note that<br />

expression of PERV in primary fibroblasts was very<br />

low (0.2%) when compared with the expression in<br />

virus-producing PK-15 cells (defined as 100%).<br />

After showing that pLVTHM-pol2 and RRL-<br />

PGK-GFP-pol2 were able to inhibit PERV expression<br />

in foetal fibroblasts, their effect in another cell<br />

type was studied. Epithelial PK-15 cells were transduced<br />

with both lentiviral vectors, RRL-PGK-GFPpol2<br />

and pLVTHM-pol2, at an efficiency of 95%<br />

and showed an inhibited PERV expression of 65.8<br />

and 73.5%, respectively (Fig. 3B), indicating that<br />

PERV expression can be inhibited in different cell<br />

types. PERV expression was also reduced in PHAtreated<br />

PBMCs, which were difficult to analyse<br />

since the transduction efficiency was relatively<br />

low. However, after enrichment of GFP-expressing<br />

PBMCs by fluorescence-activated cell sorting, an<br />

inhibition of PERV full-length mRNA expression<br />

of 75.2% was observed.<br />

In order to analyse whether reduced expression<br />

of PERV mRNA leads to a reduced expression of<br />

viral proteins, protein samples of foetal fibroblasts<br />

P1 F10 and PK-15 cells were analysed using a<br />

Western blot assay. Analysing protein lysates of<br />

transduced and non-transduced PK-15 cells (50 mg)<br />

by Western blot with a goat anti-p27Gag-antiserum<br />

[8], a reduced PERV expression was detected in<br />

cells transduced with both lentiviral pol2 vectors<br />

(Fig. 3C). Loading of equal protein amounts was<br />

verified using antibodies against b-actin (Sigma).<br />

However, even if 80 mg total protein lysate of fetal<br />

fibroblasts P1 F10 was used, no Gag protein was<br />

detected even in the non-transduced control cells,<br />

1<br />

Fig. 3. Expression of PERV in cells transduced with the<br />

lentiviral vectors RRL-PGK-GFP-pol2 and pLVTHM-pol2<br />

and in cells transduced with control vector as well as in<br />

non-transduced cells (set 100%). A Three batches of foetal<br />

fibroblasts and B PK-15 cells transduced with both vectors<br />

were analysed at day 5 or at other time points after transduction<br />

using one-step RT real-time PCR. C Western blots<br />

showing expression of p27Gag in PK-15 cells transduced<br />

with RRL-PGK-GFP-pol2 and pLVTHM-pol2 and in controls.<br />

D RT activity in the supernatants of PK-15 cells<br />

transduced with RRL-PGK-GFP-pol2 and pLVTHM-pol2,<br />

and control cells


Inhibition of PERV expression by RNA interference<br />

suggesting that the PERV protein expression in<br />

these cells is under the detection level.<br />

To analyse whether viral reverse transcriptase<br />

(RT) activity was released into the supernatants of<br />

control and transduced foetal fibroblasts P1 F10<br />

and PK-15 cells, an RT activity assay was performed<br />

(Cavidi Bio Tech, Uppsala, Sweden). Transduced<br />

and control PK-15 cells were seeded at a density<br />

of 5 10 4 cells in 24-well plates and RT activity<br />

was measured in the supernatants after three<br />

days. This assay demonstrated a reduced RT activity<br />

in the supernatants of PK-15 cells transduced<br />

with both lentiviral vectors, RRL-PGK-GFP-pol2<br />

(5.78 mU=ml) and pLVTHM-pol2 (4.74 mU=ml),<br />

compared with the expression of untreated control<br />

cells (21.21 mU=ml) (Fig. 3D). In the supernatants<br />

of fetal fibroblasts P1 F10, no RT activity (0 mU=ml)<br />

was detected, confirming the low expression of<br />

PERV even in untreated cells.<br />

In this study, we show for the first time a significant<br />

inhibition of PERV expression in primary<br />

porcine cells by lentiviral vector-mediated RNA<br />

interference. These data correlate with previous experiments<br />

investigating PERV-specific RNA interference<br />

in human cells infected with PERV-B<br />

or with a recombinant PERV-A=C using synthetic<br />

siRNAs and the pSuper vector system [9, 10, 18].<br />

Lentiviral vectors were chosen, firstly, to increase<br />

the transduction efficiency, and secondly, because<br />

these vectors can be used to generate transgenic<br />

pigs by injecting them into the perivitelline space<br />

of porcine embryos [7], or transfected primary cells<br />

can be employed in somatic nuclear transfer [19].<br />

Expression of shRNA depends on the number<br />

of integrated vectors per cell and on the promotor<br />

activity. Although the expression cassette of<br />

pLVTHM-pol2 is doubled after lentiviral integration<br />

[32], no significant difference in the knockdown<br />

efficiency could be detected compared to<br />

RRL-pGK-GFP-pol2 (Fig. 3A, B). Due to the<br />

strong polymerase III H1-RNA gene promoter,<br />

which allows ubiquitous expression of shRNAs, a<br />

single shRNA expression cassette per lentiviral vector<br />

might be sufficient to express enough shRNAs<br />

to reduce the PERV mRNA to a minimum. Moreover,<br />

after integration of the lentiviral vector into<br />

the host genome, the knockdown of PERV expres-<br />

sion was stable in primary porcine cells for more<br />

than 170 days (Fig. 3A).<br />

Little is known about the expression of PERV in<br />

different tissues and organs of normal and transgenic<br />

pigs [1]. The data presented here suggest<br />

that the expression of PERV may be inhibited in<br />

different cell types including blood cells, which is<br />

important for the use of cells and organs for<br />

xenotransplantation. Systematic screening of PERV<br />

expression in the first shRNA-transgenic animals<br />

using real-time RT-PCR and of expression of the<br />

shRNA using Northern blot analysis will show<br />

whether PERV is suppressed in all tissues in vivo.<br />

Taken together, the strategy described here could<br />

lead to the generation of transgenic pigs with<br />

greatly reduced PERV expression, providing safe<br />

xenotransplantants, and is therefore of considerable<br />

medical importance.<br />

Acknowledgment<br />

Supported by Deutsche Forschungsgemeinschaft, DFG,<br />

DE729=4. We thank Deborah Mihica for excellent technical<br />

support.<br />

References<br />

1. Akiyoshi DE, Denaro M, Zhu H, Greenstein JL,<br />

Banerjee P, Fishman JA (1998) Identification of a<br />

full-length cDNA for an endogenous retrovirus of miniature<br />

swine. J Virol 72: 4503–4507<br />

2. Denner J, Specke V, Schwendemann J, Tacke SJ (2001)<br />

Porcine endogenous retroviruses (PERVs): adaptation<br />

to human cells and attempts to infect small animals<br />

and non-human primates. Ann Transplant 6: 25–33<br />

3. Ericsson T, Oldmixon B, Blomberg J, Rosa M,<br />

Patience C, Andersson G (2001) Identification of novel<br />

porcine endogenous betaretrovirus sequences in miniature<br />

swine. J Virol 75: 2765–2770<br />

4. Fiebig U, Stephan O, Kurth R, Denner J (2003) Neutralizing<br />

antibodies against conserved domains of p15E<br />

of porcine endogenous retroviruses: basis for a vaccine<br />

for xenotransplantation? Virology 307: 406–413<br />

5. Follenzi A, Ailles LE, Bakovic S, Geuna M, Naldini L<br />

(2000) Gene transfer by lentiviral vectors is limited<br />

by nuclear translocation and rescued by HIV-1 pol<br />

sequences. Nat Genet 25: 217–222<br />

6. Garkavenko O, Croxson MC, Irgang M, Karlas A,<br />

Denner J, Elliott RB (2004) Monitoring for presence<br />

of potentially xenotic viruses in recipients of pig islet<br />

xenotransplantation. J Clin Microbiol 42: 5353–5356


7. Hofmann A, Kessler B, Ewerling S, Weppert M, Vogg<br />

B, Ludwig H, Stojkovic M, Boelhauve M, Brem G, Wolf<br />

E, Pfeifer A (2003) Efficient transgenesis in farm animals<br />

by lentiviral vectors. EMBO Rep 4: 1054–1060<br />

8. Irgang M, Sauer IM, Karlas A, Zeilinger K, Gerlach JC,<br />

Kurth R, Neuhaus P, Denner J (2003) Porcine endogenous<br />

retroviruses: no infection in patients treated with<br />

a bioreactor based on porcine liver cells. J Clin Virol<br />

28: 141–154<br />

9. Karlas A, Kurth R, Denner J (2004) Inhibition of porcine<br />

endogenous retroviruses by RNA interference: increasing<br />

the safety of xenotransplantation. Virology 325: 18–23<br />

10. Karlas A (2006) Porcine endogene Retroviren und<br />

Xenotransplantation: Evaluierung des Infektionsrisikos<br />

und Strategien <strong>zur</strong> Hemmung der Virusreplikation mittels<br />

RNA-Interferenz. Logos, Berlin<br />

11. Kues WA, Petersen B, Mysegades W, Carnwath JW,<br />

Niemann H (2005) Isolation of murine and porcine fetal<br />

stem cells from somatic tissue. Biol Reprod 72(4): 1020<br />

12. Le Tissier P, Stoye JP, Takeuchi Y, Patience C, Weiss RA<br />

(1997) Two sets of human-tropic pig retrovirus. Nature<br />

389: 681–682<br />

13. Livak KJ, Schmittgen TD (2001) Analysis of relative<br />

gene expression data using real-time quantitative<br />

PCR and the 2(-Delta Delta C(T)) Method. Methods<br />

25: 402–408<br />

14. Loss M, Arends H, Winkler M, Przemeck M, Steinhoff<br />

G, Rensing S, Kaup FJ, Hedrich HJ, Winkler ME,<br />

Martin U (2001) Analysis of potential porcine endogenous<br />

retrovirus (PERV) transmission in a whole-organ<br />

xenotransplantation model without interfering microchimerism.<br />

Transpl Int 14: 31–37<br />

15. Martin U, Kiessig V, Blusch JH, Haverich A, von der<br />

Helm K, Herden T, Steinhoff G (1998) Expression of pig<br />

endogenous retrovirus by primary porcine endothelial<br />

cells and infection of human cells. Lancet 352: 692–694<br />

16. Martin U, Winkler ME, Id M, Radeke H, Arseniev L,<br />

Takeuchi Y, Simon AR, Patience C, Haverich A,<br />

Steinhoff G (2000) Productive infection of primary<br />

human endothelial cells by pig endogenous retrovirus<br />

(PERV). Xenotransplantation 7: 138–142<br />

17. McIntyre MC, Kannan B, Solano-Aguilar GI, Wilson<br />

CA, Bloom ET (2003) Detection of porcine endogenous<br />

retrovirus in cultures of freshly isolated porcine bone<br />

marrow cells. Xenotransplantation 10: 337–342<br />

18. Miyagawa S, Nakatsu S, Nakagawa T, Kondo A,<br />

Matsunami K, Hazama K, Yamada J, Tomonaga K,<br />

Miyazawa T, Shirakura R (2005) Prevention of PERV<br />

infections in pig to human xenotransplantation by the<br />

RNA interference silences gene. J Biochem (Tokyo)<br />

137: 503–508<br />

19. Niemann H, Kues WA (2003) Application of transgenesis<br />

in livestock for agriculture and biomedizine. Anim<br />

Reprod Sci 79: 391–317<br />

B. Dieckhoff et al.: Inhibition of PERV expression by RNA interference<br />

20. Oldmixon BA, Wood JC, Ericsson TA, Wilson CA,<br />

White-Scharf ME, Andersson G, Greenstein JL,<br />

Schuurman HJ, Patience C (2002) Porcine endogenous<br />

retrovirus transmission characteristics of an inbred herd<br />

of miniature swine. J Virol 76: 3045–3048<br />

21. Paradis K, Langford G, Long Z, Heneine W, Sandstrom<br />

P, Switzer WM, Chapman LE, Lockey C, Onions D,<br />

Otto E (1999) Search for cross-species transmission of<br />

porcine endogenous retrovirus in patients treated with<br />

living pig tissue. The XEN 111 Study Group. Science<br />

285: 1236–1241<br />

22. Patience C, Takeuchi Y, Weiss RA (1997) Infection of<br />

human cells by an endogenous retrovirus of pigs. Nat<br />

Med 3: 282–286<br />

23. Patience C, Switzer WM, Takeuchi Y, Griffiths DJ,<br />

Goward ME, Heneine W, Stoye JP, Weiss RA (2001)<br />

Multiple groups of novel retroviral genomes in pigs and<br />

related species. J Virol 75: 2771–2775<br />

24. Pfeifer A, Ikawa M, Dayn Y, Verma IM (2002)<br />

Transgenesis by lentiviral vectors: lack of gene silencing<br />

in mammalian embryonic stem cells and preimplantation<br />

embryos. Proc Natl Acad Sci USA 99:<br />

2140–2145<br />

25. Specke V, Rubant S, Denner J (2001) Productive infection<br />

of human primary cells and cell lines with porcine<br />

endogenous retroviruses. Virology 285: 177–180<br />

26. Tacke SJ, Kurth R, Denner J (2000) Porcine endogenous<br />

retroviruses inhibit human immune cell function: risk<br />

for xenotransplantation? Virology 268: 87–93<br />

27. Tacke SJ, Bodusch K, Berg A, Denner J (2001) Sensitive<br />

and specific immunological detection methods for porcine<br />

endogenous retroviruses applicable to experimental<br />

and clinical xenotransplantation. Xenotransplantation 8:<br />

125–135<br />

28. Tacke SJ, Specke V, Denner J (2003) Differences in<br />

release and determination of subtype of porcine endogenous<br />

retroviruses (PERV) produced by stimulated<br />

normal pig blood cells. Intervirology 46: 17–24<br />

29. Takeuchi Y, Patience C, Magre S, Weiss RA, Banerjee<br />

PT, Le Tissier P, Stoye JP (1998) Host range and<br />

interference studies of three classes of pig endogenous<br />

retrovirus. J Virol 72: 9986–9991<br />

30. Wilson CA, Wong S, VanBrocklin M, Federspiel MJ<br />

(2000) Extended analysis of the in vitro tropism of<br />

porcine endogenous retrovirus. J Virol 74: 49–56<br />

31. Winkler ME, Winkler M, Burian R, Hecker J, Loss M,<br />

Przemeck M, Lorenz R, Patience C, Karlas A, Sommer<br />

S, Denner J, Martin U (2005) Analysis of pig-to-human<br />

porcine endogenous retrovirus transmission in a triplespecies<br />

kidney xenotransplantation model. Transpl Int<br />

17: 848–858. Epub 2005 Apr 29<br />

32. Wiznerowicz M, Trono D (2003) Conditional suppression<br />

of cellular genes: lentivirus vector-mediated druginducible<br />

RNA interference. J Virol 77: 8957–8961


Parent-of-origin dependent gene-specific knock down in mouse embryos<br />

Abstract<br />

Khursheed Iqbal, Wilfried A. Kues, Heiner Niemann *<br />

Department of Biotechnology, Institute for Animal Breeding, Mariensee, 31535 Neustadt, Germany<br />

Received 10 April 2007<br />

Available online 7 May 2007<br />

In mice hemizygous for the Oct4-GFP transgene, the F1 embryos show parent-of-origin dependent expression of the marker gene. F1<br />

embryos with a maternally derived OG2 allele (OG2 mat / ) express GFP in the oocyte and during preimplantation development until the<br />

blastocyst stage indicating a maternal and embryonic expression pattern. F1-embryos with a paternally inherited OG2 allele (OG2 pat / )<br />

express GFP from the 4- to 8-cell stage onwards showing only embryonic expression. This allows to study allele specific knock down of<br />

GFP expression. RNA interference (RNAi) was highly efficient in embryos with the paternally inherited GFP allele, whereas embryos<br />

with the maternally inherited GFP allele showed a delayed and less stringent suppression, indicating that the initial levels of the target<br />

transcript and the half life of the protein affect RNAi efficacy. RT-PCR analysis revealed only minimum of GFP mRNA. These results<br />

have implications for studies of gene silencing in mammalian embryos.<br />

Ó 2007 Elsevier Inc. All rights reserved.<br />

Keywords: Oct4-GFP transgene; RNA interference; Maternal inheritance; Paternal inheritance; Short interfering RNA; Preimplantation development;<br />

Gene expression<br />

Development of early mammalian embryos is characterized<br />

by a switch from a maternal to an embryonic expression<br />

program [1]. Maternally derived transcripts and<br />

proteins are deposited in the oocyte and consumed after<br />

fertilisation at least until the embryonic genome is activated.<br />

The RNA binding protein Msy2 has been proposed<br />

to stabilize maternal mRNAs in the oocyte [2].<br />

The timing of the major onset of embryonic genome<br />

transcription is species specific. In mouse embryos it occurs<br />

at the 2-cell stage, in human embryos at the 4-cell stage,<br />

and in bovine embryos at 8–16-cell stage. A ‘‘minor’’ genome<br />

activation has been observed prior to the major genomic<br />

activation. Some genes show a biphasic expression<br />

pattern, in that maternally derived mRNAs and proteins<br />

are present in the oocyte and that the gene are also actively<br />

transcribed after embryonic activation.<br />

It has been shown that RNA interference (RNAi)-mediated<br />

gene knock down works in mammalian embryos for<br />

several genes including E-cadherin, Mos, Plat, Oct4,<br />

* Corresponding author. Fax: +49 5034 871 101.<br />

E-mail address: niemann@tzv.fal.de (H. Niemann).<br />

Biochemical and Biophysical Research Communications 358 (2007) 727–732<br />

0006-291X/$ - see front matter Ó 2007 Elsevier Inc. All rights reserved.<br />

doi:10.1016/j.bbrc.2007.04.155<br />

www.elsevier.com/locate/ybbrc<br />

Cox5a, Cox5b or Cox6b1 [3–9]. However, it is unknown<br />

whether RNAi is equally effective for maternal and embryonic<br />

transcripts. Maternal transcripts have a long half-life<br />

in the range of days, whereas embryonic transcripts last<br />

at most for minutes. In the present study, a breeding schedule<br />

was applied to produce hemizygous embryos carrying<br />

a marker transgene that permits a comparison of gene<br />

silencing of the same allele expressed either maternally or<br />

after embryonic genome activation. The zygotes analysed<br />

in this study were hemizygous for a GFP transgene driven<br />

by the promoter of the POU transcription factor Oct4<br />

[10,11]. It has been previously shown that the Oct4 promoter<br />

activates transcription in early blastomeres from<br />

the 4- to 8-cell stage onward, in the inner cell mass<br />

(ICM) of preimplantation embryos, in the epiblast and<br />

eventually in the gametes [12]. The Oct4-GFP transgene<br />

has been shown to be active in oocytes and both mRNA<br />

and GFP protein are maternally stored in MII oocytes<br />

and zygotes. Terminally differentiated spermatozoa, however,<br />

do not express Oct4. Mating of homozygous OG2<br />

males or females with non-transgenic mice, produced hemizygous<br />

OG2 F1 embryos with either embryonic GFP


728 K. Iqbal et al. / Biochemical and Biophysical Research Communications 358 (2007) 727–732<br />

expression after the 4-cell stage (OG2 pat / ), or zygotes<br />

with maternal deposits of GFP mRNA and protein<br />

(OG2 mat / ), i.e. maternal expression. Thus knock down<br />

of one of two alleles of the target gene located at the same<br />

integration site, sharing identical regulatory and structural<br />

elements, and identical secondary structure of the transcripts<br />

could be studied in two different modes of<br />

expression.<br />

Materials and methods<br />

Production of zygotes hemizygous for the OG2 transgene. OG2 transgenic<br />

mice (tg; Oct4 promoter-GFP) [10,11] were maintained as an inbred<br />

homozygous line by sib matings. For the production of hemizygous OG2<br />

zygotes, OG2 animals were mated with non-transgenic NMRI animals.<br />

OG2 and NMRI females were superovulated by intraperitoneal injections<br />

of 10 IU PMSG (pregnant mare serum gonadotropin, Intervet,<br />

Unterschleißheim, Germany), followed by injections of 10 IU hCG<br />

(human chorion gonadotropin, Intervet) 48 h later and mating with OG2<br />

or NMRI males, respectively. Zygotes were collected by flushing the<br />

oviducts of plugged females 22–24 h after hCG injection with phosphate<br />

buffered saline (PBS) containing 0.3% (w/v) bovine serum albumin (BSA).<br />

Cumulus cells were completely removed by a short treatment with 0.03%<br />

hyaluronidase and zygotes were collected in PBS/0.3% BSA on a warming<br />

plate. Animals were maintained and handled according to German<br />

guidelines.<br />

Injection of siRNA and in vitro culture. The lyophylised siRNAs were<br />

dissolved at a concentration of 20 lM in a buffer consisting of 30 mM<br />

HEPES/KOH (pH 7.4), 100 mM K-acetate, and 2 mM Mg-acetate and<br />

stored frozen at 20 °Cin10ll aliquots. Denuded zygotes were washed in<br />

PBS/0.3% BSA and transferred to a micromanipulation unit in a droplet<br />

of PBS/0.3% BSA. Individual zygotes were fixed by suction to a holding<br />

capillary, approximately 10 pl of siRNA solution was injected into the<br />

cytoplasm using an Eppendorf transjector 5246 (Eppendorf, Hamburg,<br />

Germany). Rhodamine conjugated siRNAs (Qiagen, Hilden, Germany)<br />

with 3 0 -TT overhangs against GFP (upper strand: 5 0 -GCAAGCUGA<br />

CCCUGAAGUUCAU) or with a random sequence (upper strand: 5 0 -UU<br />

CUCCGAACGUGUCACGU) with no known homology to mammalian<br />

genes were used for injection. Successful injections were verified by fluorescence<br />

microscopy.<br />

Fertilized embryos of both inheritance patterns were microinjected<br />

with either the GFP-siRNA or the control random-siRNA, untreated<br />

embryos were used as culture controls. Groups of 10 embryos were<br />

cultured for 4.5 days in 50 ll microdrops of KSOM medium overlaid<br />

with silicone oil at 37 °C in an atmosphere of 5% CO2 in air. For<br />

fluorescence microscopy, a Zeiss Axiovert 35 M microscope equipped<br />

with fluorescence optics for Hoechst 33342, GFP and rhodamine was<br />

used [13]. Images were recorded on Ektachrome 320 T film (Kodak)<br />

with a Contax167MT camera. The GFP fluorescence signals were<br />

normalized by using a fixed exposure time of 15 s. 4-cell stage embryos<br />

and blastocysts were collected for RT-PCR and stored frozen at<br />

80 °C. At the end of the in vitro culture period developmental stage,<br />

nuclei number and GFP mRNA levels were determined. For determination<br />

of the number of nuclei per blastocyst, embryos were stained<br />

with 1 lg/ml Hoechst 33342 for 5 min.<br />

Determination of GFP transcript levels by RT-PCR. Poly(A) + RNA<br />

was isolated from single embryos using a Dynabeads mRNA DIRECT<br />

Kit (Dynal, Norway, Product number 610.11) as described [14]. The<br />

mRNAs were reverse transcribed in a reaction mixture consisting of 1·<br />

RT buffer (50 mM KCl, 10 mM Tris–HCl, pH 8.3, Perkin-Elmer), 50 mM<br />

MgCl2, 1 mM dNTPs (Amersham, Brunswick, Germany), 20 IU RNase<br />

inhibitor (Perkin-Elmer) and 50 U/ll reverse transcriptase (RT, Perkin-<br />

Elmer) in a total volume of 20 ll using 2.5 ll of50lM random hexamer<br />

primers (Perkin-Elmer Vaterstetten, Germany). The RT reaction was<br />

carried out at 25 °C for 10 min and then 42 °C for 1 h followed by<br />

denaturation at 99 °C for 5 min. The polymerase chain reaction (PCR)<br />

was performed with the reverse transcribed product using 0.2 embryo<br />

equivalents. The final volume of the PCR mixture was 50 ll (1· PCR<br />

buffer, 20 mM Tris buffer–HCl (pH 8.4), 1.5 mM MgCl2, 200 lM dNTPs,<br />

0.25 lM of each primer). The GFP primers were 5 0 -AGC ACG ACT TCT<br />

TCA AGT CC and 5 0 -TGA AGT TCA CCT TGA TGC CG (annealing<br />

temperature: 58 °C, 35 cycles). For poly(A) polymerase (also known as<br />

Papola; Mouse Nomenclature Page, www.informatics.jax.org/mgihome)<br />

amplification, the primers were 5 0 -CCT CAC TGC ACA ATA TGC CGT<br />

CTT CAC CTG and 5 0 -AGC CAA TCC ACT GTT CTC CCA CCT TTA<br />

CTC (annealing temperature 54 °C, 35 cycles). The PCR cycle numbers<br />

were optimized to ensure that the reactions are in the linear range of<br />

amplification. PCR amplicons were separated by electrophoresis on a 2%<br />

agarose gel, and were recorded using a CCD camera (Quantix, Photometrics,<br />

Munich Germany). Sequence identities of amplicons were identified<br />

by sequencing.<br />

Results<br />

Hemizygous embryos with a maternally inherited Oct4-<br />

GFP allele (OG2 mat / ) expressed GFP in oocytes and<br />

zygotes; while embryos with a paternally inherited Oct4-<br />

GFP allele initiated transcription of GFP after embryonic<br />

activation and displayed GFP fluorescence at the 4–8-cell<br />

stage (Fig. 1A). Pronuclear stage zygotes of both inheritance<br />

patterns were used for siRNA injection, the siRNAs<br />

were conjugated with rhodamine, thus successful injection<br />

and metabolism of the siRNAs could be followed by fluorescence<br />

microscopy (Fig. 1B and C). The majority of<br />

embryos had reached the blastocyst stage at the end of<br />

the culture period at day 4.5. Apparently the injection procedure<br />

did not reduce the blastocyst rate compared to the<br />

untreated control group (Table 1).<br />

After injection of GFP-siRNA, green GFP fluorescence<br />

was significantly suppressed in embryos with paternally<br />

inherited Oct4-GFP during the entire observation period<br />

(Fig. 2A) and 95% of blastocysts showed down-regulation<br />

of GFP fluorescence (Table 1). On average, the GFP<br />

mRNA levels were reduced by 97.4% in siRNA treated<br />

(OG pat / )-blastocysts as determined by RT-PCR analysis<br />

(Fig. 3A and B).<br />

In contrast, embryos with a maternally inherited GFP<br />

allele (OG2 mat / ) showed a delayed reduction of GFP<br />

fluorescence after siRNA injection. All stages from zygote<br />

to morula continued to display GFP fluorescence; a significant<br />

knock down of GFP fluorescence became was only<br />

apparent at the blastocyst stage (Fig. 2B).<br />

The remaining GFP transcript levels in blastocyst stages<br />

were low and not different between embryos with a maternally<br />

(OG mat / ) or a paternal allele (OG pat / ) inherited<br />

allele (Fig. 3C) indicating that maternal and embryonic<br />

transcripts were equally susceptible to RNAi. To determine<br />

whether maternally derived GFP transcripts were degraded<br />

slower after RNAi injection, OG mat / zygotes were<br />

injected with GFP-siRNA and early 4-cell stages were<br />

analysed by RT-PCR. At this time point, embryonic transcription<br />

of the Oct4-GFP transgene had yet not started.<br />

RT-PCR indicated only minimum levels of GFP transcripts<br />

(data not shown), suggesting that predominantly


Fig. 1. (A) Expression of GFP in F1-embryos with hemizygous OG2 allele from paternal (OG2 pat / ) or maternal (OG2 mat / ) inheritance in zygotes, 2cell<br />

embryos, morulae and blastocysts. (B) Schematic drawing of rhodamine conjugated siRNA. (C) Injection of siRNA-rhodamine in (OG2 mat / )<br />

zygotes. Top, brightfield image; middle, GFP fluorescence; bottom, rhodamine fluorescence in injected zygote and injection capillary.<br />

Table 1<br />

Knock down of GFP in (OG pat / ) and (OG mat / ) mouse embryos<br />

the protein turnover time of GFP contributed to the<br />

delayed down-regulation.<br />

Injection of a random siRNA with no known murine<br />

complementary sequence did not affect GFP fluorescence.<br />

Rhodamine fluorescence was detected in the cytoplasm of<br />

all embryonic stages up to the blastocyst, indicating slow<br />

degradation of the synthetic siRNAs (data not shown).<br />

Discussion<br />

K. Iqbal et al. / Biochemical and Biophysical Research Communications 358 (2007) 727–732 729<br />

No. of<br />

zygotes<br />

No. of morulae with knockdown<br />

of GFP/total (%)<br />

Here, we show that a single injection of a synthetic siR-<br />

NA into the cytoplasm of mouse zygotes is sufficient to<br />

induce allele specific silencing during embryonic development.<br />

Employing hemizygous F1 embryos with parent-oforigin<br />

dependent expression pattern of the OG2 transgene,<br />

allowed discriminating between embryonic and maternal<br />

No. of blastocysts with<br />

knockdown of GFP/total (%)<br />

(OG pat / )<br />

GFP-siRNA injected 210 60/172 (93%) 133/140 (95%)<br />

Random-siRNA<br />

injected<br />

31 0/25 (0%) 0/19 (0%)<br />

Untreated control 60 0/53 (0%) 0/47 (0%)<br />

(OG mat / )<br />

GFP-siRNA injected 94 0/79 (0%) 66/66 (100%)<br />

Untreated control 47 0/39 (0%) 0/35 (0%)<br />

origin of gene expression. Silencing was highly efficient in<br />

embryos carrying a paternally inherited GFP allele<br />

(OG2 pat / ). In this setting the siRNA was injected 2–3 cell<br />

cycles prior to activation of the GFP allele. Nascent GFP<br />

transcripts apparently were targeted and subsequently<br />

degraded. The long-term knock down of GFP in OG2 pat /<br />

blastocysts up to day 4.5 led us to speculate that the duration<br />

of silencing might last to implantation. This suggests that it<br />

would be possible to silence genes during the peri- and<br />

postimplantation period and to dissect critical pathways in<br />

this important period of gestation. Current methodology<br />

to target these stages [15–17] is complicated and time<br />

consuming. It is not yet known at which developmental stage<br />

the unspecific interferon mediated response becomes<br />

prevalent, thus the application of siRNAs might be superior<br />

to the commonly used long double stranded (ds)RNA.


730 K. Iqbal et al. / Biochemical and Biophysical Research Communications 358 (2007) 727–732<br />

Fig. 2. Knock down of GFP expression in hemizygous embryos with either a paternally or a maternally inherited GFP allele. 2-cell stage, morula and<br />

blastocyst stages are shown under brightfield and GFP fluorescence conditions.<br />

Fig. 3. RT-PCR detection of GFP mRNA in siRNA injected embryos. (A) OG pat / embryos: lane 1, untreated embryo; lane 2, untreated, no reverse<br />

transcriptase (RT); lane 3, GFP-siRNA injected; lane 4, no RT; M, 100 bp DNA ladder; lane 5, no template. Poly(A) polymerase was amplified as control<br />

transcript. (B) OG mat / embryos: lane 1, GFP-siRNA injected; lane 2, GFP-siRNA injected, no RT; lane 3, untreated embryo; lane 4, untreated embryo,<br />

no RT; M, 100 bp ladder; lane 5, no template. (C) Knock down efficiency of GFP transcripts in hemizygous OG pat / and OG mat / blastocysts (3<br />

replicates; SD).<br />

RNAi was discovered in nematodes and plants [18,19],<br />

where long dsRNAs of 500–1000 bp specifically suppress<br />

expression of complementary transcripts. In most mammalian<br />

cell types, dsRNA causes a non-specific effect mediated<br />

by an interferon mediated response [20,21]. This non-specific<br />

interferon response is not observed in early embryonic stages<br />

of mammals. It has been shown that specific RNA interference<br />

can be achieved in mammalian embryos by the injection<br />

of long dsRNA [3–6,8]. Processing of the injected dsRNAs<br />

by cellular RNaseIII-like endoribonuclease (Dicer) leads to<br />

the release of several different short interfering (si)RNAs<br />

[22] in the embryo, of which some might be functional as specific<br />

RNAi. Commonly, only 1 out of 4 tested siRNAs gives<br />

specific gene knock down [23,24]. However, off target effects<br />

and siRNA overload [25] cannot be excluded if dsRNA is<br />

used for injection into mammalian embryos.


An important finding of this study is that the onset and<br />

efficiency of gene knock down in embryos critically depend<br />

on whether the target gene is expressed from the maternal<br />

or embryonic genome. In OG2 mat / embryos, which<br />

already carry maternal GFP mRNAs and protein at the<br />

time of injection, knock down was delayed and significant<br />

GFP fluorescence remained up to the 16-cell stage. Efficient<br />

reduction of GFP expression was observed in blastocysts,<br />

al<strong>bei</strong>t with higher residual GFP fluorescence than in siR-<br />

NA treated OG2 pat / embryos. In mouse LA-9 cells the<br />

half-life of GFP was determined to be 26 h [26], thus confirming<br />

that genes with maternal expression can be<br />

silenced, but turnover rates of maternal RNA and protein<br />

may delay the inhibitory effects of RNAi.<br />

Previously, it has been shown that both maternal and<br />

embryonic target genes could be silenced, e.g E-cadherin,<br />

Mos, Plat, Oct4, Cox5a, Cox5b or Cox6b1 [3–7]. However,<br />

due to the unique orchestration of gene expression in mammalian<br />

embryos with a switch from maternal to embryonic<br />

expression, RNAi experiments for transcripts with a biphasic<br />

expression pattern might be obscured. Phenotypic<br />

changes of RNAi in embryos could be masked in cases<br />

when proteins with long half lives or slow turnover rates<br />

are still active, although transcripts have already been<br />

knocked down.<br />

The data presented here provide important clues for gene<br />

silencing in mammalian preimplantation embryos. The<br />

siRNA approach could also be suitable for species other than<br />

mouse in which genomic sequencing is not yet complete [27].<br />

Synthesis of siRNAs requires knowledge of short sequence<br />

regions and the effects of synthetic siRNA are transient,<br />

attributed to degradation of the original molecules, unless<br />

transgenic delivery of siRNA is employed [23]. However, it<br />

seems to be worthwhile to further investigate the stability<br />

of synthetic siRNAs in embryos and to determine their<br />

efficacy in pre- and peri-implantation stages. A potential<br />

application of synthetic siRNA injection in embryos is the<br />

knockdown of gene functions with long term consequences<br />

for the developing organism. For example, it has been shown<br />

that a specific telomere elongation program acts at<br />

morula–blastocyst transition [28], a transient knock down<br />

of telomerase in this critical time window should help to<br />

elucidate the biological relevance of this mechanism.<br />

Acknowledgments<br />

The expert technical assistance of E. Lemme is gratefully<br />

acknowledged. Khursheed Iqbal is in the Ph.D. program of<br />

Hannover Biomedical Research School (HBRS). Khursheed<br />

Iqbal was supported by a grant from Goyaike<br />

S.A.A.C.I. Y F., Buenos Aires, Argentina.<br />

References<br />

K. Iqbal et al. / Biochemical and Biophysical Research Communications 358 (2007) 727–732 731<br />

[1] R.M. Schultz, The molecular foundations of the maternal to zygotic<br />

transition in the preimplantation embryo, Hum. Reprod. Update 8<br />

(2002) 323–331.<br />

[2] J. Yu, N.B. Hecht, R.M. Schutz, Expression of MSY2 in mouse<br />

oocytes and preimplantation embryos, Biol. Reprod. 65 (2001) 1260–<br />

1270.<br />

[3] P. Svoboda, P. Stein, H. Hayashi, R.M. Schultz, Selective reduction<br />

of dormant maternal mRNAs in mouse oocytes by RNA interference,<br />

Development 127 (2000) 4147–4156.<br />

[4] F. Wianny, M. Zernicka-Goetz, Specific interference with gene<br />

function by double-stranded RNA in early mouse development,<br />

Nat. Cell Biol. 2 (2000) 70–75.<br />

[5] B. Plusa, S. Frankenberg, A. Chalmers, A.K. Hadjantonakis, C.A.<br />

Moore, N. Papalopulu, V.E. Papaioannou, D.M. Glover, M.<br />

Zernicka-Goetz, Downregulation of Par3 and aPKC function directs<br />

cells towards the ICM in the preimplantation mouse embryo, J. Cell<br />

Sci. 118 (2004) 505–515.<br />

[6] F. Paradis, C. Vigneault, C. Robert, M.A. Sirard, RNA interference<br />

as a tool to study gene function in bovine oocytes, Mol. Reprod. Dev.<br />

70 (2005) 111–121.<br />

[7] K. Nganvongpanit, H. Müller, B. Rings, M. Gilles, D. Jennen, M.<br />

Hoelker, E. Tholen, K. Schellander, D. Tesfaye, Targeted suppression<br />

of E-cadherin gene expression in bovine preimplantation embryo by<br />

RNA interference technology using double stranded RNA, Mol.<br />

Reprod. Dev. 73 (2005) 153–163.<br />

[8] M.R. Shin, X.S. Cui, J.H. Jun, Y.J. Jeong, N.H. Kim, Identification<br />

of mouse blastocyst genes that are downregulated by double-stranded<br />

RNA-mediated knockdown of Oct4 expression, Mol. Reprod. Dev.<br />

70 (2005) 390–396.<br />

[9] X.S. Cui, X.Y. Li, Y.J. Jeong, J.H. Jun, N.H. Kim, Gene expression<br />

of cox5a, 5b, or 6b1 and their roles in preimplantation mouse<br />

embryos, Biol. Reprod. 72 (2006) 601–610.<br />

[10] R. Anderson, T.K. Copeland, H. Scholer, J. Heasman, C. Wylie, The<br />

onset of germ cell migration in the mouse embryo, Mech. Dev. 91<br />

(2000) 61–68.<br />

[11] P.E. Szabo, K. Huebner, H. Scholer, J.R. Mann, Allele-specific<br />

expression of imprinted genes in mouse migratory primordial germ<br />

cells, Mech. Dev. 115 (2002) 157–160.<br />

[12] Y.I. Yeom, G. Fuhrmann, C.E. Ovitt, A. Brehm, K. Ohbo, M. Gross,<br />

K. Huebner, H.R. Scholer, Germline regulatory element of Oct4<br />

specific for the totipotent cycle of embryonal cells, Development 122<br />

(1996) 881–894.<br />

[13] W.A. Kues, B. Petersen, W. Mysegades, J.W. Carnwath, H.<br />

Niemann, Isolation of murine and porcine fetal stem cells from<br />

somatic tissue, Biol. Reprod. 72 (2005) 1020–1028.<br />

[14] P.S. Yadav, W.A. Kues, D. Herrmann, J.W. Carnwath, H. Niemann,<br />

Bovine ICM derived cells express the Oct4 ortholog, Mol. Reprod.<br />

Dev. 72 (2005) 182–190.<br />

[15] G. Mellitzer, M. Hallonet, L. Chen, S.L. Ang, Spatial and temporal<br />

‘knock downs’ of gene expression by electroporation of dsRNA and<br />

morpholinos into early periimplantation embryos, Mech. Dev. 118<br />

(2002) 57–63.<br />

[16] F. Caligari, A.M. Marzesco, R. Kittler, F. Buchholz, W.B. Huttner,<br />

Tissue specific RNA interference in post-implantation mouse embryos<br />

using directional electroporation and whole embryo culture, Differentiation<br />

72 (2004) 92–102.<br />

[17] M.L. Soares, S. Haraguchi, M.E. Torres-Padilla, T. Kalmar, L.<br />

Carpenter, G. Bell, A. Morrison, C.J.A. Ring, N.J. Clarke, D.M.<br />

Glover, M. Zernicka-Goetz, Functional studies of signalling pathways<br />

in peri-implantation development of the mouse embryo by<br />

RNAi, BMC Dev. Biol. 5 (2005) 1–11.<br />

[18] A. Fire, S. Xu, M.K. Montgomery, S.A. Kostas, S.E. Driver,<br />

C. Mello, Potent and specific genetic interference by doublestranded<br />

RNA in Caenorhabditis elegans, Nature 391 (1998)<br />

809–811.<br />

[19] R.A. Jorgensen, R.G. Atkinson, R.L. Forster, W.J. Lucas, An RNAbased<br />

information superhighway in plants, Science 279 (1998) 1486–<br />

1487.<br />

[20] P.A. Sharp, RNA interference-2001, Genes Dev. 15 (2001) 485–490.<br />

[21] R.H.A. Plasterk, RNA silencing: the genomes immune system,<br />

Science 296 (2002) 1263–1265.


732 K. Iqbal et al. / Biochemical and Biophysical Research Communications 358 (2007) 727–732<br />

[22] S.M. Elbashir, J. Harborth, W. Lendeckel, A. Yalcin, K. Weber,<br />

T. Tuschl, Duplexes of 21-nucleotide RNAs mediate RNA<br />

interference in cultured mammalian cells, Nature 411 (2001)<br />

494–498.<br />

[23] D.A. Rubinson, C.P. Dillon, A.V. Kwiatkowski, C. Sievers, L. Yang,<br />

J. Kopinja, D.L. Rooney, M.M. Ihrig, M.T. McManus, F.B. Gertler,<br />

A lentivirus based system to functionally silence genes in primary<br />

mammalian cells, stem cells and transgenic mice by RNA interference,<br />

Nat. Genet. 33 (2003) 401–406.<br />

[24] D. Gong, J.E. Ferrell Jr., Picking a winner: new mechanistic<br />

insights into the design of effective siRNAs, Trends Biotechnol. 22<br />

(2004) 451–454.<br />

[25] D. Grimm, K.L. Streetz, C.L. Jopling, T.A. Storm, K. Pandey, C.R.<br />

Davis, P. Marion, F. Salazar, M.A. Kay, Fatality in mice due to<br />

oversaturation of cellular microRNA/short hairpin RNA, Nature 441<br />

(2006) 537–541.<br />

[26] P. Corish, C. Tyler-Smith, Attenuation of green fluorescent protein<br />

half-life in mammalian cells, Protein Eng. 12 (1999) 1035–1040.<br />

[27] W.A. Kues, H. Niemann, The contribution of farm animals to human<br />

health, Trends Biotechnol. 22 (2004) 286–294.<br />

[28] S. Schaetzlein, A. Lucas-Hahn, E. Lemme, W.A. Kues, M. Dorsch,<br />

M.P. Manns, H. Niemann, K.L. Rudolph, Telomere length is reset<br />

during early mammalian embryogenesis, Proc. Natl. Acad. Sci. USA<br />

101 (2004) 8034–8038.


CLONING AND STEM CELLS<br />

Volume 9, Number 3, 2007<br />

© Mary Ann Liebert, Inc.<br />

DOI: 10.1089/clo.2006.0009<br />

Production of Viable Pigs from Fetal Somatic Stem Cells<br />

NADINE HORNEN, WILFRIED A. KUES, JOSEPH W. CARNWATH,<br />

ANDREA LUCAS-HAHN, BJÖRN PETERSEN, PETRA HASSEL, and HEINER NIEMANN<br />

ABSTRACT<br />

Fetal somatic stem cells (FSSCs) are a novel type of somatic stem cells that have recently been<br />

discovered in primary fibroblast cultures from pigs and other species. The goal of the present<br />

study was to produce viable piglets from FSSCs. NT complexes were prepared from both<br />

FSSCs and porcine fetal fibroblasts (pFF) to permit comparison of these two donor cell types.<br />

FSSCs from isolated attached colonies were compared with pFF in their ability to form blastocysts<br />

upon use in NT. Fusion and cleavage rates were similar between the two groups, while<br />

blastocyst rates were significantly higher when using pFF as donor cells. FSSCs of three different<br />

size categories derived from dissociation of spheroids yielded similar results. The use<br />

of FSSCs of 15–20 m in size yielded similar cleavage and blastocyst rates as fetal fibroblasts.<br />

In the final experiment NT complexes produced from FSSCs were transferred to foster mothers.<br />

After transfer to prepubertal gilts, three of seven recipients established pregnancies and<br />

delivered seven piglets, of which three piglets were viable and showed normal development.<br />

Results for the first time demonstrate that FSSCs are able to produce cloned embryos, and<br />

that pregnancies can be established and viable piglets can be produced.<br />

INTRODUCTION<br />

SOMATIC CELL NUCLEAR TRANSFER (NT) is a useful<br />

technology for basic research and the production<br />

of transgenic animals. NT permits targeted<br />

modification of the genome of the donor<br />

cells by homologous recombination and production<br />

of genetically identical groups of transgenic<br />

animals (Denning et al., 2001). Since the first somatic<br />

clone “Dolly” (Wilmut et al., 1997), 11 other<br />

mammalian species have been cloned successfully<br />

by the same technique (Baguisi et al., 1999;<br />

Betthauser et al., 2000; Campbell et al., 1996;<br />

Cibelli et al., 1998; Chesne et al., 2002; Galli et al.,<br />

2003; Lee et al., 2005a; Li et al., 2006; Onishi et al.,<br />

2000; Polejaeva et al., 2000; Shin et al., 2002;<br />

Wakayama et al., 1998; Woods et al., 2003; Zhou<br />

et al., 2003). However, despite intensive efforts<br />

Department of Biotechnology, Institut für Tierzucht, Mariensee Neustadt, Germany.<br />

364<br />

cloning, efficiency in most species is still low. In<br />

pigs, on average 1%–3% of reconstructed embryos<br />

survive to birth (Harrison et al., 2004; Kues<br />

and Niemann, 2004). The differentiation status of<br />

the donor cells has been assumed to contribute to<br />

the success of cloning since correct epigenetic reprogramming<br />

and the resulting changes in transcriptional<br />

control are the main processes involved<br />

in creating an embryo from a somatic<br />

nucleus (Jaenisch et al., 2002). Nuclear reprogramming<br />

is a physiological process which takes<br />

place after fertilization and in early cleavage<br />

stages to regulate gene expression during embryogenesis.<br />

The reprogramming includes remodeling<br />

of the chromatin structure, changes in<br />

DNA-methylation, regulation of the expression of<br />

imprinted genes, regeneration of telomere length,<br />

and inactivation of the X-chromosome (Han et al.,


FSSCS IN SOMATIC CELL NUCLEAR TRANSFER 365<br />

2003; Westphal, 2005). In somatic cell nuclear<br />

transfer the chromatin structure of the differentiated<br />

nucleus must be removed and remodeled to<br />

initiate the embryogenic process after fusion of<br />

the donor cell nucleus with the enucleated oocyte<br />

(Gurdon et al., 2003; Westphal, 2005). Presumably,<br />

reprogramming to the embryonic state<br />

should be completed before the onset of embryonic<br />

gene expression (EGA) (Jaenisch et al., 2002).<br />

In bovine development, EGA occurs between the<br />

8- and 16-cell stage; in the pig, it occurs at the<br />

four-cell stage (Lequarre et al., 2003; Maddox-<br />

Hyttel et al., 2001). This additional time for remodeling<br />

could explain the higher cloning efficiency<br />

in bovine nuclear transfer because the<br />

transferred nucleus would have more time to be<br />

reprogrammed.<br />

Stem cells, as well as differentiated cells, have<br />

been successfully used as donor cells, at least to<br />

the extent that they gave development to the blastocyst<br />

stage. Many of these embryos, however,<br />

failed to develop further after transfer to foster<br />

mothers (Kato et al., 2000; Wakayama and Yanagimachi,<br />

2001). The terminally differentiated cells<br />

used include lymphocytes and cardiomyocytes,<br />

but all result in rather low efficiency (Hochedlinger<br />

and Jaenisch, 2002; Inoue et al., 2005;<br />

Schwarzer et al., 2006). In mice the use of embryonic<br />

stem cells (ES cells) increased the efficiency<br />

from 1%–3% to 15% (offspring/transferred<br />

embryos) (Wakayama et al., 1999). True ES cells<br />

have not yet been isolated from livestock (Gjorret<br />

and Maddox-Hyttel, 2005).<br />

Adult stem cells are undifferentiated cells in a<br />

differentiated tissue. They are able to proliferate,<br />

self-renew, and give rise to differentiated daughter<br />

cells. In adult tissue, they are responsible for<br />

regeneration and maintainance of homeostasis.<br />

Many different types of adult tissue harbor stem<br />

cells including bone marrow, brain, epidermis,<br />

blood, liver, skin, eye, gut, pancreas, and skeletal<br />

muscle (Schoeler, 2004). We have recently described<br />

the derivation of fetal somatic stem cells<br />

(FSSCs) from explant cultures of various species<br />

(Kues et al., 2005). These cells do not show signs<br />

of senescence in high density culture, lose fibroblast<br />

markers, and express markers of pluri- or<br />

totipotency-like OCT4, AP, and STAT3. FSSCs<br />

could be more amenable to reprogramming, and<br />

would thus be a new source of donor cells in somatic<br />

cell nuclear transfer or cell therapy approaches<br />

(Kues et al., 2005). However, the isolation<br />

and integration into an existing nuclear<br />

transfer protocol had not yet been investigated,<br />

and was the subject of this research.<br />

MATERIALS AND METHODS<br />

Cell culture<br />

A porcine fetus was obtained by hysterectomy<br />

of a pregnant gilt at day 25. It was washed twice<br />

in Dulbecco phosphate-buffered saline solution<br />

(D-6650, Sigma, St. Louis, MO). Head, legs, and<br />

viscera were removed and remaining tissue was<br />

sliced into 1–2-mm sections. Tissue slices were<br />

placed in drops of recalcified bovine plasma (e.g.,<br />

9 parts bovine plasma, P-4639, Sigma (St. Louis,<br />

MO) 1 part 0.5 M CaCl2) in a tissue culture flask<br />

(#9025, TPP, Switzerland). Supplementation of<br />

plasma with calcium is required for fixation of<br />

tissue. After agglutination of the plasma drops,<br />

DMEM (Dulbecco Modified Eagle Medium)<br />

medium was added (supplemented with 2 mM<br />

glutamine, 1% nonessential amino acids, 0.1 mM<br />

mercaptoethanol, 100 U/mL penicillin, 100<br />

g/mL streptomycin, all from Sigma) 10% fetal<br />

calf serum (FCS, Gibco, Karlsruhe, Germany)],<br />

and somatic explant cultures were incubated in a<br />

humidified 95% air, 5% CO2 atmosphere at 37°C.<br />

Outgrowing cells were trypsinized and subpassaged<br />

twice prior to cryopreservation.<br />

Five to 6 days prior to nuclear transfer, cells<br />

FIG. 1. 3D colonies with surrounding monolayer of fibroblasts<br />

(original magnification 100).


366<br />

FIG. 2. Floating spheroids in suspension culture (original<br />

magnification 100).<br />

were thawed and reseeded on tissue culture test<br />

plates (92406, Fa. TPP, Switzerland) with<br />

DMEM 10% FCS. The cell cycle of the fetal fibroblasts<br />

was synchronized in G0/G1 by serum<br />

deprivation in DMEM with reduced FCS content<br />

(0.5%) 48 h before nuclear transfer. Growth of<br />

FSSCs was induced by DMEM culture medium<br />

with increased serum content (30%). FSSCs were<br />

maintained either as attached cells, in which case<br />

3D colonies formed on the confluent culture (Fig.<br />

1), or in a suspension culture which permits the<br />

growth of free floating spheroids (Fig. 2). For nuclear<br />

transfer, 3D colonies were isolated from surrounding<br />

fibroblasts by aspirating with a 100 L<br />

pipette (Fig. 3), floating spheroids were pelleted<br />

and washed with PBS. Cells were dissociated<br />

by incubation for 10–15 min in 0.025% EDTA/<br />

trypsin, pelleted, washed once in TL-HEPES containing<br />

10% FCS, once in serum-free TL-HEPES,<br />

and resuspended in TL-HEPES.<br />

Collection and culture of<br />

cumulus–oocyte complexes<br />

Ovaries from prepubertal gilts were collected<br />

from local abattoirs. Oocytes with at least three<br />

complete layers of cumulus cells were collected<br />

and matured in vitro for 40 h (Hoelker et al., 2005).<br />

Then, matured oocytes were transferred to TL-<br />

HEPES supplemented with 0.1% hyaluronidase<br />

and incubated for 5 min. Cumulus cells were re-<br />

moved by repeated pipetting. Denuded oocytes<br />

were transferred to TL-HEPES covered with mineral<br />

oil (Silicone fluid DC200, #35135, Fa. Serva<br />

Biochemica, Heidelberg).<br />

Nuclear transfer<br />

Denuded oocytes were enucleated by removing<br />

the first polar body and the metaphase plate<br />

under UV-light control after Hoechst 33342 staining<br />

of the oocytes. A single donor cell was placed<br />

into the perivitelline space (Hoelker et al., 2005).<br />

Oocyte donor cell complexes were placed between<br />

two platinum electrodes and fused with<br />

one DC pulse of 1 kV/cm for 99 sec (Multiporator,<br />

Eppendorf, Hamburg). Oocyte-donor cell<br />

complexes and 50–60 nonmanipulated oocytes<br />

(parthenogenetic controls) were activated in an<br />

activation chamber (Bügel-Fusionskammer, Fa.<br />

Krüss, Hamburg) with one DC pulse of 1 kV/cm<br />

for 45 sec (Hoelker et al., 2005). Then the complexes<br />

were activated chemically by exposure<br />

to NCSU23 with 2 mM 6-dimethylaminopurin<br />

for 3 h.<br />

Culture of embryos to blastocysts<br />

HORNEN ET AL.<br />

After activation up to 60 reconstructed embryos<br />

and parthenogenetic controls were each<br />

placed in 500 L NCSU23 in one well of a fourwell<br />

dish. Embryos were cultured for 7 days in<br />

an incubation chamber (Nuclear Incubation<br />

Chamber; Billups-Rothenberg Inc., Del Mar, CA)<br />

at 38°C with 5% CO2, 5% O2, and 90% N2. After<br />

7 days of in vitro culture, embryos were stained<br />

with 1 g/mL Hoechst 33342 to determine the<br />

number of nuclei using a fluorescence microscope.<br />

To determine the number of inner cell mass<br />

cells and trophectoderm cells, embryos were dif-<br />

FIG. 3. Isolation of 3D colonies by aspiration.


FSSCS IN SOMATIC CELL NUCLEAR TRANSFER 367<br />

ferentially stained. Blastocysts were incubated in<br />

a permeabilizing solution containing the ionic detergent<br />

Triton X-100 and the fluorochrome propidium<br />

iodide to stain trophectoderm cells for 50<br />

sec. Thereafter blastocysts were incubated in a<br />

second solution containing 100% ethanol for fixation<br />

and the fluorochrome Hoechst 33342 for<br />

staining the inner cell mass cells for 4 min. Fixed<br />

and stained blastocysts were analyzed with a fluorescence<br />

microscope at 100 magnification<br />

(Thouas et al., 2001).<br />

Embryo transfers<br />

Embryos cloned from FSSCs were surgically<br />

transferred into the oviducts of gilts that were<br />

synchronized to the reconstructed embryos<br />

(Walker et al., 2002), immediately after activation.<br />

Maintenance of pregnancies was supported by induction<br />

of accessory corpora lutea injecting 1000<br />

IU PMSG injected i.m. at day 12 after embryo<br />

transfer (Pregnant Mare’s Serum Gonadotropin,<br />

Intergonan ® i.m., Intervet, Tönisvorst) and 500 IU<br />

hCG (human Chorionic Gonadotropin, Ovogest ® ,<br />

Intervet, Tönisvorst) 72 h later. Recipients were<br />

examined ultrasonographically at days 22 and 35<br />

following embryo transfer. Piglets were delivered<br />

on days 116–119 after embryo transfer. Animal<br />

handling was in accordance with institutional<br />

guidelines.<br />

Microsatellite analysis<br />

Parentage was analyzed in piglets derived<br />

from somatic cell nuclear transfer to confirm genetic<br />

identity of the piglets with the donor cells<br />

used for somatic cell nuclear transfer. The microsatellite<br />

analysis was performed as descibed<br />

before (Hoelker et al., 2005). Eleven porcine DNA<br />

microsatellite markers were used (S0002, S0068,<br />

S0090, S0101, S0178, SW69, SW 632, SW 857, SW<br />

911, SW 951, and SW1122).<br />

Experimental design<br />

The first three experiments were designed to<br />

determine which FSSC preparation was most<br />

suitable to nuclear transfer based on in vitro culture<br />

to the blastocyst stage. The assay was to compare<br />

the development of these FSSC derived embryos<br />

to that of embryos reconstructed from<br />

porcine fetal fibroblasts which our laboratory<br />

uses routinely for nuclear transfer. In the first experiment,<br />

FSSCs from isolated attached colonies<br />

were compared with fetal fibroblasts as donor<br />

cells. In the second experiment, FSSCs of different<br />

size categories, that is, 14 m, 15–20 m,<br />

23–28 m, isolated from spheroids were compared.<br />

In the third experiment, blastocysts were<br />

cloned from medium size FSSCs (15–20 m) derived<br />

from spheroids and compared with that of<br />

fetal fibroblasts. In the final experiment, cloned<br />

zygotes produced from FSSC nuclear donor cells<br />

were transferred to foster mothers. Spheroids<br />

were not serum deprived prior to use in nuclear<br />

transfer. The different cell size of spheroids does<br />

not correspond with different cell cycle stages,<br />

but is thought to be related to the differentiation<br />

status of stem cells.<br />

Statistical analysis<br />

Fusion rate, the cleavage rate, the blastocyst<br />

rate, the number of nuclei in the blastocysts, and<br />

the percentage of ICM cells compared to the total<br />

cell number of the blastocysts were analyzed<br />

by “Sigma Stat” (Jandel Scientific Coorporation,<br />

Germany). Results are based on 4–10 replicates<br />

per experiment, and are expressed as means<br />

(mean SD). A probability value of p 0.05 was<br />

considered to be statistically significant.<br />

RESULTS<br />

Experiment 1<br />

In the experiment to determine whether<br />

FSSCs removed from selected three-dimensional<br />

colonies by trypsinization were better nuclear<br />

donors than primary fibroblasts, the blastocyst<br />

rate of FSSC reconstituted embryos was significantly<br />

lower than the blastocyst rate of the fibroblast<br />

group (6.4% 3.5% vs. 24.9% 8.6%).<br />

Parthenogenetic embryos used as control for development<br />

without nuclear transfer gave a blastocyst<br />

rate of 44.3% 9.1%. Cleaveage rates and<br />

the mean cell number of the blastocysts in all<br />

three groups were not different (Table 1). In this<br />

experiment, a wide distribution of sizes was observed<br />

within the trypsinized population of<br />

FSSCs.<br />

Experiment 2<br />

In this experiment to determine whether the<br />

size of FSSC donor cells was an important factor,<br />

FSSCs of different sizes were isolated from spher-


368<br />

oids and used as donor cells. A comparison of the<br />

blastocyst rates, cleavage rates, and mean cell<br />

number from the three different cell sizes and for<br />

the parthenogenetic controls is shown in Table 2.<br />

There was no significant difference which could<br />

be attributed to the size of the donor cells and the<br />

blastocyst rate for the parthenogenetic controls<br />

was normal (33.2% 8.8%).<br />

Experiment 3<br />

In the experiment to determine whether FSSC<br />

donor cells derived from spheroids were better<br />

or worse than fibroblasts, there was no significant<br />

difference between the two cell types with respect<br />

to blastocyst rate, cleavage rate, or mean number<br />

of cells in the blastocysts (Table 3). Differential<br />

staining, revealed no significant differences between<br />

the two cell types with respect to the number<br />

of ICM cells, the number of TE cells, total<br />

number of cells, or the number of cells in the ICM<br />

as a proportion of the total cell number (Table 4).<br />

Experiment 4<br />

TABLE 1. BLASTOCYST RATES, CLEAVAGE RATES, AND MEAN CELL NUMBER OF BLASTOCYSTS<br />

OF EMBRYOS DERIVED FROM 3D COLONIES AND FETAL FIBROBLASTS<br />

Cleavage rate Blastocyst rate Cell number of blastocysts<br />

Experimental groups n (%) (%) (mean SD)<br />

FSSCs 138 64.8 17.3 6.4 3.5 a 32 80<br />

Fetal fibroblasts 166 82.5 5.60 24.9 8.6 b 31 13<br />

Parthenogenetic 356 81.6 11.7 44.3 9.1 b 32 11<br />

a p 0.05, six replications.<br />

In this experiment to evaluate the quality of reconstituted<br />

embryos by their capacity for in vivo<br />

development, embryos cloned from FSSCs were<br />

transferred to synchronized foster mothers, and<br />

at 25 days three of the seven recipients were pregnant.<br />

A second ultrasound examination at 32 days<br />

showed that these three animals were still preg-<br />

nant. The eighth recipient died shortly after surgery<br />

for reasons unrelated to the experiment and<br />

was not included in the analysis. The three sows<br />

which remained pregnant delivered a total of<br />

seven female piglets. Three of the piglets survived<br />

and showed normal development (Figs. 4<br />

and 5). Microsatellite analysis confirmed that the<br />

piglets had been derived from the FSSCs. The<br />

other four piglets had normal birthweights and<br />

no obvious deformaties. A detailed pathological<br />

examination suggested that they had developed<br />

normally.<br />

DISCUSSION<br />

HORNEN ET AL.<br />

Successful nuclear transfer depends on correct<br />

and complete epigenetic reprogramming of the<br />

donor cell nucleus to convert the genome of a differentiated<br />

somatic cell into the totipotent state<br />

(Gurdon and Colman, 1999). It has been suggested<br />

that a less differentiated nuclear donor<br />

cell could be more efficiently reprogrammed<br />

(Jaenisch et al., 2002) and, indeed, in the mouse,<br />

cloning from pluripotent embryonic stem cells resulted<br />

in higher rates of development than<br />

cloning with differentiated donor cells (Wakayama<br />

et al., 1999). Embryonic stem cells have not<br />

yet been isolated from livestock species; however,<br />

we recently reported a cell culture system that enriches<br />

a subpopulation of FSSCs in the pig, the<br />

TABLE 2. BLASTOCYST RATES, CLEAVAGE RATES, AND MEAN CELL NUMBER OF EMBRYOS<br />

RECONSTRUCTED WITH SPHEROID FSSCS OFDIFFERENT SIZE CLASSES<br />

Cleavage rate Blastocyst rate Cell number of blastocysts<br />

Experimental groups n (%) (%) (mean SD)<br />

Small’ FSSCs 145 83.5 7.8 15.3 7.9 32 13<br />

Medium’ FSSCs 135 83.1 1.6 17.6 6.8 34 13<br />

Large’ FSSCs 142 83.2 5.8 10.4 2.7 31 16<br />

Parthenogenetic 259 88.8 8.9 33.2 8.8 34 11<br />

controls<br />

Five replications.


FSSCS IN SOMATIC CELL NUCLEAR TRANSFER 369<br />

TABLE 3. BLASTOCYST RATES, CLEAVAGE RATES, AND MEAN CELL NUMBER OF BLASTOCYSTS<br />

OF EMBRYOS DERIVED FROM SPHEROID FSSCS AND FETAL FIBROBLASTS<br />

Cleavage rate Blastocyst rate Cell number of blastocysts<br />

Experimental groups n (%) (%) (mean SD)<br />

FSSCs 348 80.7 5.9 18.4 5.6 35 11<br />

Fetal fibroblasts 322 86.1 6.7 21.4 5.6 37 13<br />

Parthenogenetic 586 84.2 7.3 36.1 5.8 34 11<br />

Ten replications.<br />

cow, and the mouse. These FSSCs exhibit many<br />

of the characteristics of pluripotent ES cells (Kues<br />

et al., 2005), including absence of senescence in a<br />

high-density culture, loss of typical fibroblast<br />

markers, and expression of typical stem cell<br />

markers such as OCT4, AP, and STAT3. Injection<br />

of mouse FSSCs into the blastocoel resulted in<br />

chimerism in fetal mesenchymal organs, and occasionally<br />

labeled cells were identified in the genital<br />

ridge (Kues et al., 2005). We hypothesized that<br />

FSSCs would require less reprogramming than<br />

more differentiated cell types, and could evolve<br />

as a novel source of donor cells in somatic nuclear<br />

transfer.<br />

In the present study, we report the birth of the<br />

first cloned piglets from somatic stem cells and<br />

demonstrate the capacity of FSSCs for in vivo development.<br />

After transferring 70–100 reconstituted<br />

embryos to each recipient, three of seven<br />

gilts (42.9%) became pregnant. Harrison et al.<br />

(2004) reported a higher efficiency with fetal fibroblasts<br />

and recent experience in our laboratory<br />

is similar.<br />

An early observation of the FSSC population<br />

in culture revealed different cellular morphologies<br />

depending on culture conditions. Before going<br />

directly to the in vivo experiments, it was necessary<br />

to evaluate the developmental competence<br />

of these different classes of FSSC. The FSSCs<br />

could be grown as three dimensional colonies by<br />

allowing them to grow first as a high density culture<br />

attached to a dish. The FSSCs could also be<br />

grown as spheroid cultures by suspension cul-<br />

ture. In both cases it was necessary to trypsinize<br />

the cell aggregates to obtain single cells for nuclear<br />

transfer, and in both cases, the trypsinization<br />

resulted in a range of cell sizes.<br />

In removing three-dimensional colonies from<br />

the culture dish prior to EDTA/trypsin treatment,<br />

it was impossible to avoid some degree of<br />

contamination by the fibroblasts which form a<br />

monolayer on the surface of the dish and provide<br />

a substrate for colony growth. Trypsinization of<br />

these colonies released some fibroblasts into the<br />

single cell suspension which must be avoided<br />

when the cells were to be used as nuclear donor<br />

cells. Spheroid cultures did not have this disadvantage,<br />

and thus were expected to provide pure<br />

FSSC samples. Another problem with using the<br />

three-dimensional cultures as sources of donor<br />

cells was the extended period of exposure to the<br />

EDTA/trypsin treatment which was necessary to<br />

give single-cell suspensions. Indeed, one explanation<br />

for the low blastocyst rate in the first experiment<br />

could be attributed to the stress of<br />

trypsin isolation on these donor cells. This idea is<br />

supported by the fact that many damaged cells<br />

could be observed in these preparations. In subsequent<br />

experiments, in addition to changing to<br />

suspension cultured FSSC cells, we also were able<br />

to shorten the EDTA/trypsin treatment.<br />

When using fetal fibroblasts as NT donor cells,<br />

it is desirable to select small round cells which<br />

have been demonstrated to be typical for the<br />

G0/G1 phase of the cell cycle (Boquest et al.,<br />

1999). Fetal fibroblast cultures have a greater<br />

TABLE 4. RESULTS OF DIFFERENTIAL STAINING OF THE BLASTOCYSTS<br />

WITH FSSCS ORFETAL FIBROBLASTS AS DONOR CELLS<br />

Number of Number of ICM cells/total<br />

Experimental Blastocysts ICM cells TE cells Total cell number cell number<br />

groups n (mean SD) (mean SD) (m) (%)<br />

FSSCs 24 16 5 22 12 39 14 41.7 14.3<br />

Fetal fibroblasts 25 17 8 24 10 39 11 42.3 17.3<br />

Six replications.


370<br />

FIG. 4. The three vital piglets 1 day after birth.<br />

number of cells with this favoured morphology<br />

after 48 h of serum deprivation. Unfortunately,<br />

cultures of FSSCs, both 3D colonies and spheroids,<br />

are heterogeneous with regard to cell size,<br />

and it is necessary to make a choice when selecting<br />

cells for nuclear donation. The FSSCs of<br />

“medium” size corresponded to the size of the fetal<br />

fibroblasts. Interestingly, the results of the second<br />

experiment showed no significant difference<br />

in development which could be related to donor<br />

cell size. This indicates that cell size is not an appropriate<br />

selection criterion for NT with FSSCs.<br />

Having determined that a more homogeneous<br />

and more viable population of FSSCs could be<br />

isolated from suspension produced spheroids<br />

and having learned that cell size was not an important<br />

trait for selection, the next step was to directly<br />

compare fibroblasts and FSSCs with respect<br />

to the developmental potential of embryos produced<br />

from these two types of nuclear donor<br />

cells. It is obvious that medium-size FSSCs isolated<br />

from spheroids have the same developmental<br />

potential after nuclear transfer as fetal fibroblasts<br />

(20%). For comparison, the blastocyst<br />

rates reported with various types of differentiated<br />

porcine nuclear donor cells range from<br />

10.3% to 37% (Boquest et al., 2002; Hyun et al.,<br />

2003; Kim et al., 2006; Lee et al., 2003a, 2003b,<br />

2005b, 2005c). It is difficult to compare the results<br />

from different laboratories due to the many<br />

methodological differences with regard oocyte<br />

HORNEN ET AL.<br />

source (in vivo vs. in vitro maturation, young or<br />

adult sows, slaughterhouse derived or fresh<br />

ovaries, etc.). In some instances, there is no visual<br />

control of enucleation success, which means that<br />

enucleation may not have been achieved in all instances<br />

(Boquest et al., 2002; Hyun et al., 2003; Lee<br />

et al., 2003b). Lee et al. (2003b) injected donor cells<br />

into the cytoplasm of oocytes and obtained a blastocyst<br />

rate of 37%. Other laboratories have looked<br />

at various types of less differentiated donor cells.<br />

For example, dedifferentiated preadipocytes<br />

gave a blastocyst rate of 20.8% (Tomii et al., 2005).<br />

Zhu et al. (2004) obtained blastocyst rates of 25%<br />

with porcine skin-derived stem cells (PSOS),<br />

Bosch et al. (2006) used porcine mesenchymal<br />

stem cells (MSC) as donors and reported a blastocyst<br />

rate of 4.1% and Sung et al. (2006) used<br />

long-term and short-term repopulating hematopoietic<br />

stem cells and hematopoietic progenitor<br />

cells as donors and reported morula/blastocyst<br />

rates of 4.1%–10.6%.<br />

The present experiments used a nuclear transfer<br />

protocol which had been optimized for fetal<br />

fibroblasts. FSSCs are different from fibroblasts<br />

in that they exhibit several different morphologies<br />

depending on culture conditions. This presumably<br />

reflects physiological differences as<br />

well, so it is likely that a nuclear transfer protocol<br />

specifically optimized for FSSC would give increased<br />

the cloning efficiency.<br />

Comparison of the quality of fetal fibroblast<br />

and FSSC derived cloned embryos (i.e., total cell<br />

number and the ratio of ICM cells to the total cell<br />

number) showed that there was no significant difference<br />

between the two types of donor cell. The<br />

total cell numbers for both groups were similar<br />

to cell counts previously reported (Park et al.,<br />

2001a, 2001b, 2001c, 2002). In general, total cell<br />

numbers of in vivo produced blastocysts or IVP<br />

FIG. 5. Piglets at age of 2.5 months.


FSSCS IN SOMATIC CELL NUCLEAR TRANSFER 371<br />

embryos are higher than those reported for<br />

cloned embryos. This may be due to the fact that<br />

optimal culture conditions have not yet been developed<br />

for cloned porcine embryos (Niemann<br />

and Rath, 2001). Indeed, the current status of<br />

porcine preimplantation embryo culture is far behind<br />

that of bovine or mouse embryo culture<br />

where significantly higher blastocyst rates are<br />

routine. We have achieved a degree of improvement<br />

in that the proportion of ICM cells to the total<br />

number of cells in the blastocysts was higher<br />

than previously reported for cloned embryos<br />

(Koo et al., 2004). Normal fetal development requires<br />

an adequate number of both ICM and TE<br />

cells for implantation (Tam, 1988). Bovine cloned<br />

blastocysts often show a higher total cell number<br />

and a higher proportion of ICM cells than IVF or<br />

in vivo developed embryos (Koo et al., 2002). A<br />

distortion in this ratio can be associated with malformations<br />

of the placenta and early fetal loss<br />

(Koo et al., 2002). In our in vivo experiment, we<br />

did not observe malformations of the placentas<br />

or abortions. In contrast to bovine cloning experiments,<br />

the number of TE cells was normal and<br />

the number of ICM cells was increased (Koo et<br />

al., 2004). The increase of ICM cells was observed<br />

for both FSSCs and fetal fibroblasts. These embryos<br />

still do not match the total cell number or<br />

the ICM/TE ratios of normal embryos developing<br />

in vivo, but the increase in ICM cells indicates<br />

that progress is <strong>bei</strong>ng made toward developing<br />

an optimized somatic nuclear transfer protocol.<br />

Stem cell markers including Oct4 and alkaline<br />

phosphatase (AP) should facilitate separation of<br />

FSSC from contaminating fibroblasts (Yeom et al.,<br />

1996; Kues et al., 2005). One possibility for improving<br />

the selection of FSSC donor cells could<br />

be to transfect the population with a gene construct<br />

consisting of Oct4 as a promotor and an appropriate<br />

flurorescent reporter gene or even an<br />

antibiotic resistance gene. This would enable one<br />

to isolate pure FSSCs with high efficiency. Another<br />

possibility would be the use of antibodies<br />

against stem cell specific surface markers either<br />

to visualize FSSCs or to purify them by FACS.<br />

Human CD133 (formerly AC133) is one such<br />

pluripotent stem cells marker (Miraglia et al.,<br />

1997; Yin et al., 1997); however, its presence on<br />

FSSCs has not yet been shown.<br />

In conclusion, results of this study show that<br />

FSSCs have at least the same in vitro developmental<br />

potential as fetal fibroblasts using an established<br />

nuclear transfer protocol and are able<br />

to generate vital piglets following embryo transfer.<br />

Improved selection and handling of the<br />

FSSCs could increase the efficiency.<br />

ACKNOWLEDGMENTS<br />

This work was supported by the H. Wilhelm<br />

Schaumann Foundation.<br />

REFERENCES<br />

Baguisi, A., Behboodi, E., Melican, D.T., et al. (1999). Production<br />

of goats by somatic cell nuclear transfer. Nat.<br />

Biotechnol. 17, 456–461.<br />

Betthauser, J., Forsberg, E., Augenstein, M., et al. (2000).<br />

Production of cloned pigs from in vitro systems. Nat.<br />

Biotechnol. 18, 1055–1059.<br />

Boquest, A.C., Day, B.N., and Prather, R.S. (1999). Flow<br />

cytometric cell cycle analysis of cultured porcine fetal<br />

fibroblast cells. Biol. Reprod. 60, 1013–1019.<br />

Boquest, A.C., Grupen, C.G., Harrison, S.J., et al. (2002).<br />

Production of cloned pigs from cultured fetal fibroblast<br />

cells. Biol. Reprod. 66, 1283–1287.<br />

Bosch, P., Pratt, S.L., and Stice, S.L. (2006). Isolation, characterization,<br />

gene modification, and nuclear reprogramming<br />

of porcine mesenchymal stem cells. Biol. Reprod.<br />

74, 46–57.<br />

Campbell, K.H.S., McWhir, J., Ritchie, W.A., et al. (1996).<br />

Sheep cloned by nuclear transfer from a cultured cell<br />

line. Nature 380, 64–66.<br />

Chesne, P., Adenot, P.G., Viglietta, C., et al. (2002). Cloned<br />

rabbits produced by nuclear transfer from adult somatic<br />

cells. Nat. Biotechnol. 20, 366–369.<br />

Cibelli, J.B., Stice, S.L., Golueke, P.J., et al. (1998). Cloned<br />

transgenic calves produced from nonquiescent fetal fibroblasts.<br />

Science 280, 1256–1258.<br />

Denning, C., Dickinson, P., Burl, S., et al. (2001). Gene targeting<br />

in primary fetal fibroblasts from sheep and pig.<br />

Cloning Stem Cells 3, 221–231.<br />

Galli, C., Lagutina, I., Crotti, G., et al. (2003). A cloned<br />

horse born to its dam twin—a birth announcement calls<br />

for a rethink on the immunological demands of pregnancy.<br />

Nature 424, 635.<br />

Gjorret, J.O., and Maddox-Hyttel, P. (2005). Attempts towards<br />

derivation and establishment of bovine embryonic<br />

stem cell-like cultures. Reprod. Fertil. Dev. 17, 113–124.<br />

Gurdon, J.B., and Colman, A. (1999). The future of<br />

cloning. Nature 402, 743–746.<br />

Gurdon, J.B., Byrne, J.A., and Simonsson, S. (2003). Nuclear<br />

reprogramming and stem cell creation. Proc. Natl.<br />

Acad. Sci. USA 100, 11819–11822.<br />

Han, Y.M., Kang, Y.K., Koo, D.B., et al. (2003). Nuclear reprogramming<br />

of cloned embryos produced in vitro.<br />

Theriogenology 59, 33–44.<br />

Harrison, S., Boquest, A., Grupen, C., et al. (2004). An efficient<br />

method for producing alpha(1,3)-galactosyl-


372<br />

transferase gene knockout pigs. Cloning Stem Cells 6,<br />

327–331.<br />

Hochedlinger, K., and Jaenisch, R. (2002). Monoclonal<br />

mice generated by nuclear transfer from mature B and<br />

T donor cells. Nature 415, 1035–1038.<br />

Hoelker, M., Petersen, B., Hassel, P., et al. (2005). Duration<br />

of in vitro maturation of recipient oocytes affects<br />

blastocyst development of cloned porcine embryos.<br />

Cloning Stem Cells 7, 35–44.<br />

Hyun, S.H., Lee, G.S., Kim, D.Y., et al. (2003). Effect of<br />

maturation media and oocytes derived from sows or<br />

gilts on the development of cloned pig embryos. Theriogenology<br />

59, 1641–1649.<br />

Inoue, K., Wakao, H., Ogonuki, N., et al. (2005). Generation<br />

of cloned mice by direct nuclear transfer from natural<br />

killer T cells. Curr. Biol. 15, 1114–1118.<br />

Jaenisch, R., Eggan, K., Humpherys, D., et al. (2002). Nuclear<br />

cloning, stem cells, and genomic reprogramming.<br />

Cloning Stem Cells 4, 389–396.<br />

Kato, Y., Tani, T., and Tsunoda, Y. (2000). Cloning of<br />

calves from various somatic cell types of male and female<br />

adult, newborn and fetal cows. J. Reprod. Fertil.<br />

120, 231–237.<br />

Kim, H.S., Lee, G.S., Kim, J.H., et al. (2006). Expression of<br />

leptin ligand and receptor and effect of exogenous leptin<br />

supplement on in vitro development of porcine embryos.<br />

Theriogenology 65, 831–844.<br />

Koo, D.B., Kang, Y.K., Choi, Y.H., et al. (2002). Aberrant<br />

allocations of inner cell mass and trophectoderm cells<br />

in bovine nuclear transfer blastocysts. Biol. Reprod. 67,<br />

487–492.<br />

Koo, D.B., Kang, Y.K., Park, J.S., et al. (2004). A paucity<br />

of structural integrity in cloned porcine blastocysts produced<br />

in vitro. Theriogenology 62, 779–789.<br />

Kues, W.A., and Niemann, H. (2004). The contribution of<br />

farm animals to human health. Trends Biotechnol. 22,<br />

286–294.<br />

Kues, W.A., Petersen, B., Mysegades, W., et al. (2005). Isolation<br />

of murine and porcine fetal stem cells from somatic<br />

tissue. Biol. Reprod. 72, 1020–1028.<br />

Lee, G.S., Kim, H.S., Hyun, S.H., et al. (2003a). Improved<br />

developmental competence of cloned porcine embryos<br />

with different energy supplements and chemical activation.<br />

Mol. Reprod. Dev. 66, 17–23.<br />

Lee, J.W., Wu, S.C., Tian, X.C., et al. (2003b). Production<br />

of cloned pigs by whole-cell intracytoplasmic microinjection.<br />

Biol. Reprod. 69, 995–1001.<br />

Lee, B.C., Kim, M.K., Jang, G., et al. (2005a). Dogs cloned<br />

from adult somatic cells. Nature 436, 641.<br />

Lee, G.S., Kim, H.S., Hyun, S.H., et al. (2005b). Effect of<br />

epidermal growth factor in preimplantation development<br />

of porcine cloned embryos. Mol. Reprod. Dev. 71,<br />

45–51.<br />

Lee, G.S., Kim, H.S., Hyun, S.H., et al. (2005c). Production<br />

of transgenic cloned piglets from genetically transformed<br />

fetal fibroblasts selected by green fluorescent<br />

protein. Theriogenology 63, 973–991.<br />

Lequarre, A.S., Marchandise, J., Moreau, B., et al. (2003).<br />

Cell cycle duration at the time of maternal zygotic transition<br />

for in vitro produced bovine embryos: effect of<br />

HORNEN ET AL.<br />

oxygen tension and transcription inhibition. Biol. Reprod.<br />

69, 1707–1713.<br />

Li, Z., Sun, X., Chen, J., et al. (2006). Cloned ferrets produced<br />

by somatic cell nuclear transfer. Dev. Biol. 293,<br />

439–448.<br />

Maddox-Hyttel, P., Dinnyes, A., Laurincik, J., et al. (2001).<br />

Gene expression during pre- and peri-implantation embryonic<br />

development in pigs. Reprod. Suppl. 58, 175–<br />

189.<br />

Miraglia, S., Godfrey, W., Yin, A.H., et al. (1997). A novel<br />

five-transmembrane hematopoietic stem cell antigen:<br />

isolation, characterization, and molecular cloning.<br />

Blood 90, 5013–5021.<br />

Niemann, H., and Rath, D. (2001). Progress in reproductive<br />

biotechnology in swine. Theriogenology 56, 1291–1304.<br />

Onishi, A., Iwamoto, M., Akita, T., et al. (2000). Pig<br />

cloning by microinjection of fetal fibroblast nuclei. Science<br />

289, 1188–1190.<br />

Park, K.W., Cheong, H.T., Lai, L.X., et al. (2001a). Production<br />

of nuclear transfer-derived swine that express<br />

the enhanced green fluorescent protein. Anim. Biotechnol.<br />

12, 173.<br />

Park, K.W., Kuhholzer, B., Lai, L.X., et al. (2001b). Development<br />

and expression of the green fluorescent protein<br />

in porcine embryos derived from nuclear transfer of<br />

transgenic granulosa-derived cells. Anim. Reprod. Sci.<br />

68, 111–120.<br />

Park, K.W., Lai, L.X., Cheong, H.T., et al. (2001c). Developmental<br />

potential of porcine nuclear transfer embryos<br />

derived from transgenic fetal fibroblasts infected with<br />

the gene for the green fluorescent protein: comparison<br />

of different fusion/activation conditions. Biol. Reprod.<br />

65, 1681–1685.<br />

Park, K.W., Lai, L.X., Cheong, H.T., et al. (2002). Mosaic<br />

gene expression in nuclear transfer-derived embryos<br />

and the production of cloned transgenic pigs from earderived<br />

fibroblasts. Biol. Reprod. 66, 1001–1005.<br />

Polejaeva, I.A., Chen, S.H., Vaught, T.D., et al. (2000).<br />

Cloned pigs produced by nuclear transfer from adult<br />

somatic cells. Nature 407, 86–90.<br />

Schoeler, H.R. (2004). [The potential of stem cells. A<br />

status update]. Bundesgesundheitsblatt. Gesundheitsforschung.<br />

Gesundheitsschutz. 47, 565–577.<br />

Schwarzer, M., Carnwath, J.W., Lucas-Hahn, A., et al.<br />

(2006). Isolation of bovine cardiomyocytes for reprogramming<br />

studies based on nuclear transfer. Cloning<br />

Stem Cells 8, 150–158.<br />

Shin, T., Kraemer, D., Pryor, J., et al. (2002). A cat cloned<br />

by nuclear transplantation. Nature 415, 859.<br />

Sung, L.Y., Gao, S., Shen, H., et al. (2006). Differentiated<br />

cells are more efficient than adult stem cells for cloning<br />

by somatic cell nuclear transfer. Nat. Genet. 38, 1323–<br />

1328.<br />

Tam, P.P.L. (1988). Postimplantation development of<br />

mitomycin-C-treated mouse blastocysts. Teratology 37,<br />

205–212.<br />

Thouas, G.A., Korfiatis, N.A., French, A.J., et al. (2001).<br />

Simplified technique for differential staining of inner<br />

cell mass and trophectoderm cells of mouse and bovine<br />

blastocysts. Reprod. Biomed. Online. 3, 25–29.


FSSCS IN SOMATIC CELL NUCLEAR TRANSFER 373<br />

Tomii, R., Kurome, M., Ochiai, T., et al. (2005). Production<br />

of cloned pigs by nuclear transfer of preadipocytes<br />

established from adult mature adipocytes. Cloning<br />

Stem Cells 7, 279–288.<br />

Wakayama, T., and Yanagimachi, R. (2001). Mouse<br />

cloning with nucleus donor cells of different age and<br />

type. Mol. Reprod. Dev. 58, 376–383.<br />

Wakayama, T., Perry, A.C.F., Zuccotti, M., et al. (1998). Fullterm<br />

development of mice from enucleated oocytes injected<br />

with cumulus cell nuclei. Nature 394, 369–374.<br />

Wakayama, T., Rodriguez, I., Perry, A.C.F., et al. (1999).<br />

Mice cloned from embryonic stem cells. Proc. Natl.<br />

Acad. Sci. USA 96, 14984–14989.<br />

Walker, S.C., Shin, T., Zaunbrecher, G.M., et al. (2002). A<br />

highly efficient method for porcine cloning by nuclear<br />

transfer using in vitro-matured oocytes. Cloning Stem<br />

Cells 4, 105–112.<br />

Westphal, H. (2005). Restoring stemness. Differentiation<br />

73, 447–451.<br />

Wilmut, I., Schnieke, A.E., McWhir, J., et al. (1997). Viable<br />

offspring derived from fetal and adult mammalian<br />

cells. Nature 385, 810–813.<br />

Woods, G.L., White, K.L., Vanderwall, D.K., et al. (2003).<br />

A mule cloned from fetal cells by nuclear transfer. Science<br />

301, 1063.<br />

Yeom, Y.I., Fuhrmann, G., Ovitt, C.E., et al. (1996).<br />

Germline regulatory element of Oct-4 specific for the<br />

totipotent cycle of embryonal cells. Development 122,<br />

881–894.<br />

Yin, A.H., Miraglia, S., Zanjani, E.D., et al. (1997). AC133,<br />

a novel marker for human hematopoietic stem and progenitor<br />

cells. Blood 90, 5002–5012.<br />

Zhou, Q., Renard, J.P., Le Friec, G., et al. (2003). Generation<br />

of fertile cloned rats by regulating oocyte activation.<br />

Science 302, 1179.<br />

Zhu, H., Craig, J.A., Dyce, P.W., et al. (2004). Embryos derived<br />

from porcine skin-derived stem cells exhibit enhanced<br />

preimplantation development. Biol. Reprod. 71,<br />

1890–1897.<br />

Address reprint requests to:<br />

Heiner Niemann<br />

Department of Biotechnology<br />

Institut für Tierzucht<br />

D-31535 Neustadt, Germany<br />

E-mail: niemann@tzv.fal.de


CSIRO PUBLISHING<br />

Reproduction, Fertility and Development, 2007, 19, 762–770 www.publish.csiro.au/journals/rfd<br />

Transgenic farm animals: an update<br />

Heiner Niemann A,B and Wilfried A. Kues A<br />

A Department of Biotechnology, Institute for Animal Breeding, Mariensee, 31535 Neustadt, Germany.<br />

B Corresponding author. Email: niemann@tzv.fal.de<br />

Abstract. The first transgenic livestock species were reported in 1985. Since then microinjection of foreign DNA<br />

into pronuclei of zygotes has been the method of choice. It is now <strong>bei</strong>ng replaced by more efficient protocols based on<br />

somatic nuclear transfer that also permit targeted genetic modifications. Lentiviral vectors and small interfering ribonucleic<br />

acid (siRNA) technology are also becoming important tools for transgenesis. In 2006 the European Medicines Agency<br />

(EMEA) gave green light for the commercialistion of the first recombinant protein produced in the milk of transgenic<br />

animals. Recombinant antithrombin III will be launched as ATryn for prophylactic treatment of patients with congenital<br />

antithrombin deficiency. This important milestone will boost the research activities in farm animal transgenesis. Recent<br />

developments in transgenic techniques of farm animals are discussed in this review.<br />

Additional keywords: conditional transgene expression, gene pharming, lentiviral mediated transgenesis, pronuclear<br />

DNA injection, RNA interference, somatic cloning, xenotransplantation.<br />

Introduction<br />

The first transgenic livestock were produced in 1985 via microinjection<br />

of foreign DNA into pronuclei of zygotes (Hammer<br />

et al. 1985). As microinjection has several significant shortcomings,<br />

including random integration into the host genome, low<br />

efficiency and frequent incidence of mosaicism, research has<br />

focussed on alternate methodologies for improving the generation<br />

of transgenic livestock (Kues and Niemann 2004; Table 1).<br />

These include sperm mediated DNA transfer (Gandolfi 1998;<br />

Squires 1999; Chang et al. 2002; Lavitrano et al. 2002), the<br />

intracytoplasmic injection (ICSI) of sperm heads carrying foreign<br />

DNA (Perry et al. 1999), injection or infection of oocytes<br />

or embryos by viral vectors (Haskell and Bowen 1995; Chan<br />

et al. 1998; Hofmann et al. 2003; Whitelaw et al. 2004) or the<br />

use of nuclear transfer (Schnieke et al. 1997; Cibelli et al. 1998;<br />

Baguisi et al. 1999). Further qualitative improvements may be<br />

derived from technologies that allow targeted modifications of<br />

the genome (Dai et al. 2002; Lai et al. 2002), or conditional<br />

instead of a constitutive expression of transgenes (Kues et al.<br />

2006). Other important developments are the usage of large<br />

genomic constructs to optimise expression, the application of<br />

RNA interference to knock down specific genes and the use of<br />

episomal vectors that might allow a more consistent expression<br />

owing to the absence of position effects (Manzini et al. 2006).<br />

Pharmaceutical production in the mammary gland<br />

of transgenic animals<br />

Gene ‘pharming’ entails the production of recombinant human<br />

pharmaceutical proteins in the mammary gland of transgenic<br />

animals.This technology can overcome the limitations of current<br />

recombinant production systems for pharmaceutical proteins<br />

(Meade et al. 1999; Rudolph 1999). The mammary gland is<br />

the preferred production site mainly because of the quantities<br />

of protein that can be produced by employing various mammary<br />

gland-specific promoter and the ease of production and<br />

purification (Meade et al. 1999; Rudolph 1999).<br />

Numerous proteins have been produced at large amounts<br />

in the mammary gland of transgenic sheep, goat, cattle, pigs<br />

and rabbits and have been advanced to clinical trials (Kues and<br />

Niemann 2004). Phase III trials for antithrombin III (ATIII)<br />

(ATryn from GTC Biotherapeutics, Framingham, MA, USA)<br />

produced in the mammary gland of transgenic goats have been<br />

completed and the recombinant product has been approved as<br />

a drug by the European Medicines Agency (EMEA) in August<br />

2006. The protein is expected to be on the market as a fully<br />

registered drug in 2007. Successful drug registration of ATryn<br />

demonstrated the usefulness and solidity of this approach and<br />

will accelerate registration of further products from this process,<br />

as well as stimulate research and commercial activity in<br />

this area. Other recombinant proteins from the udder are recombinant<br />

C1 inhibitor (Pharming BV, Leiden, The Netherlands)<br />

produced in the milk of transgenic rabbits, α-antitrypsin (α-AT)<br />

and tissue plasminogen activator (tPA), which have all progressed<br />

to advanced clinical trials (Kues and Niemann 2004).<br />

The enzyme α-glucosidase (Pharming BV) from the milk of<br />

transgenic rabbits has orphan drug registration and has been<br />

successfully used for the treatment of Pompe’s disease (Van den<br />

Hout et al. 2001). An overview of the current list of recombinant<br />

pharmaceutical proteins advanced state to commercial exploitation<br />

is given in Table 2 (Echelard et al. 2006). The use of somatic<br />

nuclear transfer accelerates production of transgenic animals<br />

with high mammary gland specific synthesis of recombinant<br />

proteins (Fig. 1).<br />

Another production site for recombinant products is the blood<br />

of transgenic animals.Transgenic cattle were cloned that produce<br />

© CSIRO 2007 10.1071/RD07040 1031-3613/07/060762


Transgenic farm animals: an update Reproduction, Fertility and Development 763<br />

Table 1. Milestones (live offspring) in transgenesis and somatic cloning in farm animals<br />

Modified from Niemann et al. 2005<br />

Year Milestone Strategy Reference<br />

1985 First transgenic sheep and pigs Microinjection of DNA into one pronucleus of a zygote Hammer et al. 1985<br />

1986 Embryonic cloning of sheep Nuclear transfer using embryonic cells as donor cells Willadsen 1986<br />

1997 Cloning of sheep with somatic donor cells Nuclear transfer using adult somatic donor cells Wilmut et al. 1997<br />

1997 Transgenic sheep produced by nuclear transfer Random integration of the construct Schnieke et al. 1997<br />

1998 Transgenic cattle produced from fetal fibroblasts and Random integration of the construct Cibelli et al. 1998<br />

nuclear transfer<br />

1998 Generation of transgenic cattle by MMLV injection Infection of oocytes with helper viruses Chan et al. 1998<br />

2000 Gene targeting in sheep Gene replacement and nuclear transfer McCreath et al. 2000<br />

2002 Trans-chromosomal cattle Artificial chromosome Kuroiwa et al. 2002<br />

2002 Knockout in pigs Heterozygous knockout Dai et al. 2002, Lai et al. 2002<br />

2003 Homozygous gene knockout in pigs Homozygous knock-out Phelps et al. 2003<br />

2003 Transgenic pigs via lentiviral injection Gene transfer into zygotes via lentiviruses Hofmann et al. 2003<br />

2006 Conditional transgene expression in pigs (tet-off) Pronuclear DNA injection and crossbreeding Kues et al. 2006<br />

Table 2. List of therapeutic proteins produced in the milk of transgenic animal that are currently in<br />

commercial development<br />

HAE, Hereditary angioedema; AFP, α-fetoprotein; HGH, human growth hormone; Mab, monoclonal antibody.<br />

Modified from Echelard et al. 2006<br />

Products Companies Production Developmental stage<br />

animal<br />

ATryn (recombinant human GTC Biotherapeutics Goat EU: EMEA approved<br />

antithrombin III) US: Phase 3<br />

C1 Esterase Inhibitor Pharming Rabbit Phase 3 for HAE<br />

MM-093 (AFP) Merrimack and GTC Biotherapeutics Goat Phase 2 for RA<br />

Alpha-Glucosidase Pharming Rabbit Phase 2, on hold<br />

HGH BioSidus Cow Preclinical<br />

Albumin GTC Biotherapeutics Cow Preclinical<br />

Fibrinogen Pharming Cow Preclinical<br />

Collagen Pharming Cow Preclinical<br />

Lactoferrin Pharming Cow Preclinical<br />

Alpha-1 Antitrypsin GTC Biotherapeutics Goat Preclinical<br />

Malaria vaccine GTC Biotherapeutics Goat Preclinical<br />

CD 137 (4–1BB) MAb GTC Biotherapeutics Goat Preclinical<br />

a recombinant bi-specific antibody in their blood. This antibody<br />

is stable after purification from serum and active in mediating<br />

target-cell-restricted T-cell stimulation and tumour cell killing<br />

(Grosse-Hovest et al. 2004).<br />

To ensure safety of recombinant products, guidelines developed<br />

by the Food and Drug Administration (FDA) of the USA<br />

require monitoring of the animals’ health, validation of the gene<br />

construct, characterisation of the isolated recombinant protein,<br />

as well as performance of the transgenic animals over several<br />

generations. This has been taken into account when developing<br />

the ‘gene pharming’ concept, for example by using exclusively<br />

animals from prion-disease-free countries (New Zealand) and<br />

keeping the production animals under strict hygienic conditions.<br />

Porcine-to-human xenotransplantation<br />

The progress in organ transplantation technology has led to an<br />

acute shortage of appropriate organs, and cadaveric or live organ<br />

donation can by far not meet the demand in western societies.<br />

To close the growing gap between demand and availability of<br />

appropriate organs, porcine xenografts from domesticated pigs<br />

are considered the solution of choice (Bach 1998; Platt and<br />

Lin 1998; Kues and Niemann 2004). Essential prerequisites for<br />

successful xenotransplantation are: (1) overcoming the immunological<br />

hurdles; (2) prevention of transmission of pathogens from<br />

the donor animal to the human recipient; and (3) compatibility<br />

of the donor organs with the human organ in terms of anatomy<br />

and physiology.<br />

The main strategies that have been successfully explored for<br />

a long-term suppression of the first immunological hurdle, i.e.<br />

the hyperacute rejection (HAR) of porcine xenografts, are the<br />

synthesis of human complement regulatory proteins (RCAs) in<br />

transgenic pigs (Cozzi and White 1995; Bach 1998; Platt and Lin<br />

1998) and/or the knockout of antigenic structures on the surface<br />

of the porcine organ that cause HAR, i.e. the α-gal-epitopes<br />

(Phelps et al. 2003; Kuwaki et al. 2005; Yamada et al. 2005).<br />

Survival rates after transplantation of porcine RCA transgenic


764 Reproduction, Fertility and Development H. Niemann and W. A. Kues<br />

Generate construct/cell lines<br />

Nuclear transfer<br />

F0 Female goat births<br />

F0 Induced lactation<br />

F0 Natural lactation<br />

Research material<br />

Pre-clinical material<br />

Phase 1 material<br />

F1 Females/male births<br />

F1 Females natural lactation<br />

F2 Females natural lactaton<br />

Bridging/Phase 3 material<br />

DNA 0.5 yr 1 yr 1.5 yr 2 yr 2.5 yr 3 yr 3.5 yr 4 yr<br />

Fig. 1. Timeline for the production of recombinant proteins in the mammary gland of transgenic goats via somatic<br />

cloning (modified from Echelard et al. 2006).<br />

hearts or kidneys to immune-suppressed non-human primates<br />

ranged from 23 to 135 days. Overall, current data show that HAR<br />

can be overcome in a clinically acceptable manner by expressing<br />

human complement regulators in transgenic pigs (Bach 1998).<br />

The usefulness of porcine organs with a knockout for the<br />

1,3-α-galactosyltransferase (α-gal) gene in a xenotransplantation<br />

situation has been demonstrated. Survival rates reached 2–6<br />

months upon transplantation of porcine hearts or kidneys from<br />

α-gal-knockout pigs to baboons (Kuwaki et al. 2005; Yamada<br />

et al. 2005). However, donor pigs with multiple transgenes critical<br />

in other immunological barriers are thought to be required<br />

for a prolonged survival of >6 months of porcine organs in nonhuman<br />

primates. We have recently produced triple transgenic<br />

pigs to suppress the coagulatory disorder frequently found after<br />

porcine-to-primate xenotransplantation by expressing human<br />

thrombomodulin (hTM) or human heme-oxygenase-1 (hHO-1)<br />

on top of two RCAs (Petersen et al. 2006). A consistent survival<br />

of porcine xenografts for more than 6 months in non-human primate<br />

recipients is thought to be necessary for starting clinical<br />

trials with human patients.<br />

Recent research has revealed that the risk of porcine endogenous<br />

retrovirus (PERV) transmission to human patients is low,<br />

paving the way for preclinical testing of xenografts (Switzer et al.<br />

2001; Irgang et al. 2003). RNA interference (RNAi) is a promising<br />

approach to knock down PERV expression in porcine cells<br />

(Dieckhoff et al. 2007), which could then be used in nuclear<br />

transfer. Guidelines for the clinical application of porcine xenotransplants<br />

are already available in the USA and are currently<br />

<strong>bei</strong>ng developed in several other countries. The compatibility<br />

between porcine and human organs, remains to a large extent an<br />

enigma, but preliminary information regarding the function of<br />

porcine kidneys and heart in non-human primates are promising<br />

in this respect (Ibrahim et al. 2006).<br />

Transgenic animals in agriculture<br />

Agricultural applications of transgenic animals are lagging<br />

behind the biomedical area (Kues and Niemann 2004). Table 3<br />

shows an overview on the various attempts to produce animals<br />

transgenic for agricultural traits. An important step towards the<br />

production of healthier pork has been made by the first pigs<br />

transgenic for a desaturase gene derived either from spinach<br />

or Caenorhabditis elegans. These pigs produced a higher ratio<br />

of polyunsaturated versus saturated fatty acids in their muscles<br />

clearly indicating the potential of rendering more healthier pork<br />

in the near future (Niemann 2004; Saeki et al. 2004; Lai et al.<br />

2006).A human diet rich in non-saturated fatty acids is correlated<br />

with a reduced risk of stroke and coronary diseases. In the pig, it<br />

has been shown that transgenic expression of a bovine lactalbumin<br />

construct in sow milk resulted in higher lactose contents and<br />

greater milk yields, which correlated with a better survival and<br />

development of the piglets (Wheeler et al. 2001). The increased<br />

survival of piglets at weaning provides significant benefits to<br />

animal welfare and the producer. Phytase transgenic pigs have<br />

been developed to address the problem of manure-based environmental<br />

pollution. These pigs carry a bacterial phytase gene<br />

under the transcriptional control of a salivary gland specific<br />

promoter, which allows the pigs to digest plant phytate. Without<br />

the bacterial enzyme, the phytate phosphorus passes undigested<br />

into manure and pollutes the environment. With the bacterial<br />

enzyme, the fecal phosphorus output was reduced up to 75%<br />

(Golovan et al. 2001). These environmentally friendly pigs are<br />

expected to enter commercial production chains within the next<br />

few years.<br />

The physicochemical properties of milk are mainly affected<br />

by the ratio of casein variants making them a prime target for the<br />

improvement of milk composition. The bovine casein ratio can<br />

be altered by overexpression of β- and κ-casein clearly underpinning<br />

the potential for improvements in the functional properties<br />

of bovine milk (Brophy et al. 2003). Mastitis is a preeminent<br />

health problem in modern dairy cattle production causing significant<br />

economic losses. Lysostaphin has been shown to confer<br />

specific resistance against mastitis caused by Staphylococcus<br />

aureus. Cows have been cloned from transgenic donor cells that<br />

express a lysostaphin gene construct in the mammary gland,


Transgenic farm animals: an update Reproduction, Fertility and Development 765<br />

Table 3. Overview on successful transgenic livestock for agricultural production<br />

Transgenic trait Key molecule Construct Gene transfer method Species Reference<br />

Increased growth rate, less body fat Growth hormone (GH) hMT-pGH Microinjection Pig Nottle et al. 1999<br />

Increased growth rate, less body fat Insulin-like growth factor-1 (IGF-1) mMT-hIGF-1 Microinjection Pig Pursel et al. 1999<br />

Increased level of poly-un-saturated Desaturase (from spinach) maP2-FAD2 Microinjection Pig Saeki et al. 2004<br />

fatty acids in pork<br />

Increased level of poly-un-saturated Desaturase (from C. elegans) CAGGS-hfat-1 Somatic cloning Pig Lai et al. 2006<br />

fatty acids in pork<br />

Phosphate metabolism Phytase PSP-APPA Microinjection Pig Golovan et al. 2001<br />

Milk composition (lactose increase) α-lactalbumin genomic bovine Microinjection Pig Wheeler et al. 2001<br />

α-lactalbumin<br />

Influenza resistance Mx protein mMx1-Mx Microinjection Pig Müller et al. 1992<br />

Enhanced disease resistance IgA α,K-α,K Microinjection Pig, sheep Lo et al. 1991<br />

Wool growth Insulin-like growth factor-1 (IGF-1) Ker-IGF-1 Microinjection Sheep Damak et al. 1996a,<br />

1996b<br />

Visna Virus resistance Visna virus envelope visna LTR-env Microinjection Sheep Clements et al. 1994<br />

Ovine prion locus Prion protein (PrP) targeting vector Somatic cloning Sheep Denning et al. 2001<br />

(homol. recomb.) (animals dead<br />

shortly after birth)<br />

Milk fat composition Stearoyl desaturase β-lactogl-SCD Microinjection Goat Reh et al. 2004<br />

Milk composition (increase of β-casein κ-casein genomic CSN2 Somatic cloning Cattle Brophy et al. 2003<br />

whey proteins) CSN-CSN-3<br />

Milk composition (increase of human lactoferrin α-s2cas-mLF Microinjection Cattle Platenburg et al. 1994<br />

lactoferrin)<br />

Staph. aureus mastitis resistance Lysostaphin ovine β-lactogl- Somatic cloning Cattle Wall et al. 2005<br />

lysostaphin<br />

rendering the animals mastitis-resistant (Wall et al. 2005). The<br />

advent to detailed genomic maps and appropriate tools for the<br />

genetic engineering of livestock species will accelerate and<br />

enhance transgenic animal production for specific production<br />

purposes.<br />

Transgenic pets<br />

As can be seen above, transgenic approaches have focussed on<br />

livestock species either for biomedical or agricultural purposes.<br />

Recent applications of transgenic technology relate to the development<br />

of new varieties of ornamental fish. Transgenic medaka<br />

(Oryzias latipes, rice fish) with a fluorescent green colour were<br />

described in 2001 (Chou et al. 2001) and subsequently approved<br />

for sale in Taiwan. The fluorescent medaka is currently marketed<br />

by the Taiwanese company Taikong (www.azoo.com.tw).<br />

GloFish is a trademarked transgenic zebra fish (Danio rerio)<br />

expressing a red fluorescent protein from a sea anemone under<br />

the transcriptional control of a muscle-specific promoter (Gong<br />

et al. 2003). In addition, green and yellow fluorescent proteins<br />

have been introduced into stable transgenic lines, resulting<br />

in fishes with different flurorescence (Fig. 2). Yorktown Technologies<br />

(www.glofish.com) started commercial sales of the<br />

transgenic zebrafish in the USA at retail prices of approximately<br />

US$5.00. Thus, GloFish has become the first transgenic animal<br />

readily available in the USA.A recent report of the FDA revealed<br />

no evidence that Glofish represent a risk (US FDA 2003). Commercialisation<br />

of fluorescent fish has gone forward in several<br />

countries other than the USA, such as Taiwan, Malaysia, and<br />

Hong Kong, whereas marketing in Australia, Canada and the<br />

European Union is prohibited.<br />

Transgenic application in other pet species, such as dog or cat,<br />

is hypothetical at present, but may become more widespread<br />

when more efficient gene transfer technologies are developed<br />

and specific genetic traits can be targeted. Some of the emerging<br />

technologies relevant in this context are described below.<br />

Emerging technologies<br />

Lentiviral transgenesis<br />

Lentiviruses have been shown to be efficient vectors for the delivery<br />

of genes into oocytes and zygotes. Injection of lentiviruses<br />

into the perivitelline space of porcine zygotes resulted in a high<br />

proportion of transgenic offspring, from which expressing transgenic<br />

lines could be established (Hofmann et al. 2003; Whitelaw<br />

et al. 2004). The production of transgenic cattle via lentiviral<br />

vectors required the injection into the perivitelline space of<br />

oocytes (Hofmann et al. 2004). Lentiviral gene transfer in livestock<br />

is compatible with unprecedented high rates of transgenic<br />

animals. Whether multiple integration of lentiviral vectors into<br />

the host genome is associated with unwanted side effects, such<br />

as oncogene activation or insertional mutagenesis, remains to be<br />

investigated. However, the silencing of lentiviral sequences in<br />

transgenic lines (Hofmann et al. 2006) and the high frequency<br />

of mosaicism in founder animals requires further research.<br />

Conditional transgenesis in farm animals<br />

Recently, we reported the first tetracycline-controlled transgene<br />

expression in a farm animal (Kues et al. 2006). An autoregulative<br />

tetracycline-responsive bicistronic expression cassette<br />

(NTA) was introduced into the pig genome via pronuclear DNA<br />

injection (Fig. 3a). The NTA cassette was designed to give


766 Reproduction, Fertility and Development H. Niemann and W. A. Kues<br />

Fig. 2. Transgenic zebrafishes expressing different fluorescent proteins. The image was taken under natural white<br />

light (reproduced with permission from www.glofish.com). Owing to high expression of the fluorescent proteins<br />

the fishes appear brightly coloured.<br />

ubiquitous expression of human RCAs independent of cellular<br />

transcription cofactors. Expression from this construct could be<br />

inhibited reversibly by feeding the animals doxycycline (tet-off<br />

system). In 10 transgenic pig lines carrying one NTA cassette,<br />

the transgene was silenced in all tissues with the exception of<br />

skeletal muscle, where the transgene was expressed in discrete<br />

muscle fibres (Niemann and Kues 2003). Interestingly, crossbreeding<br />

to produce animals with two NTA cassettes resulted in<br />

reactivation of the silenced NTA cassettes and a strong and broad<br />

tissue-independent, tetracycline-sensitive RCA expression pattern<br />

(Fig. 3b). The extent of reactivation in the double transgenic<br />

pigs correlated inversely with the methylation status of the NTA<br />

cassettes, and probably reflected effects of the different integration<br />

sites. This shows that selective crossbreeding of transgenic<br />

pig lines can overcome epigenetic silencing.This approach highlights<br />

the importance to study epigenetic (trans)-gene regulation<br />

in the pig (Fig. 3c).<br />

Artificial chromosomes as transgene vectors<br />

Artificial chromosomes have the potential to carry very large<br />

pieces of DNA that are maintained as episomal entities (Niemann<br />

and Kues 2003). A human artificial chromosome (HAC) containing<br />

the entire sequences of the human immunoglobulin<br />

heavy and light chain loci has been introduced into bovine<br />

fibroblasts, which were then used in nuclear transfer. Transchromosomal<br />

bovine offspring were obtained that expressed<br />

human immunoglobulin in their blood. This system could be<br />

a significant step forward in the production of human therapeutic<br />

polyclonal antibodies (Kuroiwa et al. 2002). Follow-up<br />

studies showed that the HAC was maintained in most animals<br />

over several years in the first generation cattle (Robl et al. 2007).<br />

Whether HACs separate correctly during meiotic cell divisions<br />

remains to be shown.<br />

Spermatogonial transgenesis<br />

Transplantation of primordial germ cells into the testes is<br />

an alternative approach to generate transgenic animals. Initial<br />

experiments in mice showed that the depletion of endogenous<br />

spermatogenonial stem cells by treatment with the chemical<br />

busulfan before transplantation is effective and compatible with<br />

re-colonisation by donor cells. Transmission of the donor haplotype<br />

to the next generation after germ-cell transplantation has<br />

been achieved in goats (Honaramooz et al. 2003). Current major<br />

obstacles of this strategy are the lack of efficient in vitro culture<br />

methods for primordial germ/prospermatogonial cells and the<br />

lack of efficient gene transfer techniques into these cells.<br />

RNA interference mediated gene knock down<br />

RNA interference is a conserved post-transcriptional gene<br />

regulatory process in most biological systems, including<br />

fungi, plants and animals. The common mechanistic elements<br />

are double stranded small interfering RNAs (siRNA) with<br />

19–23 nucleotides, which specifically bind to complementary<br />

sequences of their target mRNAs. The target mRNAs are then<br />

degraded by endonuclease activity and no protein is translated<br />

(Plasterk 2002). RNA interference is involved in gene regulation<br />

and specifically to control/suppress the translation of mRNAs<br />

from endogenous and exogenous viral elements and can be used<br />

for therapeutic purposes (Dallas and Vlassow 2006).<br />

For a transient gene knock down, synthetic siRNAs are transfected<br />

in cells or early embryos (Clark and Whitelaw 2003; Iqbal<br />

et al. 2007). For stable gene repression the siRNA sequences<br />

must be incorporated into a gene construct. The combination of<br />

siRNA with the lentiviral vector technology is a highly effective<br />

tool in this respect. RNAi knockdown of PERV has been<br />

shown in porcine primary cells (Dieckhoff et al. 2007), as<br />

well as the knockdown of the prion protein gene (PRNP) in


Transgenic farm animals: an update Reproduction, Fertility and Development 767<br />

(a)<br />

(b)<br />

(c)<br />

Genotype<br />

L13<br />

L15<br />

2.3 kb<br />

Closed<br />

(heterochromatin)<br />

Open<br />

(euchromatin)<br />

RCA<br />

Transactivator<br />

ptTA RCA IRES tTA pA<br />

Inactivated by<br />

exogenous<br />

tetracycline<br />

(AAA) n<br />

1 2 3 4 5 6 7 8 9<br />

<br />

<br />

DNA demethylation<br />

Histone acetylation<br />

Transcription<br />

DNA methylation<br />

Histone deacetylation<br />

Fig. 3. Conditional expression of transgenes in pigs and epigenetic regulation.<br />

(a) Design of an autoregulative cassette, encoding a regulator of<br />

complement activation (RCA) and a tetracycline sensitive transactivator<br />

(tTA), separated by an internal ribosomal entry sequence (IRES). Basal<br />

expression of the construct results in the translation of tTA and subsequently<br />

in autoregulative up-regulation, by binding of tTA to the promoter (ptTA).<br />

Exogenous tetracycline binds and inactivates the transactivator (tet-off),<br />

resulting in down-regulation of transcription. (b) Reactivation of silenced<br />

transgenes by line crossbreeding. Animals of lines 13 and 15, which are<br />

hemizygous carriers for one autoregulative cassette, were crossbred to obtain<br />

double transgenic offspring. Specific detection of RCA mRNA levels by<br />

Northern blotting of muscle biopsies from nine siblings. Animals that carried<br />

one cassette, either from line 13 (L13) or line 15 (L15), expressed low<br />

levels of the transgene (lane 1, 3, 4, 6 and 9); these low expression levels<br />

were also found in the parent animals. Siblings that carried both cassette<br />

(lane 2, 5, 7 and 8) showed 50–100 times higher transcript levels. (c) Epigenetic<br />

regulation of gene expression. Simplified diagram showing that DNA<br />

regions may switch between a closed and an open confirmation. Epigenetic<br />

modification of the DNA by methylation of CG-dinucleotids and of the histones<br />

by acetylation contribute to these changes and may fix the DNA in<br />

one mode. In the open confirmation (euchromatin) the DNA is accessible<br />

to DNA binding protein, like transcription factors and RNA polymerase II,<br />

and transcription can be activated. In the closed mode (heterochromatin) a<br />

gene locus is silenced.<br />

cattle embryos (Golding et al. 2006). Germline transmission of<br />

lentiviral siRNA has been shown in rats over three generations<br />

(Tenenhaus Dann et al. 2006). In contrast to knock out technologies,<br />

which require time consuming breeding strategies, RNAi<br />

can easily be integrated into existing breeding lines.<br />

Health and welfare of transgenic farm animals<br />

Owing to integration and expression of transgenes, insertional<br />

mutagenesis and unwanted pathological side effects cannot be<br />

ruled out and concerns have been raised about the health of<br />

transgenic farm animals (Van Reenen et al. 2001). Transgenic<br />

farm animals have been produced since 1985 by pronuclear<br />

DNA injection, but only few systematic studies investigated<br />

the health status in cohorts of these animals. Overexpression<br />

of human growth hormone in pigs and sheep was associated<br />

with specific pathological phenotypes, which could be avoided<br />

by advanced constructs (Nottle et al. 1999). In pigs transgenic<br />

for human decay accelerating factor (hDAF), maintained<br />

at qualified pathogen free conditions, the haematology and<br />

blood chemistry were normal compared with transgenic animals<br />

(Tucker et al. 2002). With the exception of a slightly exceeded<br />

growth rate, no abnormalities were found. A detailed pathomorphological<br />

examination of nine lines of hemizygous pigs<br />

expressing human RCAs revealed no adverse effects correlating<br />

with transgene expression (Deppenmeier et al. 2006), providing<br />

clear evidence that transgenesis per se does not compromise<br />

animal health and welfare. Several of the investigated animals<br />

carried a bicistronic expression cassette, driving hCD59 and the<br />

tetracycline transactivator (Kues et al. 2006), suggesting that<br />

multi-transgenic animals are compatible with a normal health<br />

status. The hemizygous lines were fertile and produced normal<br />

litter sizes. However, attempts to breed these lines to homozygosity<br />

resulted in very small litter sizes in the first generation,<br />

likely reflecting inbreeding depression. In mice, several inbreeding<br />

rounds are necessary for extinction of a specific line (Taft<br />

et al. 2006).<br />

Cloning of mammals by somatic nuclear transfer is still<br />

tainted by low success rate of live offspring, typically with a<br />

range of 1–5% of transferred embryos surviving to term (Kues<br />

and Niemann 2004). Recently, we have obtained significant<br />

improvement of porcine cloning success by better selection and<br />

optimised treatment of the recipients, specifically by providing a<br />

24-h asynchrony between the pre-ovulatory oviducts of the recipients<br />

and the reconstructed embryos. Presumably, this gave the<br />

embryos additional time to achieve the necessary level of nuclear<br />

reprogramming. A second major improvement was a switch to<br />

peripubertal recipients synchronised with Altrenogest feeding<br />

over 15 days (Regumate, Cilag-Janssen, Neuss, Germany, 20 mg<br />

per animal per day) rather than using the first oestrus of prepubertal<br />

gilts (Walker et al. 2002). This resulted in pregnancy rates of<br />

∼80% and only slightly reduced litter size (Petersen et al. 2007).<br />

High embryonic and fetal attrition rates of cloned animals are<br />

specifically observed in ruminants that seem to be particularly<br />

susceptible to epigenetic aberrations. Cardiopulmonary and placental<br />

lesions and lesions in the skeletal and muscle system have<br />

been observed in cloned cattle (Hill et al. 1999). The observed<br />

adverse effects seem to be related to the cloning procedure


768 Reproduction, Fertility and Development H. Niemann and W. A. Kues<br />

per se and not primarily to transgenic modification, as similar<br />

adverse effects are also seen in cloned cattle without genetic<br />

modification (Panarace et al. 2006). In contrast, transgenic<br />

cloned pigs that survived the neonatal period are apparently normal<br />

in all haematologial and biochemical parameters (Carter<br />

et al. 2002).<br />

Studies with sufficient numbers of transgenic farm animals<br />

produced by emerging technologies, such as lentiviral and spermatogonial<br />

transgenesis, or expressing new constructs, like HAC<br />

and siRNA, will be necessary to assess the safety of these<br />

transgenic approaches.<br />

Outlook<br />

Recent advances in genomic technologies in farm animals<br />

present not only a major opportunity to address significant<br />

limitations in agricultural production, but also offer exciting<br />

prospects for medical research, significantly extending the<br />

important role of animals as models of human health and disease<br />

(Niemann et al. 2007). Developments in animal genomics<br />

have broadly followed the route that the human genome field<br />

has led, both in terms of completeness of data and in developing<br />

tools and applications as demonstrated by the recent availability<br />

of advanced drafts of the maps of the bovine, chicken, dog and<br />

bee genomes. The possibility of extending knowledge through<br />

deliberately engineering change in the genome is unique to animals<br />

and is currently <strong>bei</strong>ng developed through the integration<br />

of advanced molecular tools and breeding technologies. However,<br />

the full realisation of this exciting potential is handicapped<br />

by gaps in our understanding of embryo genomics and the epigenetic<br />

changes that are critical in the production of healthy<br />

offspring.<br />

The convergence of the recent advances in reproductive<br />

technologies with the tools of molecular biology opens a new<br />

dimension for animal breeding. Major prerequisites are the continuous<br />

refinement of reproductive biotechnologies and a rapid<br />

completion and refinement of livestock genomic maps. Embryonic<br />

stem cells (ES-cells) play a critical role in this context.<br />

Despite numerous efforts, no ES-cells with germ line contribution<br />

have been established from mammals other than the mouse.<br />

ES-like cells, i.e. cells that share several parameters with true EScells,<br />

have been reported in several species. The time period over<br />

which these cells could be maintained in culture varied from 13<br />

weeks to 3 years (Gjørret and Maddox-Hyttel 2005; Yadav et al.<br />

2005). Derivation of true ES cells in farm animals will permit<br />

exploitation of the full power of recombinant DNA technology<br />

in animal breeding and will thus be critical to develop sustainable<br />

and diversified animal production systems. We anticipate<br />

genetically modified animals will soon play a significant role in<br />

the biomedical field. Agricultural application might be further<br />

down the track given the complexity of some of the economically<br />

important traits.<br />

Acknowledgements<br />

The financial support by a grant from the DFG (FOR 535) is gratefully<br />

acknowledged. This review contains in parts data from Niemann, H., Kues,<br />

W. A., and Carnwath, J. W. (2005). Transgenic farm animals: present and<br />

future. Rev. Sci. Tech. OIE 24, 285–298.<br />

References<br />

Bach, F. H. (1998). Xenotransplantation: problems and prospects.Annu. Rev.<br />

Med. 49, 301–310. doi:10.1146/ANNUREV.MED.49.1.301<br />

Baguisi, A., Behboodi, E., Melican, D., Pollock, J. S., Destrempes, M. M.,<br />

et al. (1999). Production of goats by somatic cell nuclear transfer.<br />

Nat. Biotechnol. 17, 456–461. doi:10.1038/8632<br />

Brophy, B., Smolenski, G.,Wheeler,T.,Wells, D., L’Huillier, P., and Laible, G.<br />

(2003). Cloned transgenic cattle produce milk with higher levels of bcasein<br />

and k-casein. Nat. Biotechnol. 21, 157–162. doi:10.1038/<br />

NBT783<br />

Carter, B. D., Lai, L., Park, K.-W., Samuel, M., Lattimer, J. C., Jordan, K. R.,<br />

Estes, D. M., Besch-Williford, C., and Prather, R. (2002). Phenotyping<br />

of transgenic cloned pigs. Cloning Stem Cells 4, 131–145.<br />

doi:10.1089/153623002320253319<br />

Chan, A. W., Homan, E. J., Ballou, L. U., Burns, J. C., and Bremel, R. D.<br />

(1998). Transgenic cattle produced by reverse-transcribed gene transfer<br />

in oocytes. Proc. Natl. Acad. Sci. USA 95, 14028–14033.<br />

doi:10.1073/PNAS.95.24.14028<br />

Chang, K., Qian, J., Jiang, M., Liu, Y. H., Wu, M. C., et al.<br />

(2002). Effective generation of transgenic pigs and mice by<br />

linker based sperm-mediated gene transfer. BMC Biotechnol. 2,<br />

5. http://www.biomedcentral.com/1472-6750/2/5. doi:10.1186/1472-<br />

6750-2-5<br />

Cibelli, J. B., Stice, S. L., Golueke, P. L., Kane, J. J., Jerry, J., Blackwell, C.,<br />

Ponce de Leon, F. A., and Robl, J. M. (1998). Transgenic bovine<br />

chimeric offspring produced from somatic cell-derived stem like cells.<br />

Nat. Biotechnol. 16, 642–646. doi:10.1038/NBT0798-642<br />

Chou, C. Y., Horng, L. S., and Tsai, H. J. (2001). Uniform GFP-expression<br />

in transgenic medaka (Oryzias latipes) at the F0 generation. Transgenic<br />

Res. 10, 303–315. doi:10.1023/A:1016671513425<br />

Clark, J., and Whitelaw, B. (2003). A future for transgenic livestock. Nat.<br />

Rev. Genet. 4, 825–833. doi:10.1038/NRG1183<br />

Clements, J. E., Wall, R. J., Narayan, O., Hauer, D., Schoborg, R., Sheffer, D.,<br />

Powell, A., Carruth, L. M., Zink, M. C., and Rexroad, C. E. (1994).<br />

Development of transgenic sheep that express the visna virus envelope<br />

gene. Virology 200, 370–380. doi:10.1006/VIRO.1994.1201<br />

Cozzi, E., and White, D. J. G. (1995). The generation of transgenic<br />

pigs as potential organ donors for humans. Nat. Med. 1, 964–966.<br />

doi:10.1038/NM0995-964<br />

Dai, Y., Vaught, T. D., Boone, J., Chen, S. H., Phelps, C. J., et al. (2002).<br />

Targeted disruption of the alpha1,3-galactosyltransferase gene in cloned<br />

pigs. Nat. Biotechnol. 20, 251–255. doi:10.1038/NBT0302-251<br />

Dallas, A., and Vlassow, A. (2006). RNAi: A novel antisense technology and<br />

its therapeutic potential. Med. Sci. Monit. 12, RA67–RA74.<br />

Damak, S., Jay, N. P., Barrell, G. K., and Bullock, D. W. (1996a). Targeting<br />

gene expression to the wool follicle in transgenic sheep. Biotechnology<br />

(N. Y.) 14, 181–184. doi:10.1038/NBT0296-181<br />

Damak, S., Su, H., Jay, N. P., and Bullock, D. W. (1996b). Improved wool<br />

production in transgenic sheep expressing insulin-like growth factor 1.<br />

Biotechnology (N. Y.) 14, 185–188. doi:10.1038/NBT0296-185<br />

Denning, C., Burl, S., Ainslie, A., Bracken, J., Dinnyes, A., et al. (2001).<br />

Deletion of the alpha(1,3)galactosyl transferase (GGTA1) gene and<br />

the prion protein (PrP) gene in sheep. Nat. Biotechnol. 19, 559–562.<br />

doi:10.1038/89313<br />

Deppenmeier, S., Bock, O., Mengel, M., Niemann, H., Kues, W., Lemme, E.,<br />

Wirth, D.,Wonigeit, K., and Kreipe, H. (2006). Health status of transgenic<br />

pig lines expressing human complement regulator protein CD59. Xenotransplantation<br />

13, 345–356. doi:10.1111/J.1399-3089.2006.00317.X<br />

Dieckhoff, B., Karlas, A., Hofmann, A., Kues, W. A., Petersen, B.,<br />

Pfeifer, A., Niemann, H., Kurth, R., and Denner, J. (2007). Inhibition<br />

of porcine endogenous retroviruses (PERVs) in primary porcine cells<br />

by RNA interference using lentiviral vectors. Arch. Virol. 152, 629–634.<br />

doi:10.1007/S00705-006-0868-Y


Transgenic farm animals: an update Reproduction, Fertility and Development 769<br />

Echelard, Y., Ziomek, C., and Meade, H. (2006). Production of recombinant<br />

therapeutic proteins in the milk of transgenic animals. BioPharm Intern.<br />

2 (August 2006). http://www.biopharminternational.com/biopharm/<br />

content. pp. 1–7.<br />

Gandolfi, F. (1998). Spermatozoa, DNA binding and transgenic animals.<br />

Transgenic Res. 7, 147–155. doi:10.1023/A:1008828627735<br />

Gjørret, J. O., and Maddox-Hyttel, P. (2005). Attempts towards derivation<br />

and establishment of bovine embryonic stem cell-like cultures. Reprod.<br />

Fertil. Dev. 17, 113–124. doi:10.1071/RD04117<br />

Golding, M. C., Long, C. R., Carmell, M. A., Hannon, G. J., and<br />

Westhusin, M. E. (2006). Suppression of prion protein in livestock<br />

by RNA interference. Proc. Natl. Acad. Sci. USA 103, 5285–5290.<br />

doi:10.1073/PNAS.0600813103<br />

Golovan, S. P., Meidinger, R. G.,Ajakaiye,A., Cottrill, M.,Wiederkehr, M. Z.,<br />

et al. (2001). Pigs expressing salivary phytase produce low-phosphorus<br />

manure. Nat. Biotechnol. 19, 741–745. doi:10.1038/90788<br />

Gong, W., Wan, H., Tay, T. L., Wang, H., Chen, M., andYan, T. (2003). Development<br />

of transgenic fish for ornamental and bioreactor by strong expression<br />

of fluorescent proteins in the skeletal muscle. Biochem. Biophys.<br />

Res. Commun. 308, 58–63. doi:10.1016/S0006-291X(03)01282-8<br />

Grosse-Hovest, L., Muller, S., Minoia, R., Wolf, E., Zakhartchenko, V., et al.<br />

(2004). Cloned transgenic farm animals produce a bispecific antibody<br />

for T cell-mediated tumor cell killing. Proc. Natl. Acad. Sci. USA 101,<br />

6858–6863. doi:10.1073/PNAS.0308487101<br />

Hammer, R. E., Pursel, V. G., Rexroad, C. E., Jr, Wall, R. J., Bolt, D. J.,<br />

Ebert, K. M., Palmiter, R. D., and Brinster, R. L. (1985). Production<br />

of transgenic rabbits, sheep and pigs by microinjection. Nature 315,<br />

680–683. doi:10.1038/315680A0<br />

Haskell, R. E., and Bowen, R. A. (1995). Efficient production of transgenic<br />

cattle by retroviral infection of early embryos. Mol. Reprod. Dev. 40,<br />

386–390. doi:10.1002/MRD.1080400316<br />

Hill, J. R., Roussel, A. J., Cibelli, J. B., Edwards, J. F., Hooper, N. L.,<br />

et al. (1999). Clinical and pathologic features of cloned transgenic<br />

calves and fetuses (13 case study). Theriogenology 51, 1451–1465.<br />

doi:10.1016/S0093-691X(99)00089-8<br />

Hofmann, A., Kessler, B., Ewerling, S., Weppert, M., Vogg, B., et al. (2003).<br />

Efficient transgenesis in farm animals by lentiviral vectors. EMBO Rep.<br />

4, 1054–1060. doi:10.1038/SJ.EMBOR.7400007<br />

Hofmann, A., Zakhartchenko, V., Weppert, M., Sebald, H., Wenigerkind, H.,<br />

Brem, G., Wolf, E., and Pfeifer, A. (2004). Generation of transgenic<br />

cattle by lentiviral gene transfer into oocytes. Biol. Reprod. 71, 405–409.<br />

doi:10.1095/BIOLREPROD.104.028472<br />

Hofmann, A., Kessler, B., Ewerling, S., Kabermann, A., Brem, G.,<br />

Wolf, E., and Pfeifer, A. (2006). Epigenetic regulation of lentiviral<br />

transgene vectors in a large animal model. Mol. Ther. 13, 59–66.<br />

doi:10.1016/J.YMTHE.2005.07.685<br />

Honaramooz, A., Behboodi, E., Megee, S. O., Overton, S. A.,<br />

Galantino-Homer, H., Echelard, Y., and Dobrinski, I. (2003). Fertility<br />

and germline transmission of donor haplotype following germ cell<br />

transplantation in immunocompetent goats. Biol. Reprod. 69, 1260–<br />

1264. doi:10.1095/BIOLREPROD.103.018788<br />

Ibrahim, Z., Busch, J., Awwad, M., Wagner, R., Wells, K., and<br />

Cooper, D. K. C. (2006). Selected physiologic compatibilities and<br />

incompatibilities between human and porcine organ systems. Xenotransplantation<br />

13, 488–499. doi:10.1111/J.1399-3089.2006.00346.X<br />

Iqbal, K., Kues, W., and Niemann, H. (2007). Parent-of-origin dependent<br />

gene specific knockdown in mouse embryos. Bioph. Bioch. Res. Com.<br />

358, 727–732. doi:10.1016/JBBRC.2007.04.155<br />

Irgang, M., Sauer, I. M., Karlas, A., Zeilinger, K., Gerlach, J., Kurth, R.,<br />

Neuhaus, P., and Denner, J. (2003). Porcine endogenous retroviruses: no<br />

infection in patients treated with a bioreactor based on porcine liver cells.<br />

J. Clin. Virol. 28, 141–154. doi:10.1016/S1386-6532(02)00275-5<br />

Kues, W. A., and Niemann, H. (2004). The contribution of farm animals to<br />

human health.Trends Biotechnol. 22, 286–294. doi:10.1016/J.TIBTECH.<br />

2004.04.003<br />

Kues, W. A., Schwinzer, R., Wirth, D., Verhoeyen, E., Lemme, E.,<br />

Herrmann, D., Barg-Kues, B., Hauser, H., Wonigeit, H., and Niemann, H.<br />

(2006). Epigenetic silencing and tissue independent expression of a novel<br />

tetracycline inducible system in double-transgenic pigs. FASEB J. 20,<br />

E1–E10. doi:10.1096/FJ.05-5415FJE<br />

Kuroiwa, Y., Kasinathan, P., Choi, Y. J., Naeem, R., Tomizuka, K.,<br />

et al. (2002). Cloned transchromosomic calves producing human<br />

immunoglobulin. Nat. Biotechnol. 20, 889–894. doi:10.1038/<br />

NBT727<br />

Kuwaki, K., Tseng, Y. L., Dor, F. J., Shimizu, A., House, S. L., et al.<br />

(2005). Heart transplantation in baboons usion α1,3-glactosyltransferase<br />

knock out pigs as donors: initial experiments. Nat. Med. 11, 29–31.<br />

doi:10.1038/NM1171<br />

Lai, L., Kolber-Simonds, D., Park, K. W., Cheong, H. T., Greenstein, J. L.,<br />

et al. (2002). Production of alpha-1,3-galactosyltransferase knockout<br />

pigs by nuclear transfer cloning. Science 295, 1089–1092.<br />

doi:10.1126/SCIENCE.1068228<br />

Lai, L., Kang, J. X., Li, R., Wang, J., Witt, W. T., et al. (2006). Generation of<br />

cloned transgenic pigs rich in omega-3 fatty acids. Nat. Biotechnol. 24,<br />

435–436. doi:10.1038/NBT1198<br />

Lavitrano, M., Bacci, M. L., Forni, M., Lazzereschi, D., Di Stefano, C.,<br />

et al. (2002). Efficient production by sperm-mediated gene transfer<br />

of human decay accelerating factor (hDAF) transgenic pigs for<br />

xenotransplantation. Proc. Natl. Acad. Sci. USA 99, 14230–14235.<br />

doi:10.1073/PNAS.222550299<br />

Lo, D., Pursel, V., Linton, P. J., Sandgren, E., Behringer, R., Rexroad, C.,<br />

Palmiter, R. D., and Brinster, R. L. (1991). Expression of mouse IgA<br />

by transgenic mice, pigs and sheep. Eur. J. Immunol. 21, 1001–1006.<br />

doi:10.1002/EJI.1830210421<br />

Manzini, S., Vargiolu, A., Stehle, I. M., Bacci, M. L., Cerrito, M. G.,<br />

et al. (2006). Genetically modified pigs produced with a nonviral<br />

episomal vector. Proc. Natl. Acad. Sci. USA 103, 17672–17677.<br />

doi:10.1073/PNAS.0604938103<br />

McCreath, K. J., Howcroft, J., Campbell, K. H., Colman, A., Schnieke, A. E.,<br />

and Kind, A. J. (2000). Production of gene-targeted sheep by<br />

nuclear transfer from cultured somatic cells. Nature 405, 1066–1069.<br />

doi:10.1038/35016604<br />

Meade, H. M., Echelard, Y., Ziomek, C. A., Young, M. W., Harvey, M.,<br />

Cloe, E. S., Groet, S., Smith, T. E., and Curling, J. M. (1999). Expression<br />

of recombinant proteins in the milk of transgenic animals. In ‘Gene<br />

Expression Systems. Using Nature for the Art of Expression’. (Eds<br />

J. M. Fernandez and J. P. Hoeffler.) pp. 399–427. (Academic Press: San<br />

Diego, CA.)<br />

Müller, M., Brenig, B., Winnacker, E. L., and Brem, G. (1992). Transgenic<br />

pigs carrying cDNA copies encoding the murine Mx1 protein which<br />

confers resistance to influenza virus infection. Gene 121, 263–270.<br />

doi:10.1016/0378-1119(92)90130-H<br />

Niemann, H. (2004). Transgenic pigs expressing plant genes. Proc. Natl.<br />

Acad. Sci. USA 101, 7211–7212. doi:10.1073/PNAS.0402011101<br />

Niemann, H., and Kues, W. A. (2003). Application of transgenesis on livestock<br />

for agriculture and biomedicine. Anim. Reprod. Sci. 79, 291–317.<br />

doi:10.1016/S0378-4320(03)00169-6<br />

Niemann, H., Kues, W. A., and Carnwath, J. W. (2005). Transgenic farm<br />

animals: present and future. OIE-Report: Biotechnology applications<br />

in animal health and production, Revue scientifique et technique 24,<br />

285–291.<br />

Niemann, H., Carnwath, J. W., and Kues, W. (2007). Application of<br />

array technology to mammalian embryos. Theriogenology, in press.<br />

doi:10.1016/J.THERIOGENOLOGY.2007.05.041


770 Reproduction, Fertility and Development H. Niemann and W. A. Kues<br />

Nottle, M. B., Nagashima, H., Verma, P. J., Du, Z. T., Grupen, C. G.,<br />

et al. (1999). Production and analysis of transgenic pigs containing<br />

a metallothionein porcine growth hormon gene construct. In ‘Transgenic<br />

Animals in Agriculture’. (Eds J. D. Murray, G. B. Anderson,<br />

A. M. Oberbauer and M. M. McGloughlin.) pp. 145–156. (CABI<br />

Publishing: New York.)<br />

Panarace, M., Agüero, J. I., Garrote, M., Jauregui, G., Segovia, A.,<br />

et al. (2006). How healthy are clones and their progeny: 5<br />

years of field experience. Theriogenology 67, 142–151. doi:10.1016/<br />

J.THERIOGENOLOGY.2006.09.036<br />

Perry, A. C. F., Wakayama, T., Kishikawa, H., Kasai, T., Okabe, M.,<br />

Toyoda, Y., and Yanagimachi, R. (1999). Mammalian transgenesis<br />

by intracytoplasmic sperm injection. Science 284, 1180–1183.<br />

doi:10.1126/SCIENCE.284.5417.1180<br />

Petersen, B., Kues, W. A., Lucas-Hahn, A., Queisser, A. L., Lemme, E.,<br />

Hoelker, M., Carnwath, J. W., and Niemann, H. (2006). Generation<br />

of pigs transgenic for hCD59/DAF and human thrombomodulin by<br />

somatic nuclear transfer. Reprod. Fertil. Dev. 18, 142. doi:10.1071/<br />

RDV18N2AB67<br />

Petersen, B., Lucas-Hahn, A., Lemme, E., Hornen, N., Hassel, P., and<br />

Niemann, H. (2007). Preovulatory embryo transfer increases cloning<br />

efficiencies in pigs. Reprod. Fertil. Dev. 19, 155–156. [Abstract]<br />

Phelps, C. J., Koike, C., Vaught, T. D., Boone, J., Wells, K. D., et al. (2003).<br />

Production of alpha 1,3-galactosyltransferase deficient pigs. Science<br />

299, 411–414. doi:10.1126/SCIENCE.1078942<br />

Plasterk, R. H. (2002). RNA silencing: the genomes immune system. Science<br />

296, 1263–1265. doi:10.1126/SCIENCE.1072148<br />

Platenburg, G. J., Kootwijk, E. P., Kooiman, P. M., Woloshuk, S. L.,<br />

Nuijens, J. H., Krimpenfort, P. J., Pieper, F. R., de Boer, H. A., and Strijker,<br />

R. (1994). Expression of human lactoferrin in milk of transgenic<br />

mice. Transgenic Res. 3, 99–108. doi:10.1007/BF01974087<br />

Platt, J. L., and Lin, S. S. (1998). The future promises of xenotransplantation.<br />

Ann. N. Y. Acad. Sci. 862, 5–18. doi:10.1111/J.1749-<br />

6632.1998.TB09112.X<br />

Pursel, V. G., Wall, R. J., Mitchell, A. D., Elsasser, T. H., Solomon, M. B.,<br />

Coleman, M. E., Mayo, F., and Schwartz, R. J. (1999). Expression of<br />

insulin-like growth factor-1 in skeletal muscle of transgenic pigs. In<br />

‘Transgenic Animals in Agriculture’. (Eds J. D. Murray, G. B. Anderson,<br />

A. M. Oberbauer and M. M. McGloughlin.) pp. 131–144. (CABI<br />

Publishing: New York.)<br />

Reh, W. A., Maga, E. A., Collette, N. M., Moyer, A., Conrad-Brink, J. S.,<br />

Taylor, S. J., DePeters, E. J., Oppenheim, S., Rowe, J. D.,<br />

BonDurant, R. H., Anderson, G. B., and Murray, J. D. (2004). Hot<br />

topic: using a stearoyl-CoA desaturase transgene to alter milk fatty acid<br />

composition. J. Dairy Sci. 87, 3510–3514.<br />

Robl, J. M., Wang, Z., Kasinathan, P., and Kuroiwa, Y. (2007). Transgenic<br />

animal production and animal biotechnology. Theriogenology 67, 127–<br />

133. doi:10.1016/J.THERIOGENOLOGY.2006.09.034<br />

Rudolph, N. S. (1999). Biopharmaceutical production in transgenic<br />

livestock. Trends Biotechnol. 17, 367–374. doi:10.1016/S0167-<br />

7799(99)01341-4<br />

Saeki, K., Matsumoto, K., Kinoshita, M., Suzuki, I.,Tasaka,Y., et al., (2004).<br />

Functional expression of a Delta12 fatty acid desaturase gene from<br />

spinach in transgenic pigs. Proc. Natl. Acad. Sci. USA 101, 6361–6366.<br />

doi:10.1073/PNAS.0308111101<br />

Schnieke, A. E., Kind, A. J., Ritchie, W. A., Mycock, K., Scott, A. R.,<br />

Ritchie, M., Wilmut, I., Colman, A., and Campbell, K. H. (1997).<br />

Human factor IX transgenic sheep produced by transfer of nuclei<br />

from transfected fetal fibroblasts. Science 278, 2130–2133. doi:<br />

10.1126/SCIENCE.278.5346.2130<br />

Squires, E. J. (1999). Status of sperm-mediated delivery methods for gene<br />

transfer. In ‘Transgenic Animals in Agriculture’. (Eds J. D. Murray,<br />

http://www.publish.csiro.au/journals/rfd<br />

G. B. Anderson, A. M. Oberbauer and M. M. McGloughlin.) pp. 87–96.<br />

(CABI Publishing: New York.)<br />

Switzer, W. M., Michler, R. E., Shangmugam, V., Matthews, A.,<br />

Hussain, A. I., et al. (2001). Lack of cross-species transmission of<br />

porcine endogenous retrovirus infection to nonhuman primate recipients<br />

of porcine cells, tissues and organs. Transplantation 71, 959–965.<br />

doi:10.1097/00007890-200104150-00022<br />

Taft, R. A., Davisson, M., and Wiles, M. V. (2006). Know thy mouse. Trends<br />

Genet. 22, 649–653. doi:10.1016/J.TIG.2006.09.010<br />

Tenenhaus Dann, C., Alvarado, A. L., Hammer, R. E., and Garbers, D. L.<br />

(2006). Heritable and stable gene knockdown in rats. Proc. Natl. Acad.<br />

Sci. USA 103, 11246–11251. doi:10.1073/PNAS.0604657103<br />

Tucker, A., Belcher, C., Moloo, B., Bell, J., Mazzulli, T., Humar, Y.,<br />

Hughes, A., McArdle, P., and Talbot, A. (2002). The production of<br />

transgenic pigs for potential use in clinical xenotransplantation: baseline<br />

clinical pathology and organ size studies. Xenotransplantation 9,<br />

203–208. doi:10.1034/J.1399-3089.2002.01052.X<br />

USA Food and Drug Administration (2003). FDA statement regarding<br />

glofish. http://www.fda.gov/bbs/topics/NEWS/2003/NEW00994.html<br />

[verified May 2007].<br />

Van den Hout, J. M., Reuser, A. J., de Klerk, J. B., Arts, W. F., Smeitink, J. A.,<br />

and Van der Ploeg, A. T. (2001). Enzyme therapy for Pompe disease with<br />

recombinant human α-glucosidase from rabbit milk. J. Inherit. Metab.<br />

Dis. 24, 266–274. doi:10.1023/A:1010383421286<br />

Van Reenen, C. G., Meuwissen, T. H. E., Hopster, H., Oldenbroek, K.,<br />

Kruip, T. H., and Blokhuis, H. J. (2001). Transgenesis may affect farm<br />

animal welfare: a case for systemic risk assessment. J. Anim. Sci. 79,<br />

1763–1769.<br />

Walker, S. C., Shin, T., Zaunbrecher, G. M., Romario, J. E., Johnson, G. A.,<br />

Bazer, F. W., and Piedrahita, J. A. (2002). A highly efficient method<br />

for porcine cloning by nuclear transfer using in vitro matured oocytes.<br />

Cloning Stem Cells 4, 105–112.<br />

Wall, R. J., Powell, A., Paape, M. J., Kerr, D. E., Bannermann,<br />

D. D., Pursel, V. G., Wells, K. D., Talbot, N., and Hawk,<br />

H. (2005). Genetically enhanced cows resist intramammary Staphylococcus<br />

aureus infection. Nat. Biotechnol. 23, 445–451. doi:10.1038/<br />

NBT1078<br />

Willadsen, S. M. (1986). Nuclear transplantation in sheep embryos. Nature<br />

320, 63–65. doi:10.1038/320063A0<br />

Wilmut, I., Schnieke, A. E., McWhir, J., Kind, A. J., and Campbell, K. H.<br />

(1997). Viable offspring derived from fetal and adult mammalian cells.<br />

Nature 385, 810–813. doi:10.1038/385810A0<br />

Whitelaw, C. B., Radcliffe, P. A., Ritchie, W. A., Carlisle, A., Ellard, F. M.,<br />

Pena, R. N., Rowe, J., Clark, A. J., King, T. J., and Mitrophanous, K. A.<br />

(2004). Efficient generation of transgenic pigs using equine infectious<br />

anaemia virus (EIAV) derived vector. FEBS Lett. 571, 233–236.<br />

doi:10.1016/J.FEBSLET.2004.06.076<br />

Wheeler, M. B., Bleck, G. T., and Donovan, S. M. (2001). Transgenic alteration<br />

of sow milk to improve piglet growth and health. Reproduction<br />

58(Suppl.), 313–324.<br />

Yadav, P. S., Kues, W. A., Hermann, D., Carnwath, J. W., and Niemann, H.<br />

(2005). Bovine ICM derived cells expressed the Oct4 ortholog. Mol.<br />

Reprod. Dev. 72, 182–190.<br />

Yamada, K., Yazawa, K., Shimizu, A., Iwanaga, T., Hisashi, Y., et al. (2005).<br />

Marked prolongation of porcine renal xenograft survival in baboons<br />

through the use of α1,3-galactosyltransferase gene-knockout donors and<br />

the cotransplantation of vascularized thymic tissue. Nat. Med. 11, 32–34.<br />

doi:10.1038/NM1172<br />

Manuscript received 26 February 2007, accepted 16 April 2007


This article was published in an Elsevier journal. The attached copy<br />

is furnished to the author for non-commercial research and<br />

education use, including for instruction at the author’s institution,<br />

sharing with colleagues and providing to institution administration.<br />

Other uses, including reproduction and distribution, or selling or<br />

licensing copies, or posting to personal, institutional or third party<br />

websites are prohibited.<br />

In most cases authors are permitted to post their version of the<br />

article (e.g. in Word or Tex form) to their personal website or<br />

institutional repository. Authors requiring further information<br />

regarding Elsevier’s archiving and manuscript policies are<br />

encouraged to visit:<br />

http://www.elsevier.com/copyright


Abstract<br />

Author's personal copy<br />

Application of DNA array technology to mammalian embryos<br />

H. Niemann *, J.W. Carnwath, W. Kues<br />

Department of Biotechnology, Institute for Animal Breeding, Mariensee, D-31535 Neustadt, Germany<br />

Early embryogenesis depends on a tightly choreographed succession of gene expression patterns which define normal<br />

development. Fertilization and the first zygotic cleavage involve major changes to paternal and maternal chromatin and translation<br />

of maternal RNAs which have been sequestered in the oocyte during oogenesis. At a critical species-specific point known as the<br />

major onset of embryonic expression, there is a dramatic increase in expression from the new diploid genome. The advent of array<br />

technology has, for the first time, made possible to determine the transcriptional profile of all 20,000 mammalian genes during<br />

embryogenesis, although the small amount of mRNA in a single embryo necessitates either pooling large numbers of embryos or a<br />

global amplification procedure to give sufficient labeled RNA for analysis. Following array hybridization, various bioinformatic<br />

tools must be employed to determine the expression level for each gene, often based on multiple oligonucleotide probes and<br />

complex background estimation protocols. The grouped analysis of clusters of genes which represent specific biological pathways<br />

provides the key to understanding embryonic development, embryonic stem cell proliferation and the reprogramming of gene<br />

expression after somatic cloning. Arrays are <strong>bei</strong>ng developed to address specific biological questions related to embryonic<br />

development including DNA methylation and microRNA expression. Array technology in its various facets is an important<br />

diagnostic tool for the early detection of developmental aberrations; for improving the safety of assisted reproduction technologies<br />

for man; and for improving the efficiency of producing cloned and/or transgenic farm animals. This review discusses current<br />

approaches and limitations of DNA microarray technology with emphasis on bovine embryos.<br />

# 2007 Elsevier Inc. All rights reserved.<br />

Keywords: Preimplantation embryo; Transcriptome; RNA amplification; DNA microarray; Bioinformatics; Plasticity of gene expression;<br />

Maternal and embryonic gene expression<br />

1. Introduction<br />

Sequencing and annotation of the human genome<br />

resulted in a downward estimation of the total number<br />

of protein-coding genes to approximately 20,000 [1].<br />

This is much lower than previous estimates which<br />

ranged to as many as 150,000 genes and is similar<br />

to estimates for other vertebrates’ species including<br />

mouse, chicken and pufferfish [2–6], or bovine and dog<br />

* Corresponding author. Tel.: +49 5034 871 148;<br />

fax: +49 5034 871 101.<br />

E-mail address: niemann@tzv.fal.de (H. Niemann).<br />

Theriogenology 68S (2007) S165–S177<br />

0093-691X/$ – see front matter # 2007 Elsevier Inc. All rights reserved.<br />

doi:10.1016/j.theriogenology.2007.05.041<br />

www.theriojournal.com<br />

(http://www.hgse.bcm.tmc.edu/projects/bovine [7,8]).<br />

It is clear that complex organisms exploit a variety of<br />

regulatory mechanisms to extend the functionality of<br />

their genomes [9], including differential promoter<br />

activation, alternative RNA splicing, RNA modification,<br />

RNA editing, localization, translation and stability of<br />

RNA, expression of non-coding RNA, antisense RNA<br />

and microRNAs [10–16]. It is important to note that most<br />

of these mechanisms work together at the RNA level to<br />

produce the functional transcriptome of an organism.<br />

Tissue-specific transcriptomic profiles indicate important<br />

features of the regulatory process. During the past<br />

decade, efficient high-throughput methods have been<br />

developed for whole genome sequencing, transcriptome


S166<br />

and proteome analysis [17–19]. These technologies are<br />

crucial to improve the efficiency and safety of assisted<br />

reproduction techniques, to facilitate the development of<br />

novel therapies based on regenerative medicine and to<br />

insure the quality of biotechnological products at the<br />

molecular level.<br />

The application of array technology to the analysis of<br />

preimplantation mammalian embryos poses-specific<br />

challenges associated with the picogram levels of<br />

mRNA in a single embryo, the plasticity of the<br />

embryonic transcriptome and the difficulties in obtaining<br />

in vivo developing embryos in some species. Gene<br />

expression patterns in mammalian embryos are more<br />

complex than those in most somatic cells due to the<br />

dramatic restructuring of the paternal chromatin, the<br />

major onset of embryonic transcription, events related<br />

to the cellular differentiation at the early morula stage,<br />

blastocyst formation, expansion and hatching and<br />

finally implantation. Embryo development is first<br />

dependent on maternally stored transcripts which are<br />

gradually depleted until the embryo produces its own<br />

transcripts after a switch to the embryonic expression<br />

program. The gene expression patterns and RNA<br />

stability in oocytes and early embryos prior to the<br />

Author's personal copy<br />

H. Niemann et al. / Theriogenology 68S (2007) S165–S177<br />

activation of embryonic expression are dramatically<br />

different from what is seen after major onset of<br />

embryonic transcription. The onset of embryonic gene<br />

transcription occurs at a species-specific time point; in<br />

mice, the major activation starts in late 1-cell embryos,<br />

in pigs it occurs at 4-cell stage, in men at the 4- to 8-cell<br />

stage and in bovine embryos it is delayed until the 8- to<br />

16-cell stage [20] (Fig. 1). Due to the late onset of this<br />

maternal-to-embryonic transition in bovine embryos,<br />

they provide an excellent model in which to study early<br />

reprogramming events in detail. Major epigenetic<br />

reorganization also occurs during preimplantation<br />

development. This includes DNA demethylation and<br />

methylation as well as targeted modifications of histone<br />

methylation and acetylation all of which play a role in<br />

controlling chromatin remodeling and X chromosome<br />

inactivation [21].<br />

Another challenge for applying DNA arrays to<br />

preimplantation embryos is the high degree of expression<br />

plasticity seen in early stage embryos. Preimplantation<br />

stages of several species can be cultured in vitro.<br />

However, this is frequently associated with changes in<br />

chromatin configuration and gene expression [22–24].<br />

In vitro production of bovine embryos is usually<br />

Fig. 1. Early reprogramming steps of maternal and embryonic transcriptomes are extended in bovine embryos (bottom) compared to mouse (upper)<br />

and human embryos (middle panel).


necessary to provide the large numbers of embryos<br />

needed for an array analysis in this species but it is clear<br />

that the changes in embryonic transcription induced by<br />

the culture method itself can affect the transcriptional<br />

profile. With the advent of rather complete and<br />

validated genomic maps for farm animals, DNA arrays<br />

are increasingly <strong>bei</strong>ng used to unravel the transcriptional<br />

profile during preimplantation development.<br />

Here, we discuss the current status and limitations of<br />

DNA microarray analysis as it is applied to preimplantation<br />

embryos with an emphasis of the bovine<br />

embryonic transcriptome.<br />

2. Array analysis of preimplantation embryos<br />

2.1. Principles of array technology<br />

The most fundamental requisite for detailed expression<br />

array analysis is availability of sufficient genomic<br />

sequences to produce an array. Because they require less<br />

hybridization solution than larger array formats, microarrays<br />

are routinely used for large scale transcriptome<br />

analyses. They have been widely and successfully<br />

employed for simultaneously monitoring the expression<br />

of complete genomes, thus identifying genes that are<br />

differentially expressed in distinct cell types, developmental<br />

stages, disease states or in cells exposed to various<br />

reagents [17]. High-density oligonucleotide arrays can be<br />

used to monitor the activity of an entire genome of a<br />

specific organism, thus defining the cellular transcriptome<br />

in the form of a snapshot of the transcription profile<br />

and revealing details of the dynamic changes.<br />

Microarrays consist of hundreds to millions of<br />

microscopically small spots attached to a matrix of<br />

glass or plastic, each containing a short but gene specific<br />

nucleic acid sequence, typically a synthetic oligonucleotide<br />

or cDNA. The microarray design can be miniaturized<br />

to few millimeters, allowing sample volumes to be<br />

reduced to a few microliters. The biological principle of<br />

the DNA array method is hybridization based on<br />

complementary base pair binding of nucleic acid<br />

molecules to their sequence matched spotted probes. A<br />

sequence of 20 bases is sufficient to identify a gene with<br />

certainty. To visualize differences in transcriptional<br />

activity, a test sample of the complex nucleic acid mixture<br />

extracted from cells or tissues is chemically tagged with a<br />

fluorochrome or chemical conjugate and then hybridized<br />

to the array. This results in a signal at each gene specific<br />

spot whose intensity is due to retention of some amount of<br />

its complementary labeled nucleic acid. The hybridization<br />

signals are quantitatively recorded and the amount of<br />

complementary nucleic acid at each point is calculated<br />

Author's personal copy<br />

H. Niemann et al. / Theriogenology 68S (2007) S165–S177 S167<br />

with respect to local background. In some cases, this<br />

involves averaging the results from several independent<br />

spots all representing different locations within the target<br />

gene. The number of genes from which expression can be<br />

simultaneously detected, is potentially unlimited and<br />

whole genome arrays are commercially available. In<br />

principle, genome coverage on an array is only limited by<br />

the density and number of spots on the array.<br />

Array analysis depends first and foremost on the<br />

determination of probability that a transcript is present<br />

or absent. This is based on statistical analysis of the<br />

strength of the signal above background often using<br />

several probes for the gene and often using several<br />

repeated array hybridizations. There are a variety of<br />

different statistical methods which can be used for this<br />

step and the choice of which to use depends on the goals<br />

of the experiment. One must decide between having<br />

more false positives or false negatives. The strength of<br />

the statistical analysis is increased by the number of<br />

repeated hybridizations but can also be increased by<br />

grouping genes into clusters or pathways which can be<br />

used to smooth out individual variation.<br />

Analysis of the vast amount of data requires<br />

sophisticated bioinformatic tools [25]. The objective<br />

of a microarray experiment is usually to improve the<br />

understanding of biological mechanisms. This requires<br />

bioinformatic methods which group the detection<br />

profiles of individual genes based on current knowledge<br />

of signaling pathways or gene function. Several<br />

databases, such as Gene Ontology (GO; http://www.geneontology.org)<br />

and KEGG, the Kyoto Encyclopaedia<br />

of Genes and Genomes, (http://www.genome.jp/kegg)<br />

are available for this and can readily be accessed over<br />

the internet. Further bioinformatic tools and a variety of<br />

statistical methods have been developed to refine<br />

reliability. It is usually necessary to repeat each array<br />

analysis 2–3 times to achieve satisfactorily accurate<br />

measurements. The most important criteria for evaluating<br />

any microarray analysis are reproducibility of the<br />

data, specificity of target gene detection, local background<br />

and gene specific background. Certain standards<br />

have been agreed upon and strict adherence to criteria<br />

described as Minimal Information about a Microarray<br />

Experiment (MIAME) is absolutely essential to<br />

guarantee consistent and reproducible datasets for the<br />

scientific community [26].<br />

2.2. Technical considerations<br />

A specific problem encountered in the array analysis<br />

of embryonic gene expression is known as the loading<br />

artifact. The necessity of loading a standardized amount


S168<br />

of labeled RNA onto each array for hybridization causes<br />

a distortion of the quantitative results because of the<br />

large change in the amount and composition of mRNA<br />

which occurs throughout preimplantation development<br />

but particularly at the major onset of embryonic gene<br />

expression. The blastocyst contains approximately<br />

twice as much mRNA as the oocyte mainly due to<br />

the greater number of individual mRNAs expressed in<br />

the blastocyst. When the total amount of labeled RNA<br />

applied to the oocyte array is the same as that applied to<br />

the blastocyst array, the amount of probe for each<br />

individual transcript in the blastocyst array is halved.<br />

Blastocyst transcripts which are close to the detection<br />

limit are lost and what appears to be a 50% reduction in<br />

transcription between the oocyte and the blastocyst for<br />

an individual gene may be only due to the loading<br />

artifact. If it would be possible to establish a group of<br />

genes whose expression levels were absolutely constant<br />

throughout preimplantation development, one could<br />

correct for the loading artifact. Similarly, if it would be<br />

possible to accurately measure the total amount of<br />

mRNA at each developmental stage, this information<br />

could be used to correct for the loading artifact.<br />

However, RT-PCR analysis has not yet revealed any<br />

gene whose expression is held at a constant level over<br />

preimplantation development, making accurate measurements<br />

of the mRNA in embryos under various<br />

culture conditions problematic. In interpretation of<br />

array results, one must realize that a number of genes,<br />

transcribed at low level in the blastocyst, are not<br />

detected because of the loading artifact and there is a<br />

systematic underestimation of the number of new genes<br />

expressed during development.<br />

One of the most commonly used methods for mRNA<br />

purification from embryos is based on poly-dT coated<br />

magnetic beads. These hybridize with the poly-A tail of<br />

mRNA and pull the mRNA out of the solution. It is<br />

important to realize that the length of the poly-A tail has<br />

an influence on the extraction efficiency. Poly-A tail<br />

length changes during embryonic development, especially<br />

between the oocyte stage where mRNA is<br />

protected and stored for later use and subsequent stages<br />

where mRNAs are transcribed for immediate use. The<br />

XIST gene provides an interesting biological example.<br />

This gives transcripts of several different lengths, which<br />

are distinguished by variations at the 3 0 end. Some of<br />

these have a signal for polyadenylation and some do not<br />

[27]. Thus, extraction with poly-dT coated magnetic<br />

beads will distortion the picture of transcription from<br />

this gene. Poly-dT extraction results in a second artifact<br />

due to the fact that 5 0 fragments of mRNA molecules<br />

which have no poly-A sequence will be lost while 3 0<br />

Author's personal copy<br />

H. Niemann et al. / Theriogenology 68S (2007) S165–S177<br />

fragments will be retained. The probability of fragmentation<br />

is a function of the length of the transcript.<br />

One routine way to detect this artifact, which is a<br />

reflection of RNA quality, is to compare spots matching<br />

3 0 sequences of selected genes with the signals from 5 0<br />

spots for the same genes. An array where this distortion<br />

is noted should not be included in the analysis.<br />

A commonly used RNA amplification method for<br />

Affymetrix array hybridization is the ‘‘Whole Transcript<br />

(WT) Sense Target Labeling Assay’’ (www.affymetrix.com).<br />

The system most suitable for<br />

preimplantation embryo analysis is the ‘‘Two-Cycle<br />

Target Labeling and Control Reagent package’’ which<br />

includes a protocol for target preparation from 10 to<br />

100 ng of total RNA (see www.affymetrix.com).<br />

One of the most difficult specimens to obtain for<br />

embryonic expression analysis is a naturally developing<br />

embryo. Even embryos flushed directly from the<br />

oviduct may have altered gene expression patterns<br />

due to the nutritional status of the donor, the use of<br />

superovulation, exposure to flushing medium such as<br />

PBS, exposure to non-physiological levels of oxygen<br />

and temperature variations. There is no question that in<br />

vitro culture affects gene expression [22–24]. For<br />

example, RT-PCR analysis shows that the expression of<br />

heat shock protein increases continuously during<br />

embryonic development in vitro. The time of exposure<br />

to non-physiological conditions between flushing and<br />

freezing will influence relative expression data as it is<br />

known that different transcripts have different halflives.<br />

As a general rule, long transcripts are rare and<br />

have shorter half-lives. Experiments with embryos must<br />

be designed to control for such artifacts by minimizing<br />

the time between isolation and cryopreservation and<br />

must be interpreted with full knowledge of their<br />

limitations.<br />

The information which can be obtained by array<br />

analysis can be either the timing of (detectable) gene<br />

expression, the relative amount of individual gene<br />

expression between matched samples, the number of<br />

different genes detected in different samples using the<br />

same array and complex generalizations with respect to<br />

clustered genes and pathways. When artifacts have been<br />

minimized, the paired sample approach is a powerful<br />

method for analyzing expression. The number of arrays<br />

used in a specific analysis may be limited by practical or<br />

financial constraints but current analytical programs are<br />

quite capable of integrating the results from hundreds of<br />

arrays. The sensitivity of analysis can generally be<br />

enhanced by increasing the number of arrays used in an<br />

experiment but even when the number of arrays is<br />

restricted, it is possible to increase the sensitivity of


analysis and the information content of the study by<br />

performing multivariate statistical analysis of, for<br />

example, clusters of genes related to the same<br />

pathway. This has the advantage of averaging out<br />

the fluctuations of weak signals. Array production is<br />

consistent, so that the comparison of results from two<br />

or more individual arrays is both possible and<br />

recommended. In some systems, additional sensitivity<br />

can be obtained by applying two differentially labeled<br />

samples to a single array. One of the most robust<br />

measurements is determination of the subset of<br />

strongest expressed genes at any stage since this<br />

avoids problems of sensitivity, RNA stability and<br />

statistical fluctuation. The main limitation for this<br />

form of analysis is saturation of individual spots on the<br />

array which will affect the relative rankings of highly<br />

expressed genes. Interpretations of biological effects<br />

seeninarrayanalysismustbeconfirmedbyRT-PCR<br />

verification.<br />

2.3. Global amplification strategies<br />

Early embryogenesis requires the well orchestrated<br />

expression of genes to ensure normal development, and<br />

is associated with significant physical reorganization of<br />

the epi-genome and of the transcriptome. It is estimated<br />

that 5000–10,000 genes are simultaneously expressed<br />

at carefully controlled levels. Transcripts for key<br />

transcription factors may be present in only a small<br />

number of copies, while transcripts encoding major<br />

structural proteins, such as MSY2/fbx2 can represent<br />

2% of the mRNA pool [28].<br />

Estimations of the RNA content of mammalian<br />

oocytes range from 0.2 to 2.4 ng and blastocysts are<br />

estimated to have roughly twice as much, 0.4–5 ng<br />

[29,30]; approximately 10–20% of the total RNA<br />

consists of polyadenylated mRNA. Protocols for array<br />

hybridization typically require as much as 5–10 mg of<br />

purified nucleic acids (www.affymetrix.com). Thus<br />

samples with tiny amounts of RNA, such as laser<br />

dissected biopsies or early embryos, require unbiased<br />

global amplification prior to array analysis. The amount<br />

of mRNA, which can be isolated from a single embryo,<br />

needs to be amplified for 10,000 times to obtain<br />

sufficient material for one hybridization reaction. It is<br />

critical that this amplification does not distortion the<br />

relative proportions of mRNA found in the original<br />

mRNA pool.<br />

During the past few years several methods have been<br />

published that are compatible with efficient amplification<br />

of RNA isolated from single cells, embryos or small pools<br />

of embryos without distortion of the RNA pool ([30–33],<br />

Author's personal copy<br />

H. Niemann et al. / Theriogenology 68S (2007) S165–S177 S169<br />

http://www.affymetrix.com/, http://www.epicentre.com,<br />

http://www.genisphere.com). Crucial steps in these<br />

global amplification protocols are the conversion of<br />

mRNA into cDNA by reverse transcriptase and second<br />

strand cDNA synthesis, followed by amplification and<br />

labeling by in vitro transcription (IVT) using T7 RNA<br />

polymerase. To introduce the recognition sequence for T7<br />

RNA polymerase, synthetic oligo(dT) primers which<br />

include a T7 recognition sequence are used for the reverse<br />

transcription reaction. Some protocols combine the T7<br />

RNA polymerase reaction with the polymerase chain<br />

reaction (PCR) to achieve the necessary amplification<br />

[30,31]. Affymetrix, one of the prominent companies in<br />

array analysis, recommends two cycles of IVT by T7<br />

RNA polymerase. For any protocol, a rigorous validation<br />

of the amplification is essential to obtain meaningful<br />

hybridization data (Fig. 2). Less amplification is<br />

generally better as there will always be differences in<br />

efficiency due to variations in the length and nucleotide<br />

composition of the mRNA species. Semi-quantitative<br />

RT-PCR or Real Time PCR assays comparing amplified<br />

and non-amplified material with a broad panel of<br />

transcripts are important for validation of the amplification<br />

protocol.<br />

2.4. The value of cross-species hybridization<br />

The rapid improvements of cDNA-chip technologies<br />

and the availability of multi-species sequence databases,<br />

e.g. GO and KEGG, have made it possible to<br />

compare gene expression profiles both within a single<br />

organism and across a wide variety of organisms on a<br />

large-scale. Cross-species gene-expression comparison<br />

is a powerful tool for the discovery of evolutionarily<br />

conserved mechanisms and pathways of expression<br />

control. The advantage of cDNA microarrays is that<br />

broad areas of homology are compared and hybridization<br />

probes are sufficiently large that small inter-species<br />

differences in nucleotide sequence do not significantly<br />

affect the analytical results. This comparative genomic<br />

approach allows one to evaluate a common set of genes<br />

within a specific developmental, metabolic, or diseaserelated<br />

gene pathway in various experimental systems.<br />

Using microarrays in studies of mammalian species<br />

other than man and rodents, e.g. non-human primates,<br />

cattle, sheep and pigs, should advance our understanding<br />

of human health and disease. This is already<br />

evident in the field of drug target validation. The lack of<br />

adequate sequence data [7,34] and commercial microarrays<br />

has limited the analysis of economically and<br />

scientifically important large domestic species such as<br />

cattle, pigs and sheep.


S170<br />

This is rapidly changing and cross-species hybridization<br />

experiments based on human sequence based<br />

arrays have been shown to be suitable for comparative<br />

genome expression studies. For example, array analyses<br />

have been performed using monkey, pig, mouse and<br />

salmon RNA on human oligonucleotide arrays [35–39].<br />

An extremely important aspect of cross-species<br />

hybridization is data reproducibility. A poorly conceived<br />

cross-species hybridization experiment leads to<br />

high variability. Statistical variation can be overcome<br />

by increasing the number of samples (biological and<br />

technical replicates) and the arrays themselves can be<br />

improved by careful attention to the use of conserved<br />

(homologous) regions.<br />

We have investigated the use of cross-species<br />

hybridization for orthologous genes within a defined<br />

developmental and metabolic pathway using human and<br />

bovine fetal brain RNA as paired Cy-dye labeled targets<br />

on single human cDNA microarrays [40]. The microarrays<br />

included probes for 349 genes; each spotted 20<br />

times to give robust statistical analysis. Validation of the<br />

results was accomplished by the use of semi-quantitative<br />

RT-PCR. Other approaches to validation include<br />

Real Time PCR and Northern blot analysis. Employing<br />

high stringency hybridization and washing, followed by<br />

analysis of the array data with appropriate statistical<br />

tests, revealed only slight interspecies differences in the<br />

expression patterns and high reproducibility of the<br />

signals for many genes. The correlation coefficient of<br />

cross hybridization between orthologous genes was<br />

Author's personal copy<br />

H. Niemann et al. / Theriogenology 68S (2007) S165–S177<br />

Fig. 2. Reproducibility of a cDNA array analysis with RNA isolated from single mouse embryos. The correlation coefficient of 0.98 indicates that<br />

the array provides reproducible results (for details see Brambrink et al. [31]).<br />

0.94. Verification of the array data by semi-quantitative<br />

RT-PCR using common primer sequences permitted coamplification<br />

of both human and bovine transcripts<br />

[40]. Results of the study demonstrated the power of<br />

comparative genomics and the feasibility of using<br />

human microarrays for the identification of differentially<br />

expressed genes in other species.<br />

3. Current status of embryo transcriptomics<br />

3.1. Embryos from farm animals<br />

Farm animal breeders suffer major economic losses<br />

through high early embryonic mortality, in cattle 35–50%<br />

of the fertilized oocytes fail to develop [41]; this figure is<br />

dramatically increased in modern short-interval breeding<br />

regimens [22]. In many parts of the world, large numbers<br />

of embryos are routinely produced using in vitro<br />

fertilization of oocytes collected by non-surgical ultrasound-guided<br />

follicular aspiration or harvested from<br />

slaughterhouse ovaries. Cloning by somatic cell nuclear<br />

transfer is also well established for this species reflecting,<br />

at least in part, the enormous commercial interest in<br />

improved bovine embryo technology. The abundant<br />

availability of such material provides researchers with<br />

excellent opportunities to study the causes of embryonic<br />

loss in bovine and other species, including man<br />

which show significant similarities in preimplantation<br />

development. Specific similarities include the time<br />

course of preimplantation development; switch from


maternal to embryonic genomic activity, metabolism and<br />

certain pathologies associated with in vitro produced<br />

embryos.<br />

Studies using sensitive Reverse Transcription–Polymerase<br />

Chain Reaction (RT-PCR) revealed approximately<br />

200–250 genes that were differentially<br />

expressed in in vitro derived embryos as compared to<br />

embryos flushed from the uterus [24]. Using this<br />

technology, we also identified expression anomalies in<br />

blastocysts produced from prepubertal oocytes which<br />

appear to correlate with their reduced developmental<br />

potential [42]. Oocytes from prepubertal animals are of<br />

interest because they are the source of embryos in<br />

breeding programs which are based on shortened<br />

generation intervals.<br />

The sequencing of cDNAs, e.g. expressed sequence<br />

tags (EST), from genes specifically expressed during<br />

early development is important for in-depth studies on<br />

the molecular mechanisms underlying mammalian<br />

preimplantation development. This work is well<br />

advanced in mice where >140,000 ESTs from all<br />

developmental stages are available from public databases,<br />

however, in other species, a more limited amount<br />

of sequence information is available [43]. The main<br />

limitation for embryo transcriptomics is that the<br />

establishment of a conventional cDNA EST library<br />

requires roughly 5000 oocytes, 13,500 2-cell stages,<br />

2800 8-cell embryos or 600 blastocysts [44], which is<br />

difficult for a uniparous species.<br />

The advent of cDNA array technology has for the<br />

first time made possible to compare the global<br />

transcription pattern of 20,000 genes during mammalian<br />

embryogenesis limited only by the availability<br />

of cloned cDNA from these early stages. Application of<br />

cDNA array technology has been reported for human,<br />

mouse, bovine and porcine preimplantation embryos. A<br />

study of the transcriptional profile of 2-cell mouse<br />

embryos included 4000 genes from a cDNA-library<br />

and revealed many previously unknown transcripts,<br />

including mobile genetic elements that appear to be<br />

associated with activation of the embryonic genome<br />

[45]. The mRNA expression profile of porcine oocytes,<br />

4- and 8-cell stages and blastocysts was studied with the<br />

aid of a 15 K cDNA-array specific for porcine<br />

reproductive organs. The transcription profile of in<br />

vivo developing embryos was compared with that of in<br />

vitro produced 4-cell, 8-cell and blastocyst stage<br />

embryos. This revealed 1409 and 1696 differentially<br />

expressed (in vivo/in vitro) genes at the 4- and 8-cell<br />

stage, respectively [46].<br />

Studies of bovine embryonic mRNA expression<br />

based on subtractive hybridization and cDNA-arrays<br />

Author's personal copy<br />

H. Niemann et al. / Theriogenology 68S (2007) S165–S177 S171<br />

have identified differentially expressed genes between<br />

bovine oocytes derived from large or small follicles. A<br />

total of 21 upregulated genes were found out of 1000<br />

clones [47]. Using a similar subtraction hybridization<br />

protocol, the sequencing of 185 clones revealed 146<br />

non-redundant genes which were associated with<br />

bovine oocyte maturation [48].<br />

Cross-species hybridization of a human cDNA-<br />

Array with bovine RNA, derived from rather large pools<br />

of >200 oocytes yielded a first expression profile for<br />

bovine oocytes before and after maturation. The relative<br />

amount of mRNA remained constant during in vitromaturation.<br />

Approximately 300 genes could positively<br />

be identified [49]. Recently, we have investigated the<br />

mRNA expression profiles of bovine oocytes and<br />

blastocysts by using a cross-species hybridization<br />

approach employing an array consisting of 15,529<br />

human cDNAs (The ENSEMBL chip) as probe, thus<br />

allowing identifying conserved genes during human and<br />

bovine preimplantation development [50]. The analysis<br />

revealed 419 genes that were expressed in both, oocytes<br />

and blastocysts. The expression of 1324 genes was<br />

detected exclusively in the blastocyst, in contrast to 164<br />

in the oocyte, including a significant number of novel<br />

genes. Typically genes indicative for transcriptional and<br />

translational control such as ELAVL4, TACC3 were<br />

over-expressed in the oocyte, whereas genes indicative<br />

for numerous pathways were over-expressed in blastocysts.<br />

Transcripts implicated in chromatin remodeling<br />

were found in both oocytes (NASP SMARCA2, DNMT1)<br />

and blastocysts (H2AFY, HDAC7A). Pathway analysis<br />

revealed differential expression of genes comprising<br />

107 distinct signaling and metabolic pathways. Expression<br />

patterns in bovine and human blastocysts were to a<br />

large extent identical [50]. A major advantage of our<br />

cross-species hybridization approach is that developmentally<br />

conserved novel and annotated marker genes<br />

could be identified for human and bovine oocytes and<br />

blastocysts.<br />

Analysis with a self constructed bovine array<br />

containing 2000 clones with ESTs from different<br />

embryonic developmental stages (Blue Chip) provided<br />

insight into the expression profile during bovine<br />

embryonic development [51]. Using this Blue Chip,<br />

it was possible to link a specific embryonic transcriptional<br />

profile, determined from embryo biopsies with<br />

the outcome of embryo transfers to recipients. Embryos<br />

leading to pregnancies specifically over-expressed<br />

genes needed for implantation and placenta formation,<br />

carbohydrate metabolism, and growth factor signaling<br />

[52]. Combining subtractive hybridization with this<br />

bovine cDNA-array allowed the identification of


S172<br />

previously unknown bovine oocyte specific genes. From<br />

a total of 850 clones, 491 known genes were identified<br />

in bovine oocytes, plus 188 previously non-characterized<br />

sequences, 105 unknown sequences and 86 noninformative<br />

sequences [53]. Among the 450 clones<br />

derived from bovine blastocysts, 301 known genes, 10<br />

unknown sequences, and 139 non-informative<br />

sequences were positively identified [51]. Other<br />

laboratories have also designed home made arrays<br />

based on either oligonucleotide or cDNA probes<br />

[54,55], which cover only a portion of bovine genome.<br />

These data illustrate the enormous challenges to arrive<br />

at a complete transcriptomic map for bovine preimplantation<br />

development.<br />

Affymetrix commercially distributes a bovine<br />

genomic array covering probe sets of approximately<br />

23,000 transcripts and 19,000 UniGene clusters,<br />

which is based on the bovine genome sequence data.<br />

Recently, the transcriptomes of bovine in vitro matured<br />

oocytes and in vitro produced 8-cell stages were<br />

analyzed by this Affymetrix bovine array and genes<br />

critically involved in the major activation of the<br />

embryonic genome were determined [56]. Approximately<br />

370 genes with functions in chromatin<br />

modification, apoptosis, signal transduction, metabolism<br />

and immune response were differentially<br />

expressed between oocytes and 8-cell stages, including<br />

DNMT1, DNMT2 and MeCP2 that were specifically<br />

upregulated in 8-cell stages.<br />

Large scale transcriptomic analyses are useful for<br />

unraveling the consequences of in vitro production<br />

techniques such as in vitro fertilization and culture and<br />

somatic cloning by comparison with in vivo derived<br />

embryos. Interestingly, the transcriptomes of cloned<br />

bovine embryos were found to be more similar to that of<br />

in vivo blastocysts than to in vitro produced embryos<br />

[57], suggesting that the cloned embryos successfully<br />

underwent major reprogramming. It is well documented<br />

that the in vitro conditions induce alterations in<br />

embryonic transcript profiles [22,23]. Recently, we<br />

have hybridized RNA from in vivo produced bovine<br />

oocytes and all stages of bovine preimplantation<br />

development until the blastocyst stage on the bovine<br />

Affymetrix array. For the first time this analysis will<br />

yield a complete picture of mRNA expression in the<br />

processes of oocyte maturation, fertilization, embryonic<br />

activation, and cell lineage differentiation in a large<br />

mammalian species (unpublished data).<br />

The power of cDNA array technology for unraveling<br />

critical reproductive pathways has also been demonstrated<br />

[58–60]. The developmental potential of bovine<br />

oocytes was correlated with the level of follistatin<br />

Author's personal copy<br />

H. Niemann et al. / Theriogenology 68S (2007) S165–S177<br />

transcripts [60]. Transcript levels of genes involved in<br />

transcription and translation were linked with the origin<br />

of embryos, i.e. either in vivo fertilization or in vitro<br />

production [59]. Another study showed that transcript<br />

abundances for follicle stimulating hormone receptor,<br />

estrogen receptor 2, inhibin alpha, activin A receptor<br />

type I, cyclin D2 and two pro-apoptotic factors<br />

decreased in granulosa cells during the growth of the<br />

dominant follicle [58].<br />

3.2. Embryos from mice and men<br />

Extensive transcriptomic analyses have been made in<br />

mouse oocytes and preimplantation embryos<br />

[31,33,61–64]. Surplus human oocytes or embryos<br />

discarded from IVF procedures have been used in most<br />

transcriptomic studies [65–69]. Recently, RNA was<br />

isolated from pools of human metaphase II oocytes<br />

obtained after hormonal ovarian stimulation [30]. The<br />

globally amplified transcripts were then hybridized to a<br />

human genome chip covering 47,000 transcripts. A<br />

comparison of the hybridization signals with those<br />

obtained for several somatic tissues revealed that 5331<br />

transcripts were up-regulated and 7074 transcripts were<br />

down-regulated in oocytes. Compared with data from<br />

mouse oocytes a common set of 1587 transcripts was<br />

found, which may indicate genes with conserved<br />

function in mammalian oocytes. Common sets of<br />

transcripts were also found in the transcriptomes of<br />

human oocytes and human and murine embryonic stem<br />

cells. Most prominent were transcripts encoding<br />

proteins active in the transforming growth factor<br />

(TGF)-b signaling pathway.<br />

A comprehensive genome-wide study of gene<br />

transcripts has been performed on 12 stages of murine<br />

in vivo oocytes at the germinal vesicle, and metaphase II<br />

stage and preimplantation embryos [33]. More than<br />

4000 genes were differentially expressed during<br />

preimplantation development. Two major groups of<br />

gene activity could be defined, one comprising oocytes<br />

and first cleavage stages, and the other consisting of 4cell<br />

to blastocyst stages, thus correlating with the<br />

transition of maternal to zygotic expression. Several<br />

transcripts of molecules involved in Wnt, bone<br />

morphogenetic protein (BMP) or Notch signaling<br />

pathways were identified in the embryonic transcriptome,<br />

indicating that these critical regulators of cell fate<br />

and patterning are conserved and functional in these<br />

stages of preimplantation development [61–63].<br />

A gene atlas of the protein-coding transcriptomes of<br />

mouse and man has been recently described for several<br />

somatic tissues [70]. On average, 8000 transcripts


were determined for individual mouse tissues; 1–3% of<br />

the analyzed transcripts were detected in all analyzed<br />

tissues, i.e. 79 man and 61 mice, respectively. The<br />

abundance of these ubiquitously expressed transcripts<br />

(housekeeping genes) was found to be significantly<br />

higher than that for all tissue specific genes.<br />

4. Limitations and perspectives of DNA arrays<br />

4.1. Limitations<br />

A major limitation of array technologies is our<br />

fragmentary understanding of genome organization and<br />

function. Recent data suggest that various mechanisms,<br />

including epigenetics, microRNAs, non-coding and<br />

unconventional RNAs significantly contribute to the<br />

dynamics of the transcriptome [10–16]. Even the most<br />

recent arrays with densities of >1 million targets per<br />

array do not cover this complexity. The transcriptome of<br />

somatic cells seems to be regulated in a more stochastic<br />

manner than previously thought [71]. Furthermore, the<br />

enormous plasticity and highly dynamic nature of early<br />

embryonic development may hamper attempts to obtain<br />

a true representative picture of the transcriptome. It has<br />

to be born in mind that current technologies only<br />

provide snapshots of parts of the entire embryonic<br />

transcriptome which may then be further exploited by<br />

arrays specific for certain pathways and/or differentiation<br />

status, etc. Targeted approaches to screen for novel<br />

transcripts via expressed sequence tag libraries may<br />

contribute to a more complete inventory of the<br />

transcriptome [45,72]. However, even in light of these<br />

limitations, the currently available technology allows<br />

discovery of novel genetic information.<br />

Appropriate statistical methods have been developed<br />

for controlling the quality of microarrays, to ensure the<br />

reproducibility of results and to identify differential<br />

expression [73,74]. Bioinformatic methods have been<br />

adapted and refined in the past few years and now take<br />

into account inherent limitations of the array technology.<br />

A recent comprehensive comparison across various<br />

studies indicated high correlation even if different<br />

platforms and bioinformatic tools had been employed<br />

[75], thereby confirming the validity of the microarray<br />

concept across different laboratories.<br />

4.2. Perspectives<br />

DNA array technology is based on the specific<br />

detection of sequences present in certain genes. Tiling<br />

and exon arrays will extend this technology by the<br />

comprehensive identification of individual exons [76] or<br />

Author's personal copy<br />

H. Niemann et al. / Theriogenology 68S (2007) S165–S177 S173<br />

other DNA elements. The human Exon 1.0 ST array<br />

(Affymetrix) is spotted with targets of approximately 1<br />

million known and predicted exons. Exon arrays offer a<br />

fine-tuned picture of gene expression, in view of the fact<br />

that the majority of mammalian genes is differentially<br />

spliced. In extreme cases, such as the neurexin gene<br />

family more than 1000 splice variants of three genes are<br />

known [77,78]. Recently, a close correlation between<br />

array data from exon and conventional gene arrays has<br />

been determined in a cell culture model [79]. This<br />

supports the concept of developing useful exon arrays.<br />

In the context of embryo transcriptomics, exon arrays<br />

will be important to reveal splice variants of genes<br />

critical in early development.<br />

Tiling arrays carry genomic DNA sequences; ideally<br />

they cover the whole genome of an organism. They can<br />

be used to study single nucleotide polymorphisms<br />

(SNP) or copy number variations (CNV), that are<br />

deletions, insertions and duplications from 1 kilobasepairs<br />

to several megabasepairs. Recent studies with<br />

tiling arrays investigated the CNV in the human genome<br />

and provided compelling evidence for an unexpected<br />

high frequency of CNVs [80]. Up to 3000 genes were<br />

found to be associated with CNV. The occurrence of this<br />

phenomenon in mice and ape suggests that a similar<br />

situation may exist in farm animals, which remains to be<br />

shown.<br />

Epigenetic regulation by DNA methylation is a<br />

widespread phenomenon in biology and bears substantial<br />

information with regard to cell physiology,<br />

dysfunction and differentiation. Primary targets of<br />

methylation are cytosine guanine dinucleotides (CpG);<br />

cytosine is modified to become 5 methyl cytosine. In the<br />

mammalian genome CpG sites are unevenly distributed<br />

and mostly clustered as CpG islands, which are often<br />

associated with promoter elements [81]. Cytosine<br />

methylation is a conserved epigenetic mechanism<br />

involved in silencing of transposons, genomic imprinting<br />

and regulation of gene expression. DNA methylation<br />

is critically important for normal development of<br />

mammals, as demonstrated by embryonic lethality of<br />

knock-outs of murine methyltransferase genes DNMT1<br />

or DNMT3a/b. Aberrant methylation patterns are<br />

associated with developmental anomalies and specific<br />

diseases, including cancer [82,83].<br />

Methylation sensitive arrays can detect cytosine<br />

methylation of genomic DNA and may thereby provide<br />

information on the underlying mechanisms of a specific<br />

transcriptional profile. A genome wide identification<br />

and functional analysis of methylated DNA nucleotides<br />

will provide cues on how and to which extent<br />

methylation and gene expression are connected.


S174<br />

Recently the first genome-wide DNA methylation map<br />

has been obtained in the plant Arabidopsis [84].<br />

Methylated and non-methylated genomic DNAs were<br />

fractionated by immuno-precipitation with a monoclonal<br />

antibody that specifically recognizes methylcytosine,<br />

followed by amplification and hybridization to<br />

whole genome arrays. Pericentromeric regions, repetitive<br />

sequences, and regions coding for small interfering<br />

RNA were found to be heavily methylated. Several<br />

constitutively expressed genes showed methylation<br />

within the transcribed regions, but only rarely in<br />

promoter regions, indicating that methylation marks<br />

have different functions in different gene compartments.<br />

Employing a similar strategy, a DNA methylation<br />

profile has been obtained for individual human<br />

chromosomes [85], and the DNA methylation pattern of<br />

the tumor suppressor gene p16INK4A has been<br />

determined [86].<br />

Several human health related technologies rely on<br />

the utilization of non-human primates and farm<br />

animals for vaccine development, organ transplantation,<br />

production of recombinant proteins and viral<br />

pathogenesis [87]. The application of animal products<br />

as drugs or treatments of human patients requires the<br />

absence of any unwanted contaminations. DNA<br />

microarrays provide a mean for a high throughput<br />

testing of these products, including products of<br />

transgenic animals [88].<br />

Ribonucleic acid interference (RNAi) allows<br />

detailed study of biological mechanisms underlying<br />

normal and aberrant embryonic development. RNAi is a<br />

conserved post-transcriptional gene regulatory process<br />

found in most biological systems. The common<br />

mechanistic elements are small interfering RNA<br />

(siRNAs) with 21–23 nucleotides, which specifically<br />

bind complementary sequences of their target mRNAs<br />

and shut down expression. The target mRNAs are<br />

degraded and no protein is translated [89]. The<br />

specificity of the RNAi approach is shown by reduced<br />

levels of the target transcript and unchanged levels of<br />

housekeeping mRNAs [90]. Microarray analysis offers<br />

the possibility to obtain a complete profile after RNAi<br />

application on the downstream effects as well off target<br />

effects in a comprehensive way.<br />

Array based analysis forms the basis for systematic<br />

characterization of the biological function of genes in<br />

preimplantation development by high-throughput RNAi<br />

screening. Genome wide screening in Caenorhabditis<br />

elegans and Drosophila, and to a lesser extent in<br />

mammalian cells, has proven to be a valuable tool for<br />

the dissection of biological pathways. Recent developments<br />

include strategies to combine genome-scale<br />

Author's personal copy<br />

H. Niemann et al. / Theriogenology 68S (2007) S165–S177<br />

RNAi libraries with high-throughput screening technologies<br />

and integrated high-content bioinformatic<br />

analysis [91]. Similarly, analysis of the vast database<br />

of array data obtained from cancer studies will exploit<br />

the similarities between tumor cells and embryonic cells<br />

and will increase understanding of the pathways<br />

involved in embryonic development. For example,<br />

studies in zebrafish show that embryonic and tumorigenic<br />

pathways converge via Nodal signaling [92]. The<br />

bovine embryo with its advanced in vitro techniques and<br />

established technology for recovery of in vivo embryos<br />

will be useful tool in this context. Cross-species<br />

transcriptome analyses will be instrumental to identifying<br />

evolutionary conserved pathways which are crucial<br />

for development. Alteration of specific pathways via<br />

RNAi or transgenesis will reveal physiological relationships<br />

and the genetic factors involved in aberrant<br />

development.<br />

5. Conclusions<br />

Broad determination of the embryonic transcriptome<br />

at each developmental stage through array technology is<br />

critical for understanding the critical regulatory pathways<br />

in early development. This is leading to the<br />

discovery of new regulatory mechanisms and is<br />

identifying new roles for genes known in other contexts.<br />

At present, our understanding of the results of array<br />

analysis is limited by our fragmentary understanding of<br />

pathways involved in the regulation of chromatin<br />

structure and the pathways involved in the regulation of<br />

the transcription of individual genes. Recent developments,<br />

including exon gene arrays, microRNA arrays<br />

and methylation sensitive arrays address some of these<br />

challenges. Furthermore, it should be taken into account<br />

that ultimately the protein determines the function of a<br />

specific gene, and future studies will unravel the<br />

proteome of an embryo by appropriate protein chips. In<br />

the context of mammalian embryonic development, it is<br />

crucial to employ these new tools to obtain a complete<br />

picture of the dynamic embryo transcriptome. Array<br />

technology in its various facets is an important<br />

diagnostic tool for early detection of developmental<br />

aberrations and for improving the safety of new<br />

reproductive technologies such as transgenesis and<br />

somatic nuclear transfer [87].<br />

Acknowledgement<br />

The authors gratefully acknowledge the financial<br />

support from DFG for the embryonic transcriptomic<br />

studies cited in this review (Ni 256/27-1, FOR 478).


References<br />

[1] Goodstadt L, Ponting CP. Phylogenetic reconstruction of orthology,<br />

paralogy, and conserved synteny for dog and human. PLoS<br />

Comput Biol 2006;2:e133.<br />

[2] Aparicio S, Chapman J, Stupka E, Putnam N, Chia JM, Dehal P,<br />

et al. Whole-genome shotgun assembly and analysis of the<br />

genome of Fugu rubripes. Science 2002;297:1301–10.<br />

[3] Gregory SG, Sekhon M, Schein J, Zhao S, Osoegawa K, Scott<br />

CE, et al. A physical map of the mouse genome. Nature<br />

2002;418:743–50.<br />

[4] Okazaki Y, Furuno M, Kasukawa T, Adachi J, Bono H, Kondo S,<br />

et al. Analysis of the mouse transcriptome based on functional<br />

annotation of 60,770 full-length cDNAs. Nature 2002;420:<br />

563–73.<br />

[5] Abril JF, Agarwal P, Alexandersson M, Antonarakis SE,<br />

Baertsch R, Berry E, et al. Initial sequencing and comparative<br />

analysis of the mouse genome. Nature 2002;420:520–62.<br />

[6] Wallis JW, Aerts J, Groenen MA, Crooijmans RP, Layman D,<br />

Graves TA, et al. A physical map of the chicken genome. Nature<br />

2004;432:761–4.<br />

[7] Fadiel A, Anidi I, Eichenbaum KD. Farm animal genomics and<br />

informatics: an update. Nucleic Acids Res 2005;33:6308–18.<br />

[8] Lindblad-Toh K, Wade CM, Mikkelsen TS, Karlsson EK, Jaffe<br />

DB, Kamal M, et al. Genome sequence, comparative analysis<br />

and haplotype of the domestic dog. Nature 2005;438:803–19.<br />

[9] Hayashizaki Y, Kanamori M. Dynamic transcriptome of mice.<br />

Trends Biotechnol 2004;22:161–7.<br />

[10] Biemont C, Vieira C. Junk DNA as evolutionary force. Nature<br />

2006;443:521–4.<br />

[11] Rassoulzadegan M, Grandjean V, Gounon P, Vincent S, Gillot I,<br />

Cuzin F. RNA-mediated non-mendelian inheritance of an epigenetic<br />

change in the mouse. Nature 2006;441:469–74.<br />

[12] Segal E, Fondufe-Mittendorf Y, Chen L, Thastrom A, Field Y,<br />

Moore IK, et al. A genomic code for nucleosome positioning.<br />

Nature 2006;442:772–8.<br />

[13] Venkatesh B, Kirkness EF, Loh YH, Halpern AL, Lee AP,<br />

Johnson J, et al. Ancient noncoding elements conserved in<br />

the human genome. Science 2006;314:765.<br />

[14] Werner A. Natural antisense transcripts. RNA Biol 2005;2:<br />

53–62.<br />

[15] Heintzman ND, Ren B. The gateway to transcription: identifying,<br />

characterizing and understanding promoters in the eukaryotic<br />

genome. Cell Mol Life Sci 2007;64:386–400.<br />

[16] Jirtle RL, Skinner MK. Environmental epigenomics and disease<br />

susceptibility. Nat Rev Genet 2007;8:253–62.<br />

[17] Schena M, Shalon D, Davis RW, Brown PO. Quantitative<br />

monitoring of gene expression patterns with a complementary<br />

DNA microarray. Science 1995;270:467–70.<br />

[18] Bennett ST, Barnes C, Cox A, Davies L, Brown C. Toward the<br />

1,000 dollars human genome. Pharmacogenomics 2005;6:<br />

373–82.<br />

[19] Ragoussis J, Elvidge GP, Kaur K, Colella S. Matrix-assisted laser<br />

desorption/ionization, time-of-flight mass spectrometry in genomics<br />

research. PLoS Genet 2006;2:e100.<br />

[20] Telford NA, Watson AJ, Schultz GA. Transition from maternal to<br />

embryonic control in early development: A comparison of<br />

several species. Mol Reprod Dev 1990;26:90–100.<br />

[21] Kiefer J. Epigenetics in development. Dev Dyn 2007;236:<br />

1144–56.<br />

[22] Niemann H, Wrenzycki C. Alterations of expression of developmentally<br />

important genes in preimplantation bovine embryos<br />

Author's personal copy<br />

H. Niemann et al. / Theriogenology 68S (2007) S165–S177 S175<br />

by in vitro culture conditions: Implications for subsequent<br />

development. Theriogenology 2000;53:21–34.<br />

[23] Wrenzycki C, Herrmann D, Lucas-Hahn A, Korsawe K, Lemme<br />

E, Niemann H. Epigenetic reprogramming throughout preimplantation<br />

development and consequences for assisted reproductive<br />

technologies. Birth Defects Res C 2005;75:1–9.<br />

[24] Wrenzycki C, Herrmann D, Lucas-Hahn A, Korsawe K, Lemme<br />

E, Niemann H. Messenger RNA expression patterns in bovine<br />

embryos derived from in vitro procedures and their implications<br />

for development. Reprod Fertil Dev 2005;17:23–35.<br />

[25] Ge H, Walhout AJM, Vidal M. Integrating ‘‘omic’’ information:<br />

a bridge between genomics and systems biology. Trends Genet<br />

2003;19:551–60.<br />

[26] Owens J. Gene expression: Do microarrays match up? Nat Rev<br />

Genet 2005;6:432–3.<br />

[27] Ma M, Strauss WM. Analysis of the Xist RNA isoforms suggests<br />

two distinctly different forms of regulation. Mamm Genome<br />

2005;16:391–4.<br />

[28] Yu J, Hecht NB, Schultz RM. RNA-binding properties and<br />

translation repression in vitro by germ cell-specific MSY2<br />

protein. Biol Reprod 2002;67:1093–8.<br />

[29] Bilodeau-Goeseels S, Schultz GA. Changes in ribosomal ribonucleic<br />

acid content within in vitro-produced bovine embryos.<br />

Biol Reprod 1997;56:1323–9.<br />

[30] Kocabas AM, Crosby J, Ross PJ, Out HH, Beyhan Z, Can H,<br />

et al. The transcriptome of human oocytes. Proc Natl Acad Sci<br />

USA 2006;103:14027–32.<br />

[31] Brambrink T, Wabnitz P, Halter R, Klocke R, Carnwath JW,<br />

Kues W, et al. Application of cDNA arrays to monitor mRNA<br />

profiles in single preimplantation mouse embryos. Biotechniques<br />

2002;33:376–85.<br />

[32] Klein CA, Seidl S, Petat-Dutter K, Offner S, Geigl JB, Schmidt-<br />

Kittler O, et al. Combined transcriptome and genome analysis of<br />

single micrometastatic cells. Nat Biotechol 2002;20:387–92.<br />

[33] Wang QT, Piotrowska K, Ciemerych MA, Milenkovic L, Scott<br />

MP, Davis RW, et al. A genome wide study of gene activity<br />

reveals developmental signalling pathways in the preimplantation<br />

mouse embryo. Dev Cell 2004;6:133–44.<br />

[34] Niemann H, Kues WA. Application of transgenesis on livestock<br />

for agriculture and biomedicine. Anim Reprod Sci 2003;79:<br />

291–317.<br />

[35] Chismar JD, Mondala T, Fox HS, Roberts E, Langford D,<br />

Masliah E, et al. Analysis of result variability from high-density<br />

oligonucleotide arrays comparing same-species and cross-species<br />

hybridisations. BioTechniques 2002;33:516–24.<br />

[36] Moody DE, Zou Z, McIntyre L. Cross-species hybridisation of<br />

pig RNA to human nylon microarrays. BMC Genomics<br />

2002;3:27.<br />

[37] Medhora M, Bousamra M, Zhu D, Somberg L, Jacobs ER.<br />

Upregulation of collagens detected by gene array in a model<br />

of flow-induced pulmonary vascular remodeling. Am J Physiol<br />

Heart Circ Physiol 2002;282:H414–22.<br />

[38] Huang GS, Yang SM, Hong MY, Yang PC, Liu YC. Differential<br />

gene expression of livers from ApoE deficient mice. Life Sci<br />

2000;68:19–28.<br />

[39] Tsoi SC, Cale JM, Bird IM, Ewart V, Brown LL, Douglas S. Use<br />

of human cDNA microarrays for identification of differentially<br />

expressed genes in Atlantic Salmon liver during Aeromonas<br />

salmonicida infection. Mar Biotechnol (NY) 2003;5:545–54.<br />

[40] Adjaye J, Herwig R, Herrmann D, Wruck W, BenKahla A, Brink<br />

TC, et al. Cross-species hybridisation of human and bovine<br />

orthologous genes on high density cDNA microarrays. BMC-


S176<br />

Genom 2004;5:83 (http://www.biomedcentral.com/1471-2164/<br />

5/83).<br />

[41] Sreenan JM, Diskin MG. The extent and timing of embryonic<br />

mortality in cattle. In: Sreenan JM, Diskin MG, editors.<br />

Embryonic mortality in farm animals. Dordrecht, NL: M. Nijhoff<br />

Publ; 1986. p. 1–11.<br />

[42] Oropeza A, Wrenzycki C, Herrmann D, Hadeler KG, Niemann<br />

H. Improvement of the developmental capacity of oocytes from<br />

prepuberal cattle by intraovarian Insulin-like growth factor-I<br />

application. Biol Reprod 2004;70:1634–43.<br />

[43] Ko MS. Embryogenomics of pre-implantation mammalian<br />

development: current status. Reprod Fertil Dev 2004;16:79–85.<br />

[44] Ko MS. Embryogenomics: developmental biology meets genomics.<br />

Trends Biotechnol 2001;19:511–8.<br />

[45] Evsikov AV, Graber JH, Brockman JM, Hampl A, Holbrook AE,<br />

Singh P, et al. Cracking the egg: molecular dynamics and<br />

evolutionary aspects of the transition from the fully grown<br />

oocyte to embryo. Genes Dev 2006;20:2713–27.<br />

[46] Whitworth KM, Agca C, Kim JG, Patel RV, Springer GK, Bivens<br />

NJ, et al. Transcriptional profiling of pig embryogenesis by using<br />

a 15-K member unigene set specific for pig reproductive tissues<br />

and embryos. Biol Reprod 2005;72:1437–51.<br />

[47] Donnison M, Pfeffer PL. Isolation of genes associated with<br />

developmentally competent bovine oocytes and quantitation of<br />

their levels during development. Biol Reprod 2004;71:1813–21.<br />

[48] Pennetier S, Uzbekova S, Guyader-Joly C, Humblot P, Mermillod<br />

P, Dalbies-Tran R. Genes preferentially expressed in bovine<br />

oocytes revealed by subtractive and suppressive hybridization.<br />

Biol Reprod 2005;73:713–20.<br />

[49] Dalbies-Tran R, Mermillod P. Use of heterologous complementary<br />

DNA array screening to analyze bovine oocyte transcriptome<br />

and its evolution during in vitro maturation. Biol Reprod<br />

2003;68:252–61.<br />

[50] Adjaye J, Herwig R, Brink T, Herrmann D, Greber B, Groth D,<br />

et al. Conserved molecular profiles of maternal and embryonic<br />

gene expression during bovine and human preimplantation<br />

development. Physiol Genom, submitted for publication.<br />

[51] Sirard MA, Dufort I, Vallee M, Massicotte L, Gravel C, Reghenas<br />

H, et al. Potential and limitations of bovine-specific arrays<br />

for the analysis of mRNA levels in early development: preliminary<br />

analysis using a bovine embryonic array. Reprod Fertil<br />

Dev 2005;17:47–57.<br />

[52] El-Sayed A, Hoelker M, Rings F, Salilew D, Jennen D, Tholen E,<br />

et al. Large-scale transcriptional analysis of bovine embryo<br />

biopsies in relation to pregnancy success after transfer to recipients.<br />

Physiol Genom 2006;28:84–96.<br />

[53] Vallee M, Robert C, Methot S, Palin MF, Sirard MA. Crossspecies<br />

hybridizations on a multi-species cDNA microarray to<br />

identify evolutionary conserved genes expressed in oocytes.<br />

BMC Genomics 2006;7:113. doi: 10.1186/1471-2164-7-113.<br />

[54] Long JE, Cai X, He LQ. Gene profiling of cattle blastocysts<br />

derived from nuclear transfer, in vitro fertilization and in vivo<br />

development based on a cDNA library. Anim Reprod Sci<br />

2007;100:243–56.<br />

[55] Yao J, Ren X, Ireland JJ, Coussens PM, Smith TP, Smith GW.<br />

Generation of a bovine oocyte cDNA library and microarray:<br />

resources for identification of genes important for follicular development<br />

and early embryogenesis. Physiol Genom 2004;19:84–92.<br />

[56] Misirlioglu M, Page GP, Sagirkaya H, Parrish JJ, First NL,<br />

Memili E. Dynamics of global transcriptome in bovine matured<br />

oocytes and preimplantation embryos. Proc Natl Acad Sci USA<br />

2006;103:18905–10.<br />

Author's personal copy<br />

H. Niemann et al. / Theriogenology 68S (2007) S165–S177<br />

[57] Smith SL, Everts RE, Tian XC, Du F, Sung LY, Rodriguez-Zas<br />

SL, et al. Global gene expression profiles reveal significant<br />

nuclear reprogramming by the blastocyst stage after cloning.<br />

Proc Natl Acad Sci USA 2005;102:17582–7.<br />

[58] Mihm M, Baker PJ, Ireland JL, Smith GW, Coussens PM, Evans<br />

AC, et al. Molecular evidence that growth of dominant follicles<br />

involves a reduction in follicle-stimulating hormone dependence<br />

and an increase in luteinizing hormone dependence in cattle.<br />

Biol Reprod 2006;74:1051–9.<br />

[59] Corcoran D, Fair T, Park S, Rizos D, Patel OV, Smith GW, et al.<br />

Suppressed expression of genes involved in transcription and<br />

translation in in vitro compared with in vivo cultured bovine<br />

embryos. Reproduction 2006;131:651–60.<br />

[60] Patel OV, Bettegowda A, Ireland JJ, Coussens PM, Lonergan P,<br />

Smith GW. Functional genomics studies of oocyte competence:<br />

evidence that reduced transcript abundance for follistatin is<br />

associated with poor developmental competence of bovine<br />

oocytes. Reproduction 2007;133:95–106.<br />

[61] Zeng F, Baldwin DA, Schultz RM. Transcript profiling during<br />

preimplantation mouse development. Dev Biol 2004;272:<br />

483–96.<br />

[62] Zeng F, Schultz RM. RNA transcript profiling during zygotic<br />

gene activation in the preimplantation mouse embryo. Dev Biol<br />

2005;283:40–57.<br />

[63] Hamatani T, Daikoku T, Wang H, Matsumoto H, Carter MG, Ko<br />

MS, et al. Global gene expression analysis identifies molecular<br />

pathways distinguishing blastocyst dormancy and activation.<br />

Proc Natl Acad Sci USA 2004;101:10326–31.<br />

[64] Hamatani T, Ko MS, Yamada M, Kuji N, Mizusawa Y, Shoji M,<br />

et al. Global gene expression profiling of preimplantation<br />

embryos. Hum Cell 2006;19:98–117.<br />

[65] Adjaye J, Daniels R, Monk M. The construction of cDNA<br />

libraries from human single preimplantation embryos and their<br />

use in the study of gene expression during development. J Assist<br />

Reprod Genet 1998;15:344–8.<br />

[66] Chen HW, Chen JJ, Yu SL, Li HN, Yang PC, Su CM, et al.<br />

Transcriptome analysis in blastocyst hatching by cDNA microarray.<br />

Hum Reprod 2005;20:2492–501.<br />

[67] Bermudez MG, Wells D, Malter H, Munne S, Cohen J, Steuerwald<br />

NM. Expression profiles of individual human oocytes using<br />

microarray technology. Reprod Biomed Online 2004;8:325–37.<br />

[68] Dobson AT, Raja R, Abeyta MJ, Taylor T, Shen S, Haqq C, et al.<br />

The unique transcriptome through day 3 of human preimplantation<br />

development. Hum Mol Genet 2004;13:1461–70.<br />

[69] Assou S, Anahory T, Pantesco V, Le Carrour T, Pellestor F, Klein<br />

B, et al. The human cumulus-oocyte gene-expression profile.<br />

Hum Reprod 2006;21:1705–19.<br />

[70] Su AI, Wiltshire T, Batalov S, Lapp H, Ching KA, Block D, et al.<br />

A gene atlas of the mouse and human protein-encoding transcriptomes.<br />

Proc Natl Acad Sci USA 2004;101:6062–7.<br />

[71] Raj A, Peskin CS, Tranchina D, Vargas DY, Tyagi S. Stochastic<br />

mRNA synthesis in mammalian cells. PLoS Biol 2006;4:e309.<br />

[72] Sharov AA, Piao Y, Matoba R, Dudekula DB, Qian Y, VanBuren<br />

V, et al. Transcriptome analysis of mouse stem cells and early<br />

embryos. PLoS Biol 2003;1:E74.<br />

[73] Peri S, Ibarrola N, Blagoev B, Mann M, Pandey A. Common<br />

pitfalls in bioinformatics-based analyses: look before you leap.<br />

Trends Genet 2001;17:541–5.<br />

[74] Nadon R, Shoemaker J. Statistical issues with microarrays:<br />

processing and analysis. Trends Genet 2002;18:265–71.<br />

[75] Shi L, Reid LH, Jones WD, Shippy R, Warrington JA, Baker SC,<br />

et al. The MicroArray Quality Control (MAQC) project shows


inter- and intraplatform reproducibility of gene expression measurements.<br />

Nat Biotechnol 2006;24:1151–61.<br />

[76] Yeakley J, Fan JB, Doucet D, Luo L, Wickham E, Ye Z, et al.<br />

Profiling alternative spliceing on fiber-optic arrays. Nat Biotechnol<br />

2002;20:353–8.<br />

[77] Ullrich B, Ushkaryov YA, Sudhof TC. Cartography of neurexins:<br />

more than 1000 isoforms generated by alternative splicing and<br />

expressed indistinct subsets of neurons.Neuron1995;14:497–507.<br />

[78] Missler M, Sudhof TC. Neurexins: three genes and 1001 products.<br />

Trends Genet 1998;14:20–6.<br />

[79] Okoniewski MJ, Hey Y, Pepper SD, Miller CJ. High correspondence<br />

between Affymetrix exon and standard expression arrays.<br />

Biotechniques 2007;42:181–5.<br />

[80] Kehrer-Sawatzki H. What a difference copy number variation<br />

makes. Bioessays 2007;29:311–3.<br />

[81] Antequera F, Bird A. Number of CpG islands and genes in<br />

human and mouse. Proc Natl Acad Sci USA 1993;90:11995–9.<br />

[82] Silva JM, Dominguez G, Villanueva MJ, Gonzalez R, Garcia JM,<br />

Corbachio C, et al. Aberrant DNA methylation of the p16INK4a<br />

geneinplasmaDNAofbreastcancer.BrJCancer1999;80:1262–4.<br />

[83] Zhou HZ, Yu BM, Wang ZW, Sun JY, Cang H, Gao F, et al.<br />

Detection of aberrant p16 methylation in the serum of colorectal<br />

cancer patients. Clin Cancer Res 1999;8:188–91.<br />

[84] Zhang X, Yazaki J, Sundaresan A. Genome-wide high resolution<br />

mapping and functional analysis of DNA methylation in Arabidopsis.<br />

Cell 2006;126:1–13.<br />

Author's personal copy<br />

H. Niemann et al. / Theriogenology 68S (2007) S165–S177 S177<br />

[85] Eckhardt F, Lewin J, Cortese R, Rakyan VK, Attwood J, Burger<br />

M, et al. DNA methylation profiling of human chromosomes 6,<br />

20 and 22. Nat Genet 2006;38:1378–85.<br />

[86] Mund C, Beier V, Bewerunge P, Dahms M, Lyko F, Hoheisel JD.<br />

Array-based analysis of genomic DNA methylation patterns of<br />

the tumour suppressor gene p16INK4A promoter in colon<br />

carcinoma cell lines. Nucl Acids Res 2005;33:e73. doi:<br />

10.1093/nar/gni072.<br />

[87] Kues WA, Niemann H. The contribution of farm animals to<br />

human health. Trends Biotechnol 2004;22:286–94.<br />

[88] Kues WA, Schwinzer R, Wirth D, Verhoeyen E, Lemme E,<br />

Herrmann D, et al. Epigenetic silencing and tissue independent<br />

expression of a novel tetracycline inducible system in doubletransgenic<br />

pigs. FASEB J Express 2006;20:E1–0. doi: 10.1096/<br />

fj.05-5415fje.<br />

[89] Plasterk RH. RNA silencing in genome’s immune system.<br />

Science 2005;296:1263–5.<br />

[90] Iqbal K, Kues WA, Niemann H. Parent-of-origin dependent gene<br />

specific knock down in mouse embryos. Biochem Biophys Res<br />

Comm, in press.<br />

[91] Fuchs F, Boutros M. Cellular phenotyping by RNAi. Funct<br />

Genom Proteomics 2006;5:52–6.<br />

[92] Topczewska JM, Postovit LM, Margaryan NY, Sam A, Hess AR,<br />

Wheaton WW, et al. Embryonic and tumorigenic pathways<br />

converge via Nodal signalling: role in melanoma aggressiveness.<br />

Nat Med 2006;12:925–32.


Review<br />

Reproductive biotechnology in farm animals goes genomics<br />

Wilfried A. Kues, Detlef Rath and Heiner Niemann*<br />

Address: Department of Biotechnology, Institute of Farm Animal Genetics (FLI), Mariensee, Höltystrasse 10, D-31535 Neustadt,<br />

Germany.<br />

*Correspondence: Heiner Niemann. Fax: +49 (0)5034-871-101. Email: heiner.niemann@fli.bund.de<br />

Received: 4 February 2008<br />

Accepted: 23 May 2008<br />

doi: 10.1079/PAVSNNR20083036<br />

The electronic version of this article is the definitive one. It is located here: http://www.cababstractsplus.org/cabreviews<br />

g CAB International 2008 (Online ISSN 1749-8848)<br />

Abstract<br />

CAB Reviews: Perspectives in Agriculture, Veterinary Science, Nutrition and Natural Resources 2008 3, No. 036<br />

The breeding of domestic animals has a longstanding and successful history, starting with<br />

domestication several thousand years ago. Modern animal breeding strategies, predominantly<br />

based on population genetics, artificial insemination (AI) and embryo transfer (ET) technologies led<br />

to significant increases in the performances in domestic animals and are the basis for the regular<br />

supply with high-quality animal-derived food at acceptable prices. A recent addition to this arsenal<br />

of technologies is the separation of spermatozoa into X- and Y-chromosomes bearing populations<br />

by means of flow cytometry, which is going to be transferred into commercial cattle breeding.<br />

Moreover, reproductive biotechnologies are currently <strong>bei</strong>ng complemented by new measures of<br />

molecular genetics. The genomes of several domestic animals have been sequenced and annotated.<br />

Important uses of genomic data include improving selective breeding strategies, studying expression<br />

patterns (transcriptomics or proteomics) and the production of transgenic animals<br />

tailored for specific production purposes. The technology for the production of transgenic<br />

animals, either by additive gene transfer or by selective deletion (knockout), has been advanced to<br />

the level of practical application. A crucial step in this context was the development of somatic cell<br />

nuclear transfer (SCNT) cloning, which has replaced the previously used rather inefficient method<br />

of microinjection for producing transgenic animals. In this review, we discuss recent advances in<br />

important areas of animal biotechnology related to agricultural applications, including sperm<br />

sexing, somatic cloning, genome and transcriptome research, transgenesis and recombinant DNA<br />

vaccines with emphasis on large domestic species.<br />

Keywords: Domestication, Reproductive techniques, Genome sequencing, Sperm sexing, Recombinant<br />

DNA, Somatic cloning, Embryo transcriptomics, Transgenic livestock, Agriculture<br />

Introduction: Evolution of Animal Breeding<br />

and Biotechnology<br />

The breeding of domestic animals has a longstanding and<br />

successful history, starting with domestication several<br />

thousand years ago, by which Man kept animals in his<br />

proximity and used products thereof [1]. Using the<br />

technical options that were available in each time period,<br />

humans have propagated those populations that were<br />

deemed useful for their respective needs and purposes.<br />

Selection mostly occurred according to the phenotype<br />

and/or because of specific traits. Scientifically based animal<br />

breeding has only existed for 50 years, mainly based on<br />

http://www.cababstractsplus.org/cabreviews<br />

population genetics and statistics. A vast number of phenotypically<br />

different breeds with desirable traits resulted<br />

from this process in a short time period (in evolutionary<br />

terms). A good example is cattle, for which nowadays<br />

>800 cattle breeds can be found worldwide [2]. These<br />

breeds are not just variations of the same archetype, but<br />

differ in qualitative and quantitative characteristics,<br />

including specific disease resistance, climate adaptation,<br />

lactation performance, fertility, meat quality and nutrient<br />

requirements [3].<br />

Modern animal-breeding strategies are based on a<br />

number of biotechnological procedures, of which artificial<br />

insemination (AI) is the most prominent example. Today,


2 Perspectives in Agriculture, Veterinary Science, Nutrition and Natural Resources<br />

AI is employed in more than 90% of all sexually mature<br />

female dairy cattle in countries with advanced breeding<br />

programs [4]. Application of AI is also steadily increasing<br />

in pigs, where now on a global scale approximately 50%<br />

of sexually matured sows are fertilized by AI. AI allows<br />

effective propagation of the genetic potential of valuable<br />

sires within a population. On an average, 200–500 insemination<br />

doses can be produced from one bull ejaculate,<br />

which can be successfully stored frozen; the corresponding<br />

figure for the boar is 10–35 [4]. A more recent<br />

development is the separation of spermatozoa into X- and<br />

Y-chromosomes bearing populations to allow the use of<br />

semen carrying predetermined sex chromosomes [5].<br />

In the second half of the 1980s, embryo transfer (ET)<br />

technology has been introduced into animal breeding. ET<br />

allowed for the first time a better exploitation of the<br />

genetic potential of female animals. On an average, 5–10<br />

viable embryos are <strong>bei</strong>ng collected from a superovulated<br />

donor cow and approximately 600 000 in vivo collected<br />

embryos are transferred worldwide, of which 50% are<br />

used after freezing and thawing; in addition, a significant<br />

number of the transferred embryos ( 300 000) are<br />

derived from in vitro production (in vitro maturation,<br />

fertilization and culture of embryos), i.e. mainly from<br />

abattoir ovaries [6]. In contrast to AI, ET is predominantly<br />

used in the top 1% of a breeding population.<br />

These breeding strategies, predominantly based on<br />

population genetics, AI and ET technologies led to significant<br />

increases in the performances in domestic animals<br />

and are the basis for the regular supply with high-quality<br />

animal-derived food at acceptable prices. However, three<br />

major disadvantages have to be taken into account:<br />

(1) the genetic progress is slow, at only 1–3% per year;<br />

(2) it is not possible to separate the desired traits from<br />

undesired ones; and<br />

(3) a targeted transfer of genetic information between<br />

different species is not possible.<br />

This situation is about to change. Reproductive biotechnologies<br />

are currently <strong>bei</strong>ng complemented by new<br />

measures of molecular genetics. The genomes of several<br />

domestic animals have been sequenced and annotated,<br />

including cattle, chicken, horse, bee and dog, and the pig<br />

genome is expected to be drafted shortly (www.<br />

ensembl.org). The technology for the production of<br />

transgenic farm animals, either by additive gene transfer<br />

or by selective deletion (knockout), has been advanced to<br />

the level of practical application ([7–11]; reviewed in<br />

[12]). A crucial step in this context was the development<br />

of somatic cell nuclear transfer (SCNT) cloning [13],<br />

which has replaced the previously used rather inefficient<br />

method of microinjection for producing transgenic animals<br />

[14]. Transgenic animals have been advanced to<br />

practical application in biomedicine, in particular for the<br />

production of pharmaceutical proteins in milk (gene<br />

pharming: GTC Biotherapeutics, Pharming, Biosidus etc.)<br />

and xenotransplantation, i.e. the production of porcine<br />

organs and tissue in human organ transplantation to cope<br />

with the shortage of human organs in terminally ill<br />

patients. Agricultural applications are now rapidly emerging.<br />

Here, we discuss recent advances in important areas of<br />

animal biotechnology related to agricultural applications,<br />

including sperm sexing, somatic cloning, genome and<br />

transcriptome research, transgenesis and recombinant<br />

DNA vaccines with emphasis on large domestic species.<br />

Sperm Sexing<br />

http://www.cababstractsplus.org/cabreviews<br />

The reliable control of the sex ratio permits faster genetic<br />

progress, higher productivity, improves animal welfare by<br />

reducing the incidence of dystocia in cattle, avoiding castration<br />

of male pigs, and producing less environmental<br />

impact through elimination of animals with the unwanted<br />

sex. In mammals, the sex-determining region (SRY) has<br />

been mapped to the short arm of the Y-chromosome<br />

[15]. SRY induces the formation of the primary testicular<br />

tissue in the genital ridge, with the further development to<br />

male reproductive organs <strong>bei</strong>ng primarily under hormonal<br />

control. Sperm from mammalian livestock species show a<br />

difference in the relative amount of DNA between X- and<br />

Y-chromosomes bearing sperm in the range of 3.5–4.2%<br />

[16]. This difference in DNA contents between X- and<br />

Y-bearing sperm can be exploited to effectively separate<br />

populations of X- and Y-sperm in mammals via high-speed<br />

flow cytometry [17]. The sample preparation requires<br />

sperm staining with the fluorochrome Hoechst 33342<br />

[18], which is a vital stain that is able to penetrate the<br />

membrane of live sperm. After saturated binding of the<br />

dye to A–T rich regions of genomic DNA, fluorescence<br />

detection of small differences in relative DNA contents is<br />

possible [18]. Damaged sperm can be excluded by sorting<br />

through prior selective staining with FD#40, a red food<br />

dye [19]. The current high-speed sperm sorters operate<br />

at sample pressures up to 4.22 kg/cm 2 and a droplet frequency<br />

of about 66 kHz, allowing identification of 30 000<br />

events/s.<br />

Correctly oriented spermatozoa are eliminated from<br />

the system through the vibrating nozzle tip embedded in<br />

small fluid droplets. The UV-light excites the fluorescent<br />

dye, emitted light is collected by the 0 and 90 detectors<br />

and the electric signals are processed by a computer.<br />

Charged droplets disintegrate from the continuous<br />

stream and pass an electro-static field (3000 V), which<br />

allows separate collection of both the streams into tubes<br />

pre-filled with a medium. Sorted spermatozoa are subsequently<br />

washed from the sheath fluid by centrifugation<br />

and the remaining sperm pellet is extended in a<br />

suitable medium for further preservation [17, 20, 21]. The<br />

current models of modified high-speed flow sorters<br />

produce sexed sperm at the maximum rate of 20 million<br />

sorted spermatozoa per hour under optimal conditions


with purities above 90%. Purity of sorted samples can<br />

be validated either by flow cytometrical reanalysis<br />

[22, 23], fluorescence in situ hybridization (FISH) [24, 25],<br />

or PCR [26].<br />

The technology was first applied commercially in the<br />

UK in 2000, and approximately two million calves have<br />

been produced from insemination with sexed sperm since<br />

then (Rath and Johnson, in press). Although the sexing<br />

technology has become more efficient over the past<br />

decade, the number of sperm needed for regular inseminations<br />

( 20 million sperm in cattle) cannot be flowsorted<br />

for large-scale commercial programmes. Thus,<br />

protocols for effective low-dose insemination have been<br />

developed. At least two million sperm are needed for<br />

commercial application of sex-separated bovine sperm<br />

[27]. However, this may be too low to obtain satisfactory<br />

pregnancy rates in light of the individual sperm differences<br />

of bulls to fertilize at such low concentrations [28]. This is<br />

even more pronounced in sorted-frozen spermatozoa<br />

that have a reduced life span, as indicated in thermotolerance<br />

tests [29]. In the pig, at least 50 million sperm<br />

are needed for deep intra-uterine AI to achieve<br />

acceptable pregnancy rates, thus preventing broader<br />

applications [30]. Reports on the fertility of sex-sorted<br />

semen are contradictory with that of some authors who<br />

find no decline of fertility and others reporting a reduced<br />

fertility after insemination with frozen sex-sorted spermatozoa<br />

[5, 31–35].<br />

The lower fertility can be caused by physical and/or<br />

chemical stress during sorting, and/or can be caused by<br />

the low number of inseminated sperm. High laser intensity<br />

may cause more damage than lower laser intensity<br />

as shown for rabbit [36] and bull spermatozoa [37].<br />

Employing the most advanced laser technology and<br />

replacing the water-cooled argon gas laser with a pulsed<br />

solid-state laser significantly reduces stress on sperm. The<br />

exposure of sperm to light pulses is much shorter than for<br />

the continuous argon laser beam. Preliminary observations<br />

with sorted ram spermatozoa using this new technology<br />

indicate increased fertility rates (WMC Maxwell,<br />

unpublished).<br />

Importantly, no changes were found in the frequency<br />

of endogenous DNA nicks after staining and UV light<br />

exposure, indicating that the dye is not genotoxic [38, 39].<br />

The hydrodynamic pressure could also exert detrimental<br />

stress on the sorted sperm. Lowering the pressure from<br />

2.59 to 2.07 mm Hg increased developmental rates of<br />

bovine in vitro fertilization (IVF) embryos produced from<br />

sex-sorted semen [40, 41].<br />

Recently, we demonstrated for the first time that<br />

bovine IVF-blastocysts derived from sex-sorted spermatozoa<br />

showed an aberrant mRNA expression pattern<br />

of developmentally important genes such as glucose<br />

transporter 3 (Glut3) and glucose-6-phosphate dehydrogenase<br />

(G6PDX), compared with their counterparts<br />

derived from unsorted semen [42]. This indicates a longer<br />

lasting effect of sorting into preimplantation development.<br />

http://www.cababstractsplus.org/cabreviews<br />

Wilfried A. Kues, Detlef Rath and Heiner Niemann 3<br />

Table 1 Birth of offspring (+) using sex sorted sperm with<br />

different fertilization techniques<br />

AI/DUI ICSI IVF<br />

Cattle + +<br />

Pig + + +<br />

Horse +<br />

Sheep + + +<br />

Goat +<br />

Dolphins +<br />

Rabbit + +<br />

Buffalo +<br />

Elk +<br />

Primates + + +<br />

Cat +<br />

AI/DUI, artificial insemination/deep intrauterine insemination;<br />

ICSI, intracytoplasmic sperm injection; IVF, in vitro fertilization.<br />

Moreover, cleavage and blastocyst rates after IVF with<br />

sex-sorted spermatozoa were significantly lower than that<br />

in embryos produced with unsorted spermatozoa of the<br />

same ejaculate [43]. Earlier reports had indicated that the<br />

number of cell cycles is reduced [35] and timing of<br />

embryo development may be disturbed after insemination<br />

with sorted semen [33, 44, 45].<br />

One approach to improve field results with sorted<br />

sperm would be to select sorted sperm populations that<br />

are specifically suited for successful insemination. With<br />

the aid of a 6-h thermotolerance test, sorted sperm<br />

populations were detected that showed improved fertility<br />

characteristics and could thus be better suited for insemination<br />

[29, 46]. The use of a novel sperm-preservation<br />

methodology (Sexcess 1 ), based on a new extender supplemented<br />

with antioxidants, significantly increased lifespan<br />

and motility of bull spermatozoa. Pregnancy rates of<br />

heifers inseminated with sex-separated semen treated<br />

with Sexcess 1 reached similar levels as for animals inseminated<br />

with non-sex-separated semen (73.6% sorted<br />

versus 76.7% controls; n=232) [47]. In the pig, it was<br />

shown that the media that had been supplemented with<br />

PSP (porcine seminal plasma) I/II heterodimers [48], which<br />

increases motility and mitochondrial activity [49], raised<br />

the proportion of viable sorted boar spermatozoa. More<br />

research is underway to improve bovine field results with<br />

sex-sorted semen by a better preselection and protection<br />

of spermatozoa during sorting and cryopreservation [50],<br />

pig [51–53], horse [54, 55].<br />

IVF, intracytoplasmic sperm injection (ICSI) and intratubal<br />

insemination are promising approaches to adjust the<br />

sorting technology in pigs and horses, where the high<br />

number of viable spermatozoa needed for regular fertilization<br />

cannot be produced, even for deep intrauterine<br />

insemination of sorted spermatozoa (Table 1) [21, 48,<br />

56–59].<br />

In conclusion, flow-cytometry-based separation of<br />

sperm into X- and Y-chromosomes bearing populations


4 Perspectives in Agriculture, Veterinary Science, Nutrition and Natural Resources<br />

Figure 1 Schematic drawing illustrating the use of somatic cloning for the production of transgenic animals. After the<br />

transfer of the transgenic donor cell into perivitelline space of the enucleated oocyte, both the components are fused usually<br />

by short, high-voltage pulses through the point of contact between the two cells. This is followed by activation of the<br />

reconstructed embryos, which is achieved either by short electrical pulses, or by brief exposure to chemical substances,<br />

such as ionomycin or dimethylaminopurin (DMAP), regulating the calcium influx into the complexes and/or the cell cycle.<br />

After a temporary in vitro culture period, the resulting blastocysts are transferred to synchronized recipient animals.<br />

already meets the requirements for field application of<br />

bovine semen when heifers are used and semen from<br />

preselected bulls is processed using the most advanced<br />

protocols. For yet-unknown reasons, cows have still<br />

lower conception rates after insemination with sex-sorted<br />

semen as compared with AI in heifers. Nevertheless,<br />

the current sexing technology is commercially available<br />

for cattle breeding and could evolve as a major tool<br />

towards the implementation of more efficient breeding<br />

and production schemes. More research is needed to<br />

allow commercial application of sex-sorted semen for<br />

other livestock species.<br />

Somatic Cloning of Farm Animals<br />

The technical aspects of nuclear transfer in mammals have<br />

already been established in the 1980s using blastomeres<br />

from early-stage embryos as donor cells [60, 61]. The<br />

assumption was that the closer the nuclear donor is<br />

developmentally to the early embryonic stage, the more<br />

successful nuclear transfer is likely to be. This was mainly<br />

based on earlier findings in frogs, in which cloning with<br />

embryonic cells was possible with the production of<br />

normal offspring, whereas cloning with adult cells was<br />

only compatible with the development into tadpoles [62].<br />

This assumption dominated biology until the birth of<br />

‘Dolly’ in 1996 – the first cloned sheep generated from an<br />

adult mammary gland cell [13]. Common somatic cloning<br />

http://www.cababstractsplus.org/cabreviews<br />

protocols involve the following major technical steps<br />

(Figure 1):<br />

(1) Enucleation of the recipient oocyte: Oocytes at the<br />

metaphase II (MII) stage rather than any other<br />

developmental stage are the most appropriate recipients<br />

for the production of viable cloned mammalian<br />

embryos. These oocytes possess high levels of<br />

maturation promoting factor (MPF), which is thought<br />

to be critical for development of the reconstructed<br />

embryo [63]. In many domesticated species, oocytes<br />

can be obtained from abattoir ovaries, thus providing<br />

a potentially unlimited source of material for cloning<br />

experiments. Oocytes are enucleated by sucking or<br />

squeezing out a small portion of the oocyte cytoplasm,<br />

specifically the portion closely apposed to the<br />

first polar body, where the MII chromosomes are<br />

usually located.<br />

(2) Preparation and subzonal transfer of the donor cell: The<br />

entire intact donor cell, i.e. nucleus plus cytoplasm, is<br />

isolated from a cell culture dish by trypsin treatment<br />

and is inserted under the zona pellucida in intimate<br />

contact with the cytoplasmic membrane of the oocyte<br />

with the aid of an appropriate micropipette. Various<br />

somatic cells, including mammary epithelial cells,<br />

cumulus cells, oviductal cells, leucocytes, hepatocytes,<br />

granulosa cells, epithelial cells, myocytes, neurons,<br />

lymphocytes and germ cells, have successfully been<br />

used as donors for the production of cloned animals


[63–66]. Fetal cells, specifically fibroblasts, have frequently<br />

been used in somatic cloning experiments,<br />

because they are thought to have less genetic damage<br />

and a higher proliferative capacity than adult somatic<br />

cells. One option to improve cloning efficiency could<br />

be using less-differentiated cells (foetal or stem cells)<br />

to minimize the complicated and error-prone reprogramming<br />

process. Somatic stem cells have been used<br />

successfully to give porcine blastocyst development<br />

and the birth of live piglets [67, 68].<br />

In most experiments, donor cells are induced to<br />

exit the cell cycle by serum deprivation, which arrests<br />

cells at the G0/G1 cell cycle stage [69]. Serum<br />

deprivation of foetal fibroblasts for more than 72 h<br />

has been found to be associated with increased<br />

apoptosis rates [70, 71]. Specific cyclin-dependent<br />

kinases, such as roscovitin, have been reported to<br />

increase the efficiency of the cloning process,<br />

although final evidence in the form of a healthy offspring<br />

is lacking [63]. Nevertheless, unsynchronized<br />

somatic donor cells have been successfully used to<br />

clone offspring in mice and cattle [72, 73].<br />

(3) Fusion of the two components: Enucleated oocyte and<br />

donor cell are fused, usually by short, high-voltage<br />

pulses through a point of contact between the two<br />

cells.<br />

(4) Activation of the reconstructed complex: Activation is<br />

achieved either by short electrical pulses, or by brief<br />

exposure to chemical substances such as ionomycin<br />

or dimethylaminopurin (DMAP), regulating the calcium<br />

influx into the complexes and/or the cell cycle.<br />

(5) Temporary culture of the reconstructed embryos: Cloned<br />

embryos can be cultured in vitro to the blastocyst<br />

stage (5–7 days) to assess the initial developmental<br />

competence prior to transfer into a foster mother.<br />

Culture systems for ruminant embryos are quite<br />

advanced and allow the routine production of 30–40%<br />

blastocysts from in vitro-fertilized oocytes isolated<br />

from abattoir ovaries [74, 75].<br />

(6) Transfer to a foster mother or storage in liquid nitrogen:<br />

Bovine embryos at the morula and blastocyst<br />

stage can be transferred non-surgically to the uterine<br />

horns of synchronized recipients using established<br />

embryo-transfer procedures. In pigs, the activated<br />

nuclear transfer complexes are immediately transferred<br />

into the oviducts of the recipient because<br />

the in vitro culture systems for embryos are not yet<br />

as effective as in cattle. This requires a surgical<br />

intervention under general anaesthesia.<br />

To date, somatic nuclear transfer has been successful in<br />

13 mammalian species (sheep [13], cow [72], mouse [76],<br />

goat [77], pig [78–80], rabbit [81], cat [82], horse [83],<br />

mule [84], rat [85], dog [86], ferret [87], wolf [88]) and<br />

also in the generation of cloned offspring from endangered<br />

species and breeds [89–92]. Attempts to preserve<br />

livestock genetic resources by somatic cell banking have<br />

http://www.cababstractsplus.org/cabreviews<br />

Wilfried A. Kues, Detlef Rath and Heiner Niemann 5<br />

been reported [93, 94]. However, the typical success rate<br />

(live births) of mammalian somatic nuclear transfer is low:<br />

usually only 1–2%. Cattle seem to be an exception<br />

to this rule as levels of 15–20% can be reached [95].<br />

Recently, we have obtained significant improvement of<br />

porcine cloning success by a better selection and an<br />

optimized treatment of the recipients, specifically by<br />

providing a 24 h asynchrony between the pre-ovulatory<br />

oviducts of the recipients and the reconstructed embryos.<br />

Presumably, this gave the embryos additional time to<br />

achieve the necessary level of nuclear reprogramming.<br />

This resulted in pregnancy rates of 80% and only slightly<br />

reduced litter size [96].<br />

Pre- and post-natal development can be compromised<br />

and a variable proportion of the offspring in ruminants<br />

and mice show aberrant developmental patterns and<br />

increased perinatal mortality. These abnormalities include<br />

a wide range of symptoms, such as extended gestation<br />

length, oversized offspring, aberrant placenta, cardiovascular<br />

problems, respiratory defects, immunological deficiencies,<br />

problems with tendons, adult obesity, kidney and<br />

hepatic malfunctions, behavioural changes and a higher<br />

susceptibility to neonatal diseases and are summarized as<br />

‘Large Offspring Syndrome’ (LOS) [97–101]. The incidence<br />

is stochastic and has not been correlated with<br />

aberrant expression of single genes or specific pathophysiology.<br />

A new term ‘Abnormal Offspring Syndrome’<br />

(AOS) with sub classification according to the outcome<br />

of such pregnancies has recently been proposed to<br />

reflect better the broad spectrum of this pathological<br />

phenomenon [102]. The general assumption is that the<br />

underlying cause for these pathologies is insufficient or<br />

faulty reprogramming of the transferred somatic cell<br />

nucleus.<br />

However, a critical survey of the published literature on<br />

cloning of animals revealed that al<strong>bei</strong>t the higher pre- and<br />

neonatal death rates of clones, most postnatal clones are<br />

healthy and develop normally [103, 104]. This is consistent<br />

with the finding that mammalian development is<br />

rather tolerant of minor epigenetic aberrations in the<br />

genome and subtle abnormalities in gene expression do<br />

not interfere with the survival of cloned animals [105].<br />

It has become clear that once cloned offspring have survived<br />

the neonatal period and are approximately 6 months<br />

of age (cattle, sheep), they are not different from agematched<br />

controls with regard to numerous biochemical<br />

blood and urine parameters [106, 107], immune status<br />

[106], body score [106], somatotrophic axis [108], reproductive<br />

parameters [109], and yields and composition<br />

of milk [110]. No differences were found in meat and milk<br />

composition of bovine clones when compared with agematched<br />

counterparts; all parameters were within the<br />

normal range [111–113]. Similar findings were reported<br />

for cloned pigs [114]. There is a unanimous consent<br />

across various regulatory agencies around the world<br />

that food derived from cloned animals is safe and there<br />

is no scientific basis for questioning this [115], according


6 Perspectives in Agriculture, Veterinary Science, Nutrition and Natural Resources<br />

Table 2 Important genomic features of selected livestock species and dog<br />

Species<br />

Genome<br />

size (10 9 bp) Chr. PC genes Pseudogenes<br />

Cow 3.25 29+X/Y 21 800 1296 2496<br />

Dog 2.39 38+X/Y 19 300 1742 2448<br />

Chicken 1.05 32+W/Z 16 700 96 655<br />

to the expert committee from the Japanese Ministry of<br />

Agriculture, Forestry and Fisheries (MAFF) [116], the<br />

FDA (Food and Drug Administration) [117] and EFSA<br />

(European Food Safety Agency) [118]. However, given the<br />

limited experience with somatic cloning (which has only<br />

been in general use since 1997) and the relatively long<br />

generation intervals in domestic animals, specific effects of<br />

cloning on longevity and senescence have not yet been<br />

fully assessed. Preliminary data indicate no pathology in<br />

serial cloning of mice and cattle [119, 120].<br />

It has to be kept in mind that the clones frequently<br />

differ in a number of characteristics, because the two<br />

cytoplasms involved contain different maternal proteins,<br />

RNAs and mitochondrial DNAs. In addition, several<br />

epigenetic and environmental factors contribute to differences<br />

amongst groups of clones. Despite current limitations,<br />

somatic cloning has already emerged as an<br />

essential novel tool in basic biological research and holds<br />

great application perspectives, including the transgenic<br />

animal production.<br />

Genome Sequencing in Livestock Species<br />

Livestock genomics has followed the path that the human<br />

genome sequencing project has set, exploiting both<br />

strategy and technology. At the start of the Human<br />

Genome (HUGO) project, it was calculated that<br />

sequencing of one DNA base pair (bp) would cost around<br />

$1, thus sequence identification of the haploid human<br />

genome with approximately 3 10 9 bp was calculated to<br />

cost US $3 billion. Over the past decade, impressive<br />

improvements in sequencing techniques and automatization<br />

have substantially reduced these costs. Today,<br />

sequencing of one mammalian genome would cost about<br />

US $100 000 and this price might be reduced even further<br />

through a better technology [121]. The sequencing and<br />

annotation of the human genome revealed a dramatically<br />

lower number of protein-coding genes than originally<br />

thought ( 20 000 versus 150 000 genes) [122], which<br />

corresponds well with estimates for other vertebrate<br />

species [123, 124]. The bovine genome is currently at<br />

the 4.0 read and 90% of the genes have been allocated<br />

to chromosomes [125] (Table 2). The chicken genomic<br />

map first became available in 2004 and is accessible<br />

from the International Chicken Genome Sequencing<br />

RNA<br />

genes<br />

Chr., autosomal chromosome number and sex chromosomes; PC genes, number of identified protein coding genes; Pseudogenes, number<br />

of annotated pseudogenes; RNA genes, number of annotated RNA genes (RNA genes are not translated and act as structural or functional<br />

transcripts). Data retrieved from Ensembl.<br />

http://www.cababstractsplus.org/cabreviews<br />

Consortium [126]. The dog genome followed in 2005<br />

[127]. A draft sequence of the horse genome has<br />

been released recently (www.broad.mit.edu.mammals.<br />

horse). Open access to comprehensive genome information<br />

is provided by Ensembl (www.ensembl.org), University<br />

of California Santa Cruz (http://genome.ucsc.edu)<br />

or the National Center for Biotechnology Information<br />

(www.ncbi.org).<br />

The Ensembl project provides genome information for<br />

47 species, including complete and pre-release low coverage<br />

of genomes of several livestock, other mammalian,<br />

vertebrate and non-vertebrate species. These genome<br />

data are valuable resources for the study of comparative<br />

genomics. Ensembl includes also polymorphism data, such<br />

as SNPs (single nucleotide polymorphisms) for 10 species.<br />

This database will be complemented by new biological<br />

data resources such as protein–DNA interaction (ChiP–<br />

chip) maps originating from the combination of genomewide<br />

chromatin immunoprecipitation (ChiP) with DNA<br />

microarray (chip) technology. The Ensembl browser<br />

allows visualization of genome data in an individualized<br />

manner. Figure 2 provides an example and shows a<br />

100 kilobase (kb) region of bovine chromosome 2<br />

including the locus of the myostatin (GDF8) gene. Loss of<br />

function mutations of this gene results in muscle overgrowth<br />

as observed in several beef cattle breeds [128–<br />

130]. In Texel sheep, a gene mutation of the GDF8 gene<br />

produces a false microRNA target site, resulting in<br />

immediate transcript degradation and muscular hypertrophy<br />

[131]. In humans, a splice-site mutation of the<br />

GDF8 gene was found to cause pathological muscular<br />

hypertrophy [132]. This illustrates the great potential of<br />

comparative genomics using the Ensembl database. Direct<br />

access to genomic data and cell clones allows us to<br />

compare genomic regions of interest within populations<br />

or to design genetic modifications with a so far unprecedented<br />

speed and accuracy.<br />

For complex trait dissection, it is equally important to<br />

characterize the genetic variation within a population.<br />

Microsatellite markers and SNPs are the most commonly<br />

used polymorphisms for this purpose and have already<br />

revealed a surprisingly high degree of genetic variability in<br />

cattle [125]. SNP chips are rapidly <strong>bei</strong>ng developed for<br />

farm animals. Affymetrix already offers a bovine 10 K SNP<br />

panel and Illuma has developed SNP chips for poultry, pigs<br />

and cattle with 7–25 K SNPs.


Figure 2 A 100 kb region of bovine chromosome 2 extracted from Ensembl. The region contains the gene locus of the<br />

myostatin gene (GDF8) encoding a major growth factor regulating muscle growth. Sequence information can be downloaded,<br />

and additional information such as GC ratio, CpG islands, SNP data and repeat elements can be assessed<br />

Genomic data can be used for improving selective<br />

breeding. Traditional methods of selection are based<br />

on selecting for quantitative traits using breeding value<br />

estimates calculated from measured traits, genetic covariances<br />

among traits and the pedigree relationships in<br />

populations. In contrast, selection using genomic information<br />

supplements phenotypic data with knowledge of<br />

the individual genotypes in a population for known relevant<br />

gene variants (=alleles). This process is significantly<br />

quicker and at the same time can easily be applied to traits<br />

with low heritability and slow genetic change. The prerequisite<br />

for genomic selection is knowledge of the trait<br />

data, the QTL (Quantitative Trait Locus) or the gene.<br />

Current SNP maps are not yet dense enough to allow<br />

full exploitation of the potential in effective breeding<br />

programmes. Another important use of genomic data<br />

is to study expression patterns (transcriptomics or proteomics)<br />

and the production of transgenic animals tailored<br />

for specific production niches or areas.<br />

Gene Expression Profiling of<br />

Preimplantation Embryos<br />

http://www.cababstractsplus.org/cabreviews<br />

Wilfried A. Kues, Detlef Rath and Heiner Niemann 7<br />

Complex organisms exploit a variety of regulatory<br />

mechanisms to extend the functionality of their genomes<br />

[133], including differential promoter activation, alternative<br />

RNA splicing, RNA modification, RNA editing,<br />

localization, translation and stability of RNA, expression<br />

of non-coding RNA, antisense RNA and microRNAs


8 Perspectives in Agriculture, Veterinary Science, Nutrition and Natural Resources<br />

Figure 3 Transcriptome complexity and coverage of DNA expression arrays. Besides protein-coding transcripts, an<br />

unexpected large number of non-translated transcripts are expressed from the mammalian genome. Non-coding RNAs<br />

(ncRNA) include functional long ncRNAs such as XIST, AIR, CTN and PINK, and structural transcripts such as telomerase<br />

RNA (TERC), transfer RNA (tRNA) and ribosomal RNAs (rRNA). Small RNAs mainly act as regulatory units and include<br />

snoRNAs, microRNAs, siRNAs and piRNAs. The number of ncRNAs and small RNAs encoded within the genome is<br />

unknown. Recent transcriptomic studies suggest the presence of over 20 000 long ncRNAs and at least as many small<br />

regulatory RNAs<br />

[134–138]. These mechanisms mostly act at the RNA<br />

level and all together form the functional transcriptome<br />

of an organism (Figure 3). Tissue-specific transcriptomic<br />

profiles indicate important features of the regulatory<br />

process.<br />

During the past decade, efficient high-throughput<br />

methods have been developed for whole genome sequencing,<br />

transcriptome and proteome analyses [121,<br />

139, 140]. The rapid improvements of DNA microarraytechnologies<br />

and the availability of multi-species sequence<br />

databases, such as the Gene Ontology database (www.<br />

geneontology.org) and the Kyoto Encyclopaedia of Genes<br />

and Genomes (KEGG) (www.genome.jp/kegg), have made<br />

it possible to compare gene expression profiles both<br />

within a single organism and across a wide variety of<br />

organisms on a large scale. These technologies are crucial<br />

to improve the efficiency and safety of oocyte-/embryorelated<br />

biotechniques, to facilitate the development of<br />

novel therapies based on regenerative medicine and to<br />

ensure the quality of biotechnologically derived products<br />

at the molecular level.<br />

In vitro culture of preimplantation embryos is frequently<br />

associated with significant alterations in chromatin configuration<br />

and gene expression [141–143]. In vitro production<br />

of bovine embryos is usually necessary to provide<br />

the large numbers of embryos needed for an array analysis<br />

http://www.cababstractsplus.org/cabreviews<br />

in this species, but it is clear that the changes in embryonic<br />

transcription induced by the culture method itself<br />

seriously affect the transcriptional profile. The application<br />

of array technology to the analysis of preimplantation<br />

embryos poses specific challenges associated with the<br />

picogram levels of mRNA in a single embryo, the plasticity<br />

of the embryonic transcriptome and the difficulties in<br />

obtaining in vivo developing embryos. Gene expression<br />

patterns in mammalian embryos are more complex than<br />

those in most somatic cells because of the dramatic<br />

restructuring of paternal chromatin, the major onset of<br />

embryonic transcription, events related to the cellular<br />

differentiation at morula stage, blastocyst hatching and<br />

implantation. Embryo development is first dependent on<br />

maternally stored transcripts, which are gradually depleted<br />

until the embryo produces its own transcripts after<br />

embryonic genome activation [144]. The gene expression<br />

patterns and RNA stability in oocytes and early embryos<br />

prior to the activation of embryonic expression are<br />

dramatically different from what is seen after the major<br />

onset of embryonic transcription [145, 146]. The onset of<br />

embryonic gene transcription occurs at a species-specific<br />

time point; in mice, the major activation starts in late 1cell<br />

embryos; in humans and pigs, it occurs in 4-cell stages<br />

and in bovine embryos, is delayed until the 8–16-cell stage<br />

[147]. Given the late onset of this maternal-to-embryonic


transition in bovine embryos, they provide an excellent<br />

model in which to study early reprogramming events in<br />

detail. Major epigenetic reorganization also occurs during<br />

preimplantation development. This includes DNA demethylation<br />

and methylation as well as targeted modifications<br />

of histone methylation and acetylation, all of which<br />

play a role in controlling chromatin remodelling and<br />

X-chromosome inactivation [148].<br />

Transcriptome profiling by DNA array technology has<br />

been reported for human, mouse, bovine and porcine<br />

preimplantation embryos [145, 146, 147–158, reviewed in<br />

159]. A study of the transcriptional profile of 2-cell mouse<br />

embryos included 4 000 genes from a cDNA-library and<br />

revealed many previously unknown transcripts, including<br />

mobile genetic elements that appear to be associated<br />

with activation of the embryonic genome [160]. The<br />

mRNA expression profile of porcine oocytes, 4- and<br />

8-cell stages and blastocysts was studied with the aid of a<br />

15 K cDNA array specific for porcine reproductive<br />

organs. The transcription profile of in vivo developing<br />

embryos was compared with that of in vitro produced<br />

4-cell, 8-cell and blastocyst-stage embryos. This revealed<br />

1409 and 1696 differentially expressed (in vivo/in vitro)<br />

genes at the 4- and 8-cell stages, respectively [155]. Studies<br />

of the bovine embryonic mRNA expression based on<br />

subtractive hybridization and cDNA arrays have identified<br />

differentially expressed genes between bovine oocytes<br />

derived from large or small follicles. A total of 21 upregulated<br />

genes were found out of 1000 clones [161].<br />

Cross-species hybridization of a human cDNA array<br />

with bovine RNA, derived from rather large pools of<br />

>200 oocytes yielded a first expression profile for bovine<br />

oocytes before and after maturation. Approximately,<br />

300 genes could positively be identified [152, 156, 162].<br />

Recently, we have investigated the mRNA expression<br />

profiles of bovine oocytes and blastocysts by using a<br />

cross-species hybridization approach employing an array<br />

consisting of 15 529 human cDNAs (Ensembl chip) as<br />

probe, thus allowing the identification of conserved genes<br />

during human and bovine preimplantation developments<br />

[157]. The analysis revealed 419 genes that were expressed<br />

in both oocytes and blastocysts. The expression<br />

of 1342 genes was detected exclusively in the blastocyst,<br />

in contrast to 164 in the oocyte, including a significant<br />

number of novel genes. Typically, genes indicative for<br />

transcriptional and translational control, such as ELAVL4<br />

and TACC3 were overexpressed in the oocyte, whereas<br />

genes indicative of numerous pathways were overexpressed<br />

in blastocysts. Transcripts implicated in chromatin<br />

remodelling were found in both oocytes (NASP<br />

SMARCA2, DNMT1) and blastocysts (H2AFY, HDAC7A).<br />

Expression patterns in bovine and human blastocysts<br />

were identical to a large extent [157]. A major advantage<br />

of this cross-species hybridization approach is that<br />

developmentally conserved novel and annotated marker<br />

genes could be identified for human and bovine oocytes<br />

and blastocysts.<br />

http://www.cababstractsplus.org/cabreviews<br />

Wilfried A. Kues, Detlef Rath and Heiner Niemann 9<br />

Affymetrix commercially distributes a bovine genomic<br />

array which is based on the bovine genome sequence data<br />

and covers probe sets of approximately 23 000 transcripts<br />

and 19 000 UniGene clusters. The transcriptomes of<br />

bovine in vitro matured oocytes and in vitro-produced<br />

8-cell stages were analysed by this Affymetrix bovine<br />

array and genes critically involved in major activation of<br />

the embryonic genome were determined [163, 164].<br />

Approximately, 370 genes with functions in chromatin<br />

modification, apoptosis, signal transduction, metabolism<br />

and immune response were differentially expressed<br />

between oocytes and 8-cell stages, including DNMT1,<br />

DNMT2 and MeCP2, which were specifically up-regulated<br />

in 8-cell stages. Recently, we have hybridized RNA from<br />

in vivo-produced bovine oocytes and all stages of bovine<br />

preimplantation development till the blastocyst stage on<br />

the bovine Affymetrix array. For the first time, this analysis<br />

will yield a complete picture of mRNA expression in<br />

the processes of oocyte maturation, fertilization,<br />

embryonic activation and cell lineage differentiation in a<br />

large mammalian species (manuscript in preparation).<br />

Transgenic Animals in Agriculture<br />

Proof-of-principle that introduced recombinant DNA can<br />

result in phenotypic changes in mammals was shown in<br />

1982 by the production of transgenic mice expressing a<br />

rat growth hormone [165]. The first transgenic livestock<br />

species were produced by microinjection of DNA into<br />

a pronucleus of fertilized eggs [14]. Despite numerous<br />

efforts, the production of transgenic farm animals is still<br />

hampered by low efficiencies and high cost. Transgenic<br />

farm animal production has been extensively reviewed by<br />

us in the past few years [12, 95, 166]. The first recombinant<br />

pharmaceutical protein (human antithrombin III,<br />

ATryn 1 ) isolated from the milk of transgenic goats was<br />

approved in June 2006 by the European Medicine Agency<br />

(EMEA), clearly showing the great potential of this<br />

transgenic approach. Whereas the biomedical applications<br />

are quite advanced, agricultural application areas are lagging<br />

behind. Table 3 gives an overview on the various<br />

attempts to produce animals transgenic for agricultural<br />

traits. Prominent examples are: environmentally friendly<br />

pigs, in which phosphorus excretion is reduced by transgenic<br />

expression of the phytase gene [7], transgenic pigs<br />

with an altered pattern of polyunsaturated fatty acids in<br />

skeletal muscle, thus providing healthier pork [10, 169],<br />

cattle with resistance against Staphylococcus aureus mastitis<br />

by transgenic expression of lysostaphin restricted to the<br />

mammary gland [180], or cattle with a knockout of the<br />

prion gene thus rendering cattle non-susceptible to<br />

bovine spongiform encephalopathy (BSE) [11], alterations<br />

of the physicochemical properties of milk by overexpression<br />

of beta- and kappa-casein clearly underpinning<br />

the potential for improvements in the functional properties<br />

of bovine milk [178].


10 Perspectives in Agriculture, Veterinary Science, Nutrition and Natural Resources<br />

Table 3 Overview on successful transgenic livestock for agricultural production<br />

Transgenic trait Key molecule Construct<br />

Gene transfer<br />

method Species Reference<br />

Increased growth rate, less<br />

body fat<br />

Growth hormone (GH) hMT-pGH Microinjection Pig [167]<br />

Increased growth rate, less Insulin-like growth mMT-hIGF-1 Microinjection Pig [168]<br />

body fat<br />

factor-1 (IGF-1)<br />

Increased level of polyunsaturated<br />

fatty acids in pork<br />

Desaturase (from<br />

spinach)<br />

maP2-FAD2 Microinjection Pig [169]<br />

Increased level of poly-<br />

Desaturase (from CAGGS-hfat-1 Somatic Pig [10]<br />

unsaturated fatty acids in pork Caenorhabditis<br />

elegans)<br />

cloning<br />

Phosphate metabolism Phytase PSP-APPA Microinjection Pig [7]<br />

Milk composition (lactose a-lactalbumin Genomic bovine Microinjection Pig [170]<br />

increase)<br />

a-lactalbumin<br />

Influenza resistance Mx protein mMx1-Mx Microinjection Pig [171]<br />

Enhanced disease resistance IgA a,K-a,K Microinjection Pig,<br />

sheep<br />

[172, 175]<br />

Wool growth Insulin-like growth<br />

factor-1 (IGF-1)<br />

Ker-IGF-1 Microinjection Sheep [173, 174]<br />

Visna virus resistance Visna virus envelope Visna LTR-env Microinjection Sheep [175]<br />

Ovine prion locus Prion protein (PRNP) Targeting vector Somatic Sheep [176]<br />

(homologous<br />

recombination)<br />

cloning<br />

1<br />

Milk fat composition Stearoyl desaturase b-lactoglobulin-SCD Microinjection Goat [177]<br />

Milk composition (increase of b-casein Genomic CSN2 Somatic Cattle [178]<br />

whey proteins)<br />

k-casein CSN-CSN-3<br />

cloning<br />

Milk composition (increase of<br />

lactoferrin)<br />

Human lactoferrin a-s2cas-mLF Microinjection Cattle [179]<br />

Staphylococcus aureus<br />

Lysostaphin Ovine b-lactoglobulin Somatic Cattle [180]<br />

mastitis resistance<br />

lysostaphin<br />

cloning<br />

1 Animals dead shortly after birth (from [12]).<br />

With further refinements of the genetic maps many<br />

more agricultural applications will be developed. Further<br />

qualitative improvements may be derived from technologies<br />

that allow precise modifications of the genome,<br />

including targeted chromosomal integration by sitespecific<br />

DNA recombinases such as Cre or flippase (FLP),<br />

or methods that allow temporally and/or spatially controlled<br />

transgene expression [181].<br />

However, it remains to be shown whether transgenic<br />

farm animals will ever be used to improve agricultural<br />

applications, considering the public anxiety against gene<br />

modification and the extraordinary high costs to establish<br />

transgenic lines. Also our current understanding of<br />

genome organization is fragmentary and needs to be<br />

augmented. Improved techniques for transgenesis are<br />

required for a more efficient production of transgenic<br />

animals and to achieve targeted expression. Successful<br />

experiments to improve transgenesis by lentiviral vectors<br />

[182, 183], to achieve conditional expression [184] and to<br />

exploit large genomic sequences [185] are important<br />

steps in this direction. A powerful tool is RNA interference,<br />

which allows the suppression of specific mRNAs<br />

[186, 187]. The potential of this endogenous regulatory<br />

mechanism is highlighted by a naturally occurring mutation<br />

in Texel sheep, which results in an illegitimate target site<br />

http://www.cababstractsplus.org/cabreviews<br />

for endogenous microRNA in the myostatin (GDF8)<br />

transcript, resulting in rapid degradation of the mRNA and<br />

muscular hypertrophy [131]. Specific degradation of<br />

myostatin transcripts can be mimicked by short interfering<br />

RNAs (siRNAs), illustrating the potential of RNAi for<br />

breeding purposes. For a transient gene knockdown,<br />

synthetic siRNAs are transfected in cells or early embryos<br />

[188, 189]. For stable gene repression, the siRNA<br />

sequences must be incorporated into a gene construct.<br />

RNAi knockdown of porcine endogenous retroviruses<br />

(PERV) has been shown in porcine primary cells and<br />

cloned piglets [190, 191], as well as knockdown of the<br />

prion protein gene (PRNP) in cattle embryos [192].<br />

Germline transmission of lentiviral siRNA has been shown<br />

in rats over three generations [193]. In contrast to<br />

knockout technologies, which require time-consuming<br />

breeding strategies to obtain homozygous knockout lines,<br />

RNAi can rapidly be integrated into existing breeding<br />

lines.<br />

The first transgenic animals to be commercially<br />

marketed are ornamental fishes expressing recombinant<br />

fluorescent proteins, which are sold for their fanciness.<br />

Transgenic medaka (Oryzias latipes, rice fish) with a<br />

fluorescent green colour [194] were approved for sale in<br />

Taiwan. Transgenic zebrafishes (Danio rerio) with several


fluorescent colours [195] are sold under the brand name<br />

Glofish 1 in the USA [196]. The fluorescent fishes might<br />

act as ‘door-opener’ for gene modifications in other<br />

species.<br />

Recombinant DNA as New Vaccines<br />

Recombinant DNA techniques and the emerging genomic<br />

sequence information will be useful for novel applications,<br />

such as the development of a new class of vaccines.<br />

For DNA vaccination the transient presence of the coding<br />

sequence of a pathogen is sufficient to stimulate an<br />

immune response of the treated organism. The ongoing<br />

biodiversity genome-sequencing project (e.g. www.<br />

ncbi.org) facilitates access to sequence data of pathogenic<br />

organisms, and combined with current DNA synthesis<br />

methods DNA vaccines can be produced within a short<br />

period of time. The first DNA vaccine was approved<br />

by the US Department of Agriculture (USDA) in 2005<br />

(www.aphis.usda.gov/lpa/news/2005/07/wnvdna_vs.html);<br />

the vaccine is designed to protect horses against West<br />

Nile Virus. The development of DNA vaccines is based on<br />

the finding that injection of plasmid DNA into skeletal<br />

muscle is compatible with the production of ectopic<br />

proteins [197]. The main advantages of DNA vaccines are:<br />

DNA plasmids can be easily and at low costs prepared<br />

in compliance with GMP standards, DNA plasmids<br />

are extremely stable in lyophilized form, design of the<br />

plasmids is flexible, unlike attenuated pathogens DNA<br />

plasmids bear no risk of pathogen formation, and DNAvaccinated<br />

animals can be discriminated from their infected<br />

counterparts [198, 199].<br />

The principle of a DNA vaccine is that a gene sequence<br />

encoding for one or more protein epitopes of the<br />

pathogen is subcloned behind a strong eukaryotic promoter<br />

in a plasmid vector. The plasmid is amplified in a<br />

bacterial mass culture and then highly purified by ion<br />

exchange columns. The plasmid is then injected into<br />

muscle tissue, taken up by some muscle fibres and<br />

translocated into myonuclei. The eukaryotic promoter<br />

drives expression of pathogenic DNA sequences, resulting<br />

in the presentation of pathogen epitopes on the<br />

cell surface followed by triggering both a humoral and a<br />

cellular immune response of the vaccinated organism.<br />

Plasmid DNA was found to be degraded within a few<br />

weeks after intramuscular injection in pigs [200]. Since<br />

only partial sequences of the pathogen’s genome are used,<br />

no functional pathogen can be produced and vaccinated<br />

animals be discriminated from infected animals, which<br />

show antibody formation against all epitopes.<br />

Outlook<br />

Development in animal genomics have broadly followed<br />

the path that the human genome field has led, both in<br />

http://www.cababstractsplus.org/cabreviews<br />

Wilfried A. Kues, Detlef Rath and Heiner Niemann 11<br />

Figure 4 A perspective for molecular breeding based on<br />

genomic data and ES cells<br />

terms of completeness of data and in developing tools<br />

and applications as demonstrated by the recent availability<br />

of advanced drafts of the maps of the bovine, chicken,<br />

dog and horse genomes. The possibility of extending<br />

knowledge through deliberately engineering change in the<br />

genome is unique to animals and is currently <strong>bei</strong>ng<br />

developed through the integration of advanced molecular<br />

tools and breeding technologies. However, the full realization<br />

of this exciting potential is handicapped by gaps in<br />

our understanding of embryo genomics and epigenetic<br />

mechanisms which are critical in the production of healthy<br />

offspring.<br />

The convergence of the recent advances in reproductive<br />

technologies with the tools of molecular biology<br />

opens a new dimension for animal breeding. Major prerequisites<br />

are the continuous refinement of reproductive<br />

biotechnologies and a rapid completion and refinement of<br />

livestock genomic maps. A critical role in this context is<br />

played by embryonic stem cells (ES cells). The availability<br />

of cells with true pluripotent properties in conjunction<br />

with exploitation of genomic data will allow molecular<br />

breeding programmes that are significantly faster and<br />

more effective than current programmes (Figure 4).<br />

Despite numerous efforts, no ES cells with germ line<br />

contribution have been established from mammals<br />

other than mouse. ES-like cells, i.e. cells that share several<br />

parameters with true ES cells, have been reported in<br />

several species [201]. The recent finding that the transgenic<br />

expression of four transcription factors (Oct4, Sox2,<br />

Klf4 and c-Myc) convert murine fibroblasts into pluripotent<br />

ES cells [202] will stimulate similar approaches in<br />

livestock (Figure 4). Genetically modified animals will<br />

play a growing role in the biomedical field. Agricultural<br />

application might be further down the track given<br />

the complexity of some of the economically important<br />

traits.<br />

The efficient use of domestic animals is urgently needed<br />

in light of the worldwide shortage of arable land and the<br />

ever increasing human population. Genomics offers great<br />

opportunities to improve livestock production in the<br />

developed world and advance the ‘livestock revolution’ in


12 Perspectives in Agriculture, Veterinary Science, Nutrition and Natural Resources<br />

the developing world [203]. In addition to the already<br />

exploited biomedical potential, biotechnology in farm<br />

animals with a strong link to genomic technologies can<br />

make a significant contribution towards a more efficient,<br />

highly diversified and targeted, and thus sustainable animal<br />

production. The challenge for scientists and practitioners<br />

is to exploit the great potential of animal biotechnology<br />

for the benefits of the mankind.<br />

References<br />

1. Larson G, Albarella U, Dobney K, Rowley-Conwy P, Schibler<br />

J, Tresset A, et al. Ancient DNA, pig domestication, and the<br />

spread of the Neolithic into Europe. Proceedings of the<br />

National Academy of Sciences of the United States of<br />

America 2007;104:15276–81.<br />

2. Beja-Pereira A, Caramelli D, Lalueza-Fox C, Vernesi C,<br />

Ferrand N, Casoli A, et al. The origin of European cattle:<br />

evidence from modern and ancient DNA. Proceedings of the<br />

National Academy of Sciences of the United States of<br />

America 2006;103:8113–8.<br />

3. Fries R, Ruvinsky A (editors). The Genetics of Cattle. CAB<br />

International, Wallingford, UK; 1999.<br />

4. Weitze KF. Update of the world wide application of swine AI.<br />

In: Johnson LA, Guthrie HD, editors. Boar Semen<br />

Preservation IV. Allen Press Inc, Lawrence, KS, USA;<br />

2000, p. 141–5.<br />

5. Seidel Jr GE. Sexing mammalian spermatozoa and embryos<br />

– state of the art. Journal of Reproduction and Fertility<br />

1999;54 Suppl.:477–87.<br />

6. Thibier M. New records in the numbers of both in vivo-derived<br />

and in vitro-produced bovine embryos around the world in<br />

2006. IETS-Newsletter 2007;25:15–20.<br />

7. Golovan SP, Meidinger RG, Ajakaiye A, Cottrill M,<br />

Wiederkehr MZ, Barney DJ, et al. Pigs expressing salivary<br />

phytase produce low-phosphorus manure. Nature<br />

Biotechnology 2001;19:741–5.<br />

8. Dai Y, Vaught TD, Boone J, Chen SH, Phelps CJ,<br />

Ball S, et al. Targeted disruption of the<br />

alpha1,3-galactosyltransferase gene in cloned pigs.<br />

Nature Biotechnology 2002;20(3):251–5.<br />

9. Lai L, Kolber-Simonds D, Park KW, Cheong HT, Greenstein<br />

JL, Im GS, et al. Production of alpha-1,3galactosyltransferase<br />

knockout pigs by nuclear transfer<br />

cloning. Science 2002;295(5557):1089–92.<br />

10. Lai L, Kang JX, Li R, Wang J, Witt WT, Yong HY, et al.<br />

Generation of cloned transgenic pigs rich in omega-3 fatty<br />

acids. Nature Biotechnology 2006;24:435–6.<br />

11. Richt J, Kasinathan P, Hamir AN, Castilla J, Sathiyaseelan T,<br />

Vargas F, et al. Production of cattle lacking prion protein.<br />

Nature Biotechnology 2007;25:132–8.<br />

12. Niemann H, Kues WA. Transgenic farm animals: an update.<br />

Reproduction Fertility and Development 2007;19:762–70.<br />

13. Wilmut I, Schnieke AE, McWhir J, Kind AJ, Campbell KH.<br />

Viable offspring derived from fetal and adult mammalian cells.<br />

Nature 1997;385:810–3.<br />

http://www.cababstractsplus.org/cabreviews<br />

14. Hammer RE, Pursel VG, Rexroad Jr CE, Wall RJ, Bolt DJ,<br />

Ebert KM, et al. Production of transgenic rabbits, sheep and<br />

pigs by microinjection. Nature 1985;315:680–3.<br />

15. Koopman P, Gubbay J, Vivian N, Goodfellow P, Lovell-Badge<br />

R. Male development of chromosomally female mice<br />

transgenic for Sry. Nature 1991;351:117–21.<br />

16. Johnson LA. Gender preselection in domestic animals using<br />

flow cytometrically sorted sperm. Journal of Animal Science<br />

1992;70 Suppl. 2 :8–18.<br />

17. Johnson LA, Flook JP, Hawk HW. Sex preselection in rabbits:<br />

live births from X and Y sperm separated by DNA and cell<br />

sorting. Biology of Reproduction 1989;41:199–203.<br />

18. Johnson LA, Flook JP, Look MV. Flow cytometry of X and<br />

Y chromosome-bearing sperm for DNA using an improved<br />

preparation method and staining with Hoechst 33342.<br />

Gamete Research 1987;17:203–12.<br />

19. Johnson LA, Welch GR. Sex preselection: high-speed flow<br />

cytometric sorting of X- and Y-sperm for maximum efficiency.<br />

Theriogenology 1999;52:1134–323.<br />

20. Johnson LA. Sex preselection in swine: altered sex ratios in<br />

offspring following surgical insemination of flow sorted X- and<br />

Y-bearing sperm. Reproduction in Domestic Animals<br />

1991;26:309–14.<br />

21. Rath D, Long CR, Dobrinsky JR, Welch GR, Schreier LL,<br />

Johnson LA. In vitro production of sexed embryos for gender<br />

preselection: high-speed sorting of X-chromosome-bearing<br />

sperm to produce pigs after embryo transfer. Journal of<br />

Animal Science 1999;77:3346–52.<br />

22. Johnson LA, Flook JP, Look MV, Pinkel D. Flow sorting of<br />

X and Y chromosome-bearing spermatozoa into two<br />

populations. Gamete Research 1987;16:1–9.<br />

23. Welch GR, Johnson LA. Sex preselection: laboratory<br />

validation of the sperm sex ratio of flow sorted X- and<br />

Y-sperm by sort reanalysis for DNA. Theriogenology<br />

1999;52:1343–52.<br />

24. Kawarasaki T, Welch GR, Long CR, Yoshida M, Johnson LA.<br />

Verification of flow cytometorically sorted X- and Y-bearing<br />

porcine spermatozoa and reanalysis of spermatozoa for DNA<br />

content using the fluorescence in situ hybridization (FISH)<br />

technique. Theriogenology 1998;50:625–35.<br />

25. Rens W, Yang F, Welch G, Revell S, O’Brien PCM, Solanky<br />

N, et al. An X-Y paint set and sperm FISH protocol that can<br />

be used for validation of cattle sperm separation procedures.<br />

Reproduction 2001;121:541–6.<br />

26. Welch GR, Waldbieser GC, Wall RJ, Johnson LA. Flow<br />

cytometric sperm sorting and PCR to confirm separation of<br />

X- and Y-chromosome bearing bovine sperm. Animal<br />

Biotechechnology 1995;6:131–9.<br />

27. Seidel Jr GE, Allen CH, Johnson LA, Holland MD, Brink Z,<br />

Welch GR, et al. Uterine horn insemination of heifers with<br />

very low numbers of non-frozen and sexed spermatozoa.<br />

Theriogenology 1997;48:1255–64.<br />

28. Den Daas JH, De Jong G, Lansbergen LM, Van Wagtendonk-<br />

De Leeuw AM. The relationship between the number of<br />

spermatozoa inseminated and the reproductive efficiency of<br />

individual dairy bulls. Journal of Dairy Science<br />

1998;81(6):1714–23.<br />

29. Klinc P, Rath D. Reduction of oxidative stress in bovine<br />

spermatozoa during flow cytometric sorting. Reproduction in<br />

Domestic Animals 2007;42:63–7.


30. Rath D, Ruiz S, Sieg B. Birth of female piglets following<br />

intrauterine insemination of a sow using flow cytometrically<br />

sexed boar semen [abstract]. Veterinary Record<br />

2003;152:400–1.<br />

31. McNutt TL, Johnson LA. Flow cytometric sorting of sperm:<br />

influence on fertilization and embryo/fetal development in the<br />

rabbit. Molecular Reproduction and Development<br />

1996;43:261–7.<br />

32. Cran DG, Cochrane DJ, Johnson LA, Wei H, Lu KH, Polge C.<br />

Separation of X- and Y-chromosome bearing bovine sperm<br />

by flow cytometry for use in IVF [abstract]. Theriogenology<br />

1994;41:183.<br />

33. Lu KH, Cran DG, Seidel GE. In vitro fertilization with flowcytometrically-sorted<br />

bovine sperm. Theriogenology<br />

1999;52:1393–405.<br />

34. Merton JS, Haring RM, Stap J, Hoebe RA, Aten JA. Effect of<br />

flow cytometrically sorted frozen/thawed semen on success<br />

rate of in vitro bovine embryo production [abstract].<br />

Theriogenology 1997;47:295.<br />

35. Beyhan Z, Johnson LA, First NL. Sexual dimorphism in<br />

IVM-IVF bovine embryos produced from X and Y<br />

chromosome-bearing spermatozoa sorted by high speed flow<br />

cytometry. Theriogenology 1999;52:35–48.<br />

36. Johnson LA, Cran DG, Welch GR, Polge C. Gender<br />

preselection in mammals. In: Miller RH, Pursel VG, Norman<br />

HD, editors. Beltsville Symposium XX: Biotechnology’s Role<br />

in the Genetic Improvement of Farm Animals. American<br />

Society of Animal Science, Savoy, IL; 1996. p. 151–4.<br />

37. Schenk JL, Seidel Jr GE. Pregnancy rates in cattle with<br />

cryopreserved sexed spermatozoa: effects of laser intensity,<br />

staining conditions and catalase. Society of Reproduction and<br />

Fertility 2007;Suppl. 64:165–77.<br />

38. Catt SL, Sakkas D, Bizzaro D, Bianchi PG, Maxwell WMC,<br />

Evans G. Hoechst staining and exposure to UV laser during<br />

flow cytometric sorting does not affect the frequency of<br />

detected endogenous DNA nicks in abnormal and normal<br />

human sperm. Molecular Human Reproduction 1997;<br />

3:821–5.<br />

39. Parrilla I, Vazquez JM, Cuello C, Gil MA, Roca J,<br />

Di Berardino D, et al. Hoechst 33342 stain and u.v. laser<br />

exposure do not induce genotoxic effects in flow-sorted boar<br />

spermatozoa. Reproduction 2004;128:615–21.<br />

40. Campos-Chillon LF, de la Torre JF. Effect on concentration of<br />

sexed bovine sperm sorted at 40 and 50 psi on<br />

developmental capacity of in vitro produced embryos<br />

[abstract]. Theriogenology 2003;59:506.<br />

41. Suh TK, Schenk JL, Seidel Jr GE. High pressure flow<br />

cytometric sorting damages sperm Theriogenology<br />

2005;64:1035–48.<br />

42. Morton KM, Herrmann D, Sieg B, Struckmann C, Maxwell<br />

WM, Rath D. Altered mRNA expression patterns in bovine<br />

blastocysts after fertilisation in vitro using flow-cytometrically<br />

sex-sorted sperm. Molecular Reproduction and Development<br />

2007;74:931–40.<br />

43. Bermejo-Alvarez P, Rizos D, Rath D, Lonergan P, Gutierrez-<br />

Adan A. Epigenetic differences between male and female<br />

bovine blastocysts produced in vitro. Physiological Genomics<br />

2008;32:264–72.<br />

44. Cran DG, Johnson LA, Miller NGA, Cochrane D, Polge CE.<br />

Production of bovine calves following separation of X- and<br />

http://www.cababstractsplus.org/cabreviews<br />

Wilfried A. Kues, Detlef Rath and Heiner Niemann 13<br />

Y-bearing sperm and in vitro fertilization. Veterinary Record<br />

1993;132:40–1.<br />

45. Morton KM, Catt SL, Maxwell WM, Evans G. An efficient<br />

method of ovarian stimulation and in vitro embryo production<br />

from prepubertal lambs. Reproduction Fertility and<br />

Development 2005;17:701–6.<br />

46. Klinc P, Frese D, Osmers H, Rath D. Insemination with sex<br />

sorted fresh bovine spermatozoa processed in the presence<br />

of antioxidative substances. Reproduction in Domestic<br />

Animals 2007;42:58–62.<br />

47. Klinc P. Improved fertility of flowcytometrically sex selected<br />

bull spermatozoa [dissertation]. Veterinary University,<br />

Hannover, Germany; 2005.<br />

48. Garcia EM, Vazquez JM, Parrilla I, Calvete JJ, Sanz L,<br />

Caballero I, et al. Improving the fertilizing ability of sex sorted<br />

boar spermatozoa. Theriogenology 2007;68:771–8.<br />

49. Centurion F, Vazquez JM, Calvete JJ, Roca J, Sanz L,<br />

Parrilla I, et al. Influence of porcine spermadhesins on the<br />

susceptibility of boar spermatozoa to high dilution. Biology<br />

of Reproduction 2003;69:640–6.<br />

50. Maxwell WM, Parrilla I, Caballero I, Garcia E, Roca J,<br />

Martinez EA, et al. Retained functional integrity of bull<br />

spermatozoa after double freezing and thawing using<br />

PureSperm density gradient centrifugation. Reproduction in<br />

Domestic Animals 2007;42:489–94.<br />

51. Grossfeld R, Klinc P, Sieg B, Rath D. Production of piglets<br />

with sexed semen employing a non-surgical insemination<br />

technique. Theriogenology 2005;63:2269–77.<br />

52. Bathgate R, Grossfeld R, Susetio D, Ruckholdt M, Heasman<br />

K, Rath D, et al. Early pregnancy loss in sows after low dose,<br />

deep uterine artificial insemination with sex-sorted, frozenthawed<br />

sperm. Animal Reproduction Science 2008;<br />

104(3–4):440–4.<br />

53. Cuello C, Berthelot F, Martinat-Botte´ F, Venturi E, Guillouet<br />

P, Va´zquez JM, et al. Piglets born after non-surgical deep<br />

intrauterine transfer of vitrified blastocysts in gilts. Animal<br />

Reproduction Science 2005;85:275–86.<br />

54. Clulow JR, Buss H, Sieme H, Rodger JA, Cawdell-Smith AJ,<br />

Evans G, et al. Field fertility of sex-sorted and non-sorted<br />

frozen-thawed stallion spermatozoa. Animal Reproduction<br />

Science 2008 [Epub ahead of print doi:10.1016/<br />

j.anireprosci.2007.08.015].<br />

55. Heer P. Adjustment of cryoperservation of stallion<br />

spermatozoa to the Beltsville Sperm Sexing Technology<br />

[dissertation]. Veterinary University, Hannover, Germany;<br />

2007.<br />

56. Rath D, Johnson LA, Welch GR. In vitro culture of porcine<br />

embryos: development to blastocysts after in vitro fertilization<br />

(IVF) with flow cytometrically sorted and unsorted semen<br />

[abstract]. Theriogenology 1993;39:293.<br />

57. Rath D, Johnson LA, Dobrinsky JR, Welch GR, Niemann H.<br />

Production of piglets preselected for sex following in vitro<br />

fertilization with X and Y chromosome-bearing spermatozoa<br />

sorted by flow cytometry. Theriogenology 1997;47:795–800.<br />

58. Abeydeera LR, Johnson LA, Welch GR, Wang WH, Boquest<br />

AC, Cantley TC. Birth of piglets preselected for gender<br />

following in vitro fertilization of in vitro matured pig oocytes by<br />

X and Y chromosome bearing spermatozoa sorted by high<br />

speed flow cytometry. Theriogenology 1998;50:981–8.


14 Perspectives in Agriculture, Veterinary Science, Nutrition and Natural Resources<br />

59. Probst S, Rath D. Production of piglets using intracytoplasmic<br />

sperm injection (ICSI) with flow cytometrically sorted boar<br />

semen and artificially activated oocytes. Theriogenology<br />

2003;59:961–73.<br />

60. Willadsen SM. Nuclear transplantation in sheep embryos.<br />

Nature 1986;320:63–5.<br />

61. Westhusin ME, Pryor JH, Bondioli KR. Nuclear transfer in the<br />

bovine embryo: a comparison of 5-day, 6-day, frozen-thawed,<br />

and nuclear transfer donor embryos. Molecular Reproduction<br />

and Development 1991;28:119–23.<br />

62. Briggs R, King TJ. Transplantation of living nuclei from<br />

blastula cells into enucleated frogs eggs. Proceedings of the<br />

National Academy of Sciences of the United States of<br />

America 1952;38:455–463.<br />

63. Miyoshi K, Rzucidlo SJ, Pratt SL, Stice SL. Improvements in<br />

cloning efficiencies may be possible by increasing uniformity<br />

in recipient oocytes and donor cells (review; published<br />

erratum appears in Biology of Reproduction 2004;70:260).<br />

Biology of Reproduction 2003;68:1079–86.<br />

64. Brem G, Ku¨hholzer B. The recent history of somatic cloning in<br />

mammals (review). Cloning and Stem Cells 2002;4:57–63.<br />

65. Hochedlinger K, Jaenisch R. On the cloning of animals from<br />

terminally differentiated cells. Nature Genetics 2007;<br />

39:136–7.<br />

66. Eggan K, Baldwin K, Tackett M, Osborne J, Gogos J,<br />

Chess A, et al. Mice cloned from olfactory sensory neurons.<br />

Nature 2004;428:44–9.<br />

67. Zhu H, Craig JA, Dyce PW, Sunnen N, Li J. Embryos derived<br />

from porcine skin-derived stem cells exhibit enhanced<br />

preimplantation development. Biology of Reproduction<br />

2004;71:1890–7.<br />

68. Hornen N, Kues WA, Carnwath JW, Lucas-Hahn A, Petersen<br />

B, Hassel P, et al. Production of viable pigs from fetal somatic<br />

stem cells. Cloning and Stem Cells 2007;9:364–73.<br />

69. Campbell KH, McWhir J, Ritchie WA, Wilmut I. Sheep cloned<br />

by nuclear transfer from a cultured cell line. Nature<br />

1996;380:64–6.<br />

70. Kues WA, Anger M, Carnwath JW, Paul D, Motlik J,<br />

Niemann H. Cell cycle synchronization of porcine fibroblasts:<br />

effects of serum deprivation and reversible cell cycle<br />

inhibitors. Biology of Reproduction 2000;62:412–9.<br />

71. Kues WA, Carnwath JW, Paul D, Niemann H. Cell cycle<br />

synchronization of porcine fetal fibroblasts by serum<br />

deprivation initiates a nonconventional form of apoptosis.<br />

Cloning and Stem Cells 2002;3:231–43.<br />

72. Cibelli JB, Stice SL, Golueke PJ, Kane JJ, Jerry J,<br />

Blackwell C, et al. Cloned transgenic calves produced from<br />

nonquiescent fetal fibroblasts. Science<br />

1998;280(5367):1256–8.<br />

73. Wakayama T, Rodriguez I, Perry AC, Yanagimachi R,<br />

Mombaerts P. Mice cloned from embryonic stem cells.<br />

Proceedings of the National Academy of Sciences of the<br />

United States of America 1999;96:14984–9.<br />

74. Rizos D, Ward F, Duffy P, Boland MP, Lonergan P.<br />

Consequences of bovine oocyte maturation, fertilization or<br />

early embryo development in vitro vs. in vivo: implications for<br />

blastocysts yields. Molecular Reproduction and Development<br />

2002;61:234–48.<br />

75. Niemann H, Wrenzycki C, Lucas-Hahn A, Brambrink T,<br />

Kues WA, Carnwath JW. Gene expression patterns in bovine<br />

http://www.cababstractsplus.org/cabreviews<br />

in vitro-produced and nuclear transfer-derived embryos and<br />

their implications for early development. Cloning and Stem<br />

Cells 2002;4:29–38.<br />

76. Wakayama T, Perry AC, Zuccotti M, Johnson KR,<br />

Yanagimachi R. Full-term development of mice from<br />

enucleated oocytes injected with cumulus cell nuclei. Nature<br />

1998;394(6691):369–74.<br />

77. Baguisi A, Behboodi E, Melican DT, Pollock JS, Destrempes<br />

MM, Cammuso C, et al. Production of goats by somatic cell<br />

nuclear transfer. Nature Biotechnology 1999;17(5):456–61.<br />

78. Betthauser J, Forsberg E, Augenstein M, Childs L, Eilertsen<br />

K, Enos J, et al. Production of cloned pigs from in vitro<br />

systems. Nature Biotechnology 2000;18(10):1055–9.<br />

79. Onishi A, Iwamoto M, Akita T, Mikawa S, Takeda K, Awata T,<br />

et al. Pig cloning by microinjection of fetal fibroblast nuclei.<br />

Science 2000;289(5482):1188–90.<br />

80. Polejaeva IA, Chen SH, Vaught TD, Page RL, Mullins J, Ball<br />

S, et al. Cloned pigs produced by nuclear transfer from adult<br />

somatic cells. Nature 2000;407(6800):86–90.<br />

81. Chesne´ P, Adenot PG, Viglietta C, Baratte M, Boulanger L,<br />

Renard JP. Cloned rabbits produced by nuclear transfer from<br />

adult somatic cells. Nature Biotechnology 2002;20(4):366–9.<br />

82. Shin T, Kraemer D, Pryor J, Liu L, Rugila J, Howe L,<br />

et al. A cat cloned by nuclear transplantation. Nature<br />

2002;415:859.<br />

83. Galli C, Lagutina I, Crotti G, Colleoni S, Turini P, Ponderato<br />

N, et al. Pregnancy: a cloned horse born to its dam twin.<br />

Nature 2003;424:635.<br />

84. Woods GL, White KL, Vanderwall DK, Li GP, Aston KI, Bunch<br />

TD, et al. A mule cloned from fetal cells by nuclear transfer.<br />

Science 2003;301:1063.<br />

85. Zhou Q, Renard JP, Le Friec G, Brochard V, Beaujean N,<br />

Cherifi Y, et al. Generation of fertile cloned rats by regulating<br />

oocyte activation. Science 2003;302:1179.<br />

86. Lee BC, Kim MK, Jang G. Dogs cloned from adult somatic<br />

cells. Nature 2005;436:641.<br />

87. Li Z, Sun X, Chen J, Liu X, Wisely SM, Zhou Q, et al. Cloned<br />

ferrets produced by somatic cell nuclear transfer.<br />

Developmental Biology 2006;293:439–48.<br />

88. Kim MK, Jang G, Oh HJ, Yuda F, Kim HJ, Hwang WS, et al.<br />

Endangered wolves cloned from adult somatic cells. Cloning<br />

Stem Cells 2007;9(1):130–7.<br />

89. Wells D, Misica P, Tervit H, Reed W. Adult somatic cell<br />

nuclear transfer is used to preserve the last surviving cow of<br />

the Enderby island cattle breed. Reproduction Fertility and<br />

Development 1999;10:369–78.<br />

90. White K, Bunch T, Mitalipov S, Reed W. Establishment of<br />

pregnancy after the transfer of nuclear transfer embryos<br />

produced from the fusion of argali (Ovis ammon) nuclei into<br />

domestic sheep (Ovis aries) enucleated oocytes. Cloning<br />

1999;1:47–54.<br />

91. Lanza R, Cibelli J, Diaz F, Moraes C, Farin C, Hammer C,<br />

et al. Cloning of an endangered species (Bos taurus) using<br />

interspecies nuclear transfer. Cloning 2000;2:79–90.<br />

92. Loi P, Ptak G, Barboni B, Fulka JJ, Cappai M, Clinton M.<br />

Genetic rescue of an endangered mammal by cross-species<br />

nuclear transfer using post mortem somatic cell. Nature<br />

Biotechnology 2001;19:962–4.


93. Ryder O. Cloning advances and challenges for conservation.<br />

Trends in Biotechnology 2002;20:231–2.<br />

94. Groeneveld E, Tinh NH, Kues W, Vien NT. A protocol for<br />

the cryoconservation of breeds by low-cost emergency cell<br />

banks – a pilot study. Animal 2008;2:1–8.<br />

95. Kues WA, Niemann H. The contribution of farm animals to<br />

human health. Trends in Biotechnology 2004;22:286–94.<br />

96. Petersen B, Lucas-Hahn A, Lemme E, Hornen N, Hassel P,<br />

Niemann H. Preovulatory embryo transfer increases success<br />

of porcine somatic cloning [abstract]. Reproduction Fertility<br />

and Development 2007;19:155–6.<br />

97. Renard JP, Chastant S, Chesne P, Richard C, Marchal J,<br />

Cordonnier N, et al. Lymphoid hypoplasia and somatic<br />

cloning. Lancet 1999;353:1489–91.<br />

98. Tamashiro KL, Wakayama T, Blanchard RJ, Blanchard DC,<br />

Yanagimachi R. Postnatal growth and behavioral<br />

development of mice cloned from adult cumulus cells. Biology<br />

of Reproduction 2000;63:328–34.<br />

99. Ogonuki N, Inoue K, Yamamoto Y, Noguchi Y, Tanemura K,<br />

Suzuki O, et al. Early death of mice cloned from somatic cells.<br />

Nature Genetics 2002;30:253–4.<br />

100. Perry AC, Wakayama T. Untimely ends and new beginnings<br />

in mouse cloning. Nature Genetics 2002;30:243–4.<br />

101. Rhind SM, King TJ, Harkness LM, Bellamy C, Wallace W,<br />

DeSousa P, et al. Cloned lambs-lessons from pathology.<br />

Nature Biotechnology 2003;21:744–5.<br />

102. Farin P, Piedrahita JA, Farin CE. Errors in development of<br />

fetuses and placentas from in vitro-produced bovine<br />

embryos. Theriogenology 2006;65:178–91.<br />

103. Cibelli JB, Campbell KH, Seidel GE, West MD, Lanza RP.<br />

The health profile of cloned animals. Nature Biotechnology<br />

2002;20:13–41.<br />

104. Panarace M, Agu¨ero JI, Garrote M, Jauregui G, Segovia A,<br />

Cane´ L, et al. How healthy are clones and their progeny: 5<br />

years of field experience. Theriogenology 2007;67:142–51.<br />

105. Humpherys D, Eggan K, Akutsu H, Hochedlinger K, Rideout<br />

III WM, Biniszkiewicz D, et al. Epigenetic instability in ES<br />

cells and cloned mice. Science 2001;293:95–7.<br />

106. Lanza RP, Cibelli JB, Faber D, Sweeney RW, Henderson B,<br />

Nevala W, et al. Cloned cattle can be healthy and normal.<br />

Science 2001;294:1893–4.<br />

107. Chavatte-Palmer P, Heyman Y, Richard C, Monget P,<br />

LeBourhis D, Kann G, et al. Clinical, hormonal, and<br />

hematologic characteristics of bovine calves derived from<br />

nuclei from somatic cells. Biology of Reproduction<br />

2002;66:1596–603.<br />

108. Govoni KE, Tian XC, Kazmer GW, Taneja M, Enright BP,<br />

Rivard AL, et al. Age-related changes of the somatotropic<br />

axis in cloned Holstein calves. Biology of Reproduction<br />

2002;66:1293–8.<br />

109. Enright BP, Taneja M, Schreiber D, Riesen J, Tian XC,<br />

Fortune JE, et al. Reproductive characteristics of cloned<br />

heifers derived from adult somatic cells. Biology of<br />

Reproduction 2002;66:291–6.<br />

110. Pace MM, Augenstein ML, Betthauser JM, Childs LA,<br />

Eilertsen KJ, Enos JM, et al. Ontogeny of cloned cattle to<br />

lactation. Biology of Reproduction 2002;67:334–9.<br />

111. Tian XC, Kubota C, Sakashita K, Izaike Y, Okano R, Tabara<br />

N, et al. Meat and milk compositions of bovine clones.<br />

http://www.cababstractsplus.org/cabreviews<br />

Wilfried A. Kues, Detlef Rath and Heiner Niemann 15<br />

Proceedings of the National Academy of Sciences of the<br />

United States of America 2005;102:6261–6.<br />

112. Yang X, Tian C, Kubota C, Page R, Xu J, Cibelli J, et al. Risk<br />

assessment of meat and milk from cloned animals. Nature<br />

Biotechnology 2007;25:77–83.<br />

113. Miller HI. Food from cloned animals is part of our brave old<br />

world. Trends in Biotechnology 2007;25:201–3.<br />

114. Archer GS, Dindot S, Friend TH, Walker S, Zaunbrecher G,<br />

Lawhorn B, et al. Hierarchical phenotypic and epigenetic<br />

variation in cloned swine. Biology of Reproduction<br />

2003;69:430–6.<br />

115. Committee on Defining Science-Based Concerns Associated<br />

with Products of Animal Biotechnology, Committee on<br />

Agricultural Biotechnology, Health, and the Environment,<br />

National Research Council. Animal Biotechnology: Science<br />

based concerns. Washington, DC, USA: National Academy<br />

Press; 2002.<br />

116. Kumugai S. Safety of animal foods that utilize cloning<br />

technology. Report to the Japanese Ministry for Agriculture,<br />

Forestry and Fishery (MAFF) 2002.<br />

117. Rudenko L, Matheson JC, Sundlof SF. Animal Cloning and<br />

the FDA: the risk assessment paradigm under public scrutiny.<br />

Nature Biotechnology 2007;25:39–43.<br />

118. European Food Safety Agency 2007. Available from: URL:<br />

http://www.efsa.europa.eu/EFSA/efsa_locale-<br />

1178620753812_1178676923092.htm<br />

119. Wakayama T, Shinkai Y, Tamashiro KL, Niida H,<br />

Blanchard DC, Blanchard RJ. Cloning of mice to six<br />

generations. Nature 2000;407:318–9.<br />

120. Kubota C, Tian XC, Yang X. Serial bull cloning by somatic cell<br />

nuclear transfer. Nature Biotechnology 2004;22:693–4.<br />

121. Bennett ST, Barnes C, Cox A, Davies L, Brown C. Toward<br />

the 1,000 dollars human genome. Pharmacogenomics<br />

2005;6:373–82.<br />

122. Goodstadt L, Ponting CP. Phylogenetic reconstruction of<br />

orthology, paralogy, and conserved synteny for dog and<br />

human. PLoS Computational Biology 2006;2:e133.<br />

123. Aparicio S, Chapman J, Stupka E, Putnam N, Chia JM, Dehal<br />

P, et al. Whole-genome shotgun assembly and analysis of<br />

the genome of Fugu rubripes. Science 2002;297:1301–10.<br />

124. Gregory SG, Sekhon M, Schein J, Zhao S, Osoegawa K,<br />

Scott CE, et al. A physical map of the mouse genome. Nature<br />

2002;418:743–50.<br />

125. Adelson DL. Insights and applications from sequencing the<br />

bovine genome. Reproduction Fertility and Development<br />

2008;20:54–60.<br />

126. Fadiel A, Anidi I, Eichenbaum KD. Farm animal genomics<br />

and informatics: an update. Nucleic Acids Research<br />

2005;33:6308–18.<br />

127. Lindblad-Toh K, Wade CM, Mikkelsen TS, Karlsson EK, Jaffe<br />

DB, Kamal M, et al. Genome sequence, comparative analysis<br />

and haplotype of the domestic dog. Nature 2005;438:803–19.<br />

128. Grobet L, Martin LJ, Poncelet D, Pirottin D, Brouwers B,<br />

Riquet J, et al. A deletion in the bovine myostatin gene<br />

causes the double-muscled phenotype in cattle. Nature<br />

Genetics 1997;17:71–4.<br />

129. Kambadur R, Sharma M, Smith TP, Bass JJ. Mutations in<br />

myostatin (GDF8) in double-muscled Belgian Blue and<br />

Piedmontese cattle. Genome Research 1997;7:910–6.


16 Perspectives in Agriculture, Veterinary Science, Nutrition and Natural Resources<br />

130. McPherron AC, Lee SJ. Double muscling in cattle due to<br />

mutations in the myostatin gene. Proceedings of the National<br />

Academy of Sciences of the United States of America<br />

1997;94:12457–61.<br />

131. Clop A, Marcq F, Takeda H, Pirottin D, Tordoir X, Bibe´ B.<br />

A mutation creating a potential illegitimate microRNA target<br />

site in the myostatin gene affects muscularity in sheep.<br />

Nature Genetics 2006;38:813–8.<br />

132. Schuelke M, Wagner KR, Stolz LE, Hu¨bner C, Riebel T,<br />

Ko¨men W, et al. Myostatin mutation associated with gross<br />

muscle hypertrophy in a child. New England Journal of<br />

Medicine 2004;350:2682–8.<br />

133. Hayashizaki Y, Kanamori M. Dynamic transcriptome of mice<br />

[review]. Trends in Biotechnology 2004;22:161–7.<br />

134. Bie´mont C, Vieira C. Genetics: junk DNA as an evolutionary<br />

force. Nature 2006;443:521–4.<br />

135. Rassoulzadegan M, Grandjean V, Gounon P, Vincent S,<br />

Gillot I, Cuzin F. RNA-mediated non-mendelian inheritance<br />

of an epigenetic change in the mouse. Nature 2006;<br />

441:469–74.<br />

136. Segal E, Fondufe-Mittendorf Y, Chen L, Thastrom A, Field Y,<br />

Moore IK, et al. A genomic code for nucleosome positioning.<br />

Nature 2006;442:772–8.<br />

137. Werner A. Natural antisense transcripts [review]. RNA<br />

Biology 2005;2:53–62.<br />

138. Jirtle RL, Skinner MK. Environmental epigenomics and<br />

disease susceptibility. Nature Reviews Genetics 2007;<br />

8:253–62.<br />

139. Schena M, Shalon D, Davis RW, Brown PO. Quantitative<br />

monitoring of gene expression patterns with a<br />

complementary DNA microarray. Science 1995;270:467–70.<br />

140. Ragoussis J, Elvidge GP, Kaur K, Colella S. Matrix-assisted<br />

laser desorption/ionisation, time-of-flight mass spectrometry<br />

in genomics research. PLoS Genetics 2006;2:e100.<br />

141. Niemann H, Wrenzycki C. Alterations of expression of<br />

developmentally important genes in preimplantation bovine<br />

embryos by in vitro culture conditions: implications for<br />

subsequent development [review]. Theriogenology<br />

2000;53:21–34.<br />

142. Wrenzycki C, Herrmann D, Lucas-Hahn A, Gebert C,<br />

Korsawe K, Lemme E, et al. Epigenetic reprogramming<br />

throughout preimplantation development and consequences<br />

for assisted reproductive technologies [review]. Birth Defects<br />

Research Part C-Embryo Today 2005;75:1–9.<br />

143. Wrenzycki C, Herrmann D, Lucas-Hahn A, Korsawe K,<br />

Lemme E, Niemann H. Messenger RNA expression patterns<br />

in bovine embryos derived from in vitro procedures and their<br />

implications for development [review]. Reproduction Fertility<br />

and Development 2005;17:23–35.<br />

144. Schultz RM. Regulation of zygotic gene activation in the<br />

mouse. Bioessays 1993;15(8):531–8.<br />

145. Wang QT, Piotrowska K, Ciemerych MA, Milenkovic L, Scott<br />

MP, Davis RW, et al. A genome-wide study of gene activity<br />

reveals developmental signaling pathways in the<br />

preimplantation mouse embryo. Developmental Cell<br />

2004;6(1):133–44.<br />

146. Zeng F, Baldwin DA, Schultz RM. Transcript profiling during<br />

preimplantation mouse development. Developmental Biology<br />

2004;272:483–96.<br />

http://www.cababstractsplus.org/cabreviews<br />

147. Telford NA, Watson AJ, Schultz GA. Transition from maternal<br />

to embryonic control in early development: A comparison of<br />

several species. Molecular Reproduction and Development<br />

1990;26:90–100.<br />

148. Kiefer JC. Epigenetics in development [review].<br />

Developmental Dynamics 2007;236:1144–56.<br />

149. Bermudez MG, Wells D, Malter H, Munne S, Chen J,<br />

Steuerwald, NM. Expression profiles of individual human<br />

oocytes using microarray technology. Reprod Biomed Online<br />

2004;8:325–337.<br />

150. Dobson AT, Raja R, Abeyta MJ, Taylor T, Shen S, Haqq C,<br />

et al. The unique transcriptome through day 3 of human<br />

preimplantation development. Human Molecular Genetics<br />

2004;13(14):1461–70.<br />

151. Hamatani T, Daikoku T, Wang H, Matsumoto H, Carter MG,<br />

Ko MS, et al. Global gene expression analysis identifies<br />

molecular pathways distinguishing blastocyst dormancy and<br />

activation. Proceedings of the National Academy of Sciences<br />

of the United States of America 2004;101(28):10326–31.<br />

152. Dalbies-Tran R, Mermillod P. Use of heterologous<br />

complementary DNA array screening to analyze bovine<br />

oocyte transcriptome and its evolution during in vitro<br />

maturation. Biology of Reproduction 2003;68:252–61.<br />

153. Valle´e M, Gravel C, Palin MF, Reghenas H, Stothard P,<br />

Wishart DS, et al. Identification of novel and known oocytespecific<br />

genes using complementary DNA subtraction and<br />

microarray analysis in three different species. Biology of<br />

Reproduction 2005;73(1):63–71.<br />

154. Mamo S, Ponsuksili S, Wimmers K, Gilles M, Schellander K.<br />

Expression of retinoid X receptor transcripts and their<br />

significance for developmental competence in<br />

in vitro-produced pre-implantation-stage bovine embryos.<br />

Reproduction in Domestic Animals 2005;40(2):177–83.<br />

155. Whitworth KM, Agca C, Kim JG, Patel RV, Springer GK,<br />

Bivens NJ, et al. Transcriptional profiling of pig<br />

embryogenesis by using a 15-K member unigene set specific<br />

for pig reproductive tissues and embryos. Biology of<br />

Reproduction 2005;72:1437–51.<br />

156. Valle´e M, Robert C, Me´thot S, Palin MF, Sirard MA. Crossspecies<br />

hybridizations on a multi-species cDNA microarray to<br />

identify evolutionarily conserved genes expressed in oocytes.<br />

BMC Genomics 2006;7:113.<br />

157. Adjaye J, Herwig R, Brink T, Herrmann D, Greber B, Groth D,<br />

et al. Conserved molecular profiles of maternal and<br />

embryonic gene expression during bovine and human<br />

preimplantation development. Physiological Genomics<br />

2007;35:315–27.<br />

158. Patel OV, Bettegowda A, Ireland JJ, Coussens PM, Lonergan<br />

P, Smith GW. Functional genomics studies of oocyte<br />

competence: evidence that reduced transcript abundance for<br />

follistatin is associated with poor developmental competence<br />

of bovine oocytes. Reproduction 2007;133(1):95–106.<br />

159. Niemann H, Carnwath JW, Kues WA. Application of DNA<br />

array technology to mammalian embryos. Theriogenology<br />

2007;68 Suppl. 1:165–77.<br />

160. Evsikov AV, Graber JH, Brockman JM, Hampl A, Holbrook<br />

AE, Singh P, et al. Cracking the egg: molecular dynamics and<br />

evolutionary aspects of the transition from the fully grown<br />

oocyte to embryo. Genes and Development 2006;<br />

20:2713–27.


161. Donnison M, Pfeffer PL. Isolation of genes associated with<br />

developmentally competent bovine oocytes and quantitation<br />

of their levels during development. Biology of Reproduction<br />

2004;71(6):1813–21.<br />

162. Robert C, Barnes F, Hue I, Sirard MA. Subtractive<br />

hybridization used to identify mRNA associated with the<br />

maturation of bovine oocytes. Molecular Reproduction and<br />

Development 2000;57:167–75.<br />

163. Misirlioglu M, Page GP, Sagirkaya H, Parrish JJ, First NL,<br />

Memili E. Dynamics of global transcriptome in bovine<br />

matured oocytes and preimplantation embryos. Proceedings<br />

of the National Academy of Sciences of the United States of<br />

America 2006;103:18905–10.<br />

164. Fair T, Carter F, Park S, Evans AO, Lonergan P. Global gene<br />

expression analysis during bovine oocyte in vitro maturation.<br />

Theriogenology 2007;68:91–7.<br />

165. Palmiter RD, Brinster RL, Hammer RE, Trumbauer ME,<br />

Rosenfeld MG, Birnberg NC. Dramatic growth of mice that<br />

develop from eggs microinjected with metallothionein-growth<br />

hormone fusion genes. Nature 1982;300:611–5.<br />

166. Niemann H, Kues WA, Carnwath JW. Transgenic farm<br />

animals: present and past [review]. Revue Scientifique et<br />

Technique (Office International des Epizooties)<br />

2005;24:285–98.<br />

167. Nottle MB, Nagashima H, Verma PJ, Du ZT, Grupen CG,<br />

Mellfatrick SM, et al. Production and analysis of transgenic<br />

pigs containing a metallothionein porcine growth hormone<br />

gene construct. In: Murray JD, Oberbauer AM, McGloughlin<br />

MM, editors. Transgenic Animals in Agriculture. CABI<br />

Publishing, New York, USA; 1999. p. 145–56.<br />

168. Pursel VG, Wall RJ, Mitchell AD, Elsasser TH, Solomon MB,<br />

Coleman ME, et al. Expression of insulin-like growth factor-1<br />

in skeletal muscle of transgenic pigs. In: Murray JD,<br />

Anderson GB, Oberbauer AM, McGloughlin MM, editors.<br />

Transgenic Animals in Agriculture. CABI Publishing, New<br />

York, NY, USA; 1999. p. 131–44.<br />

169. Saeki K, Matsumoto K, Kinoshita M, Suzuki I, Tasaka Y,<br />

Kano K, et al. Functional expression of a Delta12 fatty acid<br />

desaturase gene from spinach in transgenic pigs.<br />

Proceedings of the National Academy of Sciences of the<br />

United States of America 2004;101:6361–6.<br />

170. Wheeler MB, Bleck GT, Donovan SM. Transgenic alteration<br />

of sow milk to improve piglet growth and health. Reproduction<br />

2001;58 Suppl.:313–24.<br />

171. Mu¨ller M, Brenig B, Winnacker EL, Brem G. Transgenic pigs<br />

carrying cDNA copies encoding the murine Mx1 protein which<br />

confers resistance to influenza virus infection. Gene<br />

1992;121:263–70.<br />

172. Lo D, Pursel V, Linton PJ, Sandgren E, Behringer R, Rexroad<br />

C, et al. Expression of mouse IgA transgenic mice, pigs and<br />

sheep. European Journal of Immunology 1991;21:1001–6.<br />

173. Damak S, Jay NP, Barrell GK, Bullock DW. Targeting gene<br />

expression to the wool follicle in transgenic sheep.<br />

Biotechnology 1996b;14:181–4.<br />

174. Damak S, Su H, Jay NP, Bullock DW. Improved wool<br />

production in transgenic sheep expressing insulin-like growth<br />

factor 1. Biotechnology 1996;14:185–8.<br />

175. Clements JE, Wall RJ, Narayan O, Hauer D, Schoborg R,<br />

Sheffer D, et al. Development of transgenic sheep that<br />

express the visna virus envelope gene. Virology<br />

1994;200:370–80.<br />

http://www.cababstractsplus.org/cabreviews<br />

Wilfried A. Kues, Detlef Rath and Heiner Niemann 17<br />

176. Denning C, Burl S, Ainslie A, Bracken J, Dinnyes A, Fletcher<br />

J, et al. Deletion of the alpha(1,3)galactosyl transferase<br />

(GGTA1) gene and the prion protein (PrP) gene in sheep.<br />

Nature Biotechnology 2001;19:559–62.<br />

177. Reh WA, Maga EA, Collette NM, Moyer A, Conrad-Brink JS,<br />

Taylor SJ, et al. Hot topic: using a stearoyl-CoA<br />

desaturase transgene to alter milk fatty acid composition.<br />

Journal of Dairy Science 2004;87:3510–4.<br />

178. Brophy B, Smolenski G, Wheeler T, Wells D, L’Huillier P,<br />

Laible G. Cloned transgenic cattle produce milk with higher<br />

levels of b-casein and k-casein. Nature Biotechnology<br />

2003;21:157–62.<br />

179. Platenburg GJ, Kootwijk EAP, Kooiman PM, Woloshuk SL,<br />

Nuijens JH, Krimpenfort JA, et al. Expression of human<br />

lactoferrin in milk of transgenic mice. Transgenic Research<br />

1994;3:99–108.<br />

180. Wall RJ, Powell A, Paape MJ, Kerr DE, Bannermann DD,<br />

Pursel VG, et al. Genetically enhanced cows resist<br />

intramammary Staphylococcus aureus infection. Nature<br />

Biotechnology 2005;23:445–51.<br />

181. Niemann H, Kues WA. Application of transgenesis on<br />

livestock for agriculture and biomedicine. Animal<br />

Reproduction Science 2003;79:291–317.<br />

182. Hofmann A, Kessler B, Ewerling S, Weppert M, Vogg B,<br />

Ludwig H, et al. Efficient transgenesis in farm animals by<br />

lentiviral vectors. EMBO Reports 2003;4:1054–60.<br />

183. Whitelaw CB, Radcliffe PA, Ritchie WA, Carlisle A, Ellard FM,<br />

Pena RN, et al. Efficient generation of transgenic pigs using<br />

equine infectious anaemia virus (EIAV) derived vector. FEBS<br />

Letters 2004;571:233–6.<br />

184. Kues WA, Schwinzer R, Wirth D, Verhoeyen E, Lemme E,<br />

Herrmann D, et al. Epigenetic silencing and tissue<br />

independent expression of a novel tetracycline inducible<br />

system in double-transgenic pigs. FASEB Journal Express<br />

2006;20:E1–E10, doi: 10.1096/fj.05-5415fje.<br />

185. Kuroiwa Y, Kasinathan P, Choi YJ, Naeem R, Tomizuka K,<br />

Sullivan EJ, et al. Cloned transchromosomic calves<br />

producing human immunoglobulin. Nature Biotechnology<br />

2002;20:889–94.<br />

186. Fire A, Xu S, Montgomery MK, Kostas SA, Driver SE, Mello<br />

CC. Potent and specific genetic interference by doublestranded<br />

RNA in Caenorhabditis elegans. Nature<br />

1998;391(6669):806–11.<br />

187. Dallas A, Vlassow A. RNAi: a novel antisense technology and<br />

its therapeutic potential. Medical Science Monitor<br />

2006;12:RA67–74.<br />

188. Clark J, Whitelaw B. A future for transgenic livestock. Nature<br />

Reviews Genetics 2003;4:825–33.<br />

189. Iqbal K, Kues WA, Niemann H. Parent-of-origin dependent<br />

gene-specific knock down in mouse embryos. Biochemical<br />

and Biophysical Research Communications 2007;<br />

358:727–32.<br />

190. Dieckhoff B, Karlas A, Hofmann A, Kues WA, Petersen B,<br />

Pfeifer A, Niemann H, et al. Inhibition of porcine endogenous<br />

retroviruses (PERV) in primary porcine cells by RNA<br />

interference using lentiviral vectors. Archives of Virology<br />

2007;152:629–34.<br />

191. Dieckhoff B, Petersen B, Kues WA, Kurth R, Niemann H,<br />

Denner J. Knockdown of porcine endogenous retrovirus


18 Perspectives in Agriculture, Veterinary Science, Nutrition and Natural Resources<br />

(PERV) expression by PERV-specific shRNA in transgenic<br />

pigs. Xenotransplantation 2008;15:36–45.<br />

192. Golding MC, Long CR, Carmell MA, Hannon GJ, Westhusin<br />

ME. Suppression of prion protein in livestock by RNA<br />

interference. Proceedings of the National Academy of<br />

Sciences of the United States of America 2006;103:5285–90.<br />

193. Tenenhaus-Dann C, Alvarado AL, Hammer RE, Garbers DL.<br />

Heritable and stable gene knockdown in rats. Proceedings<br />

of the National Academy of Sciences of the United States of<br />

America 2006;103:11246–51.<br />

194. Chou CY, Horng LS, Tsai HJ. Uniform GFP-expression in<br />

transgenic medaka (Oryzias latipes) at the F0 generation.<br />

Transgenic Research 2001;10:303–15.<br />

195. Gong W, Wan H, Tay TL, Wang H, Chen M, Yan T.<br />

Development of transgenic fish for ornamental and bioreactor<br />

by strong expression of fluorescent proteins in the skeletal<br />

muscle. Biochemical and Biophysical Research<br />

Communications 2003;308:58–63.<br />

196. Food and Drug Administration. USA (2003) FDA statement<br />

regarding Glofish. Available from: URL: http://www.fda.gov/<br />

bbs/topics/NEWS/2003/NEW00994.html.<br />

http://www.cababstractsplus.org/cabreviews<br />

197. Wolff JA, Malone RW, Williams P, Chong W, Acsadi G,<br />

Jani A, et al. Direct gene transfer into mouse muscle in vivo.<br />

Science 1990;247:1465–8.<br />

198. Davis HL. CpG motifs for optimization of DNA vaccines.<br />

Developmental Biology (Basel) 2001;104:165–9.<br />

199. Lorenzen N, LaPatra SE. DNA vaccines for aquacultured<br />

fish. Revue Scientifique et Technique (International Office of<br />

Epizootics) 2005;24:201–13.<br />

200. Gravier R, Dory D, Laurentie M, Bougeard S, Cariolet R,<br />

Jestin A. In vivo tissue distribution and kinetics of a<br />

pseudorabies virus plasmid DNA vaccine after intramuscular<br />

injection in swine. Vaccine 2007;25:6930–8.<br />

201. Gjorret JO, Maddox-Hyttel P. Attempts towards derivation<br />

and establishment of bovine embryonic stem cell-like<br />

cultures. Reproduction Fertility and Development<br />

2005;17:113–24.<br />

202. Takahashi K, Yamanaka S. Induction of pluripotent stem cells<br />

from mouse embryonic and adult fibroblast cultures by<br />

defined factors. Cell 2006;126:663–76.<br />

203. Rothschild MF, Plastow GS. Impact of genomics on animal<br />

agriculture and opportunities for human health. Trends in<br />

Biotechnology 2008;26(1):21–5.


Xenotransplantation 2008: 15: 36–45<br />

Printed in Singapore. All rights reserved<br />

doi: 10.1111/j.1399-3089.2008.00442.x<br />

Knockdown of porcine endogenous<br />

retrovirus (PERV) expression by<br />

PERV-specific shRNA in transgenic pigs<br />

Dieckhoff B, Petersen B, Kues WA., Kurth R, Niemann H, Denner J.<br />

Knockdown of porcine endogenous retrovirus (PERV) expression by<br />

PERV-specific shRNA in transgenic pigs.<br />

Xenotransplantation 2008; 15: 36–45. Ó 2008 Blackwell Munksgaard<br />

Abstract: Background: Xenotransplantation using porcine cells, tissues<br />

or organs may be associated with the transmission of porcine endogenous<br />

retroviruses (PERVs). More than 50 viral copies have been identified<br />

in the pig genome and three different subtypes of PERV were<br />

released from pig cells, two of them were able to infect human cells in<br />

vitro. RNA interference is a promising option to inhibit PERV transmission.<br />

Methods: We recently selected an efficient si (small interfering) RNA<br />

corresponding to a highly conserved region in the PERV DNA, which is<br />

able to inhibit expression of all PERV subtypes in PERV-infected<br />

human cells as well as in primary pig cells. Pig fibroblasts were transfected<br />

using a lentiviral vector expressing a corresponding sh (short<br />

hairpin) RNA and transgenic pigs were produced by somatic nuclear<br />

transfer cloning. Integration of the vector was proven by PCR, expression<br />

of shRNA and PERV was studied by in-solution hybridization<br />

analysis and real-time RT PCR, respectively.<br />

Results: All seven born piglets had integrated the transgene. Expression<br />

of the shRNA was found in all tissues investigated and PERV expression<br />

was significantly inhibited when compared with wild-type control<br />

animals.<br />

Conclusion: This strategy may lead to animals compatible with PERV<br />

safe xenotransplantation.<br />

Introduction<br />

Xenotransplantation using porcine cells, tissues or<br />

organs may offer a potential solution for the<br />

shortage of allogeneic human organs [1]. Pigs are<br />

the most favored donor species for several reasons:<br />

(i) similar physiology and size of organs, (ii)<br />

unlimited availability, (iii), short generation interval,<br />

high number of progeny (up to 15 to 18<br />

piglets); (iv) relatively low costs, (v) maintenance<br />

under high hygienic conditions is possible, i.e.,<br />

specified pathogen-free (spf) breeding and gnotobiotic<br />

delivery is possible, (vi) tools for the genetic<br />

modification to reduce the immunogenicity are<br />

available and (vii) somatic cloning is possible and<br />

thus production of transgenic animals can be<br />

36<br />

Copyright Ó 2008 Blackwell Munksgaard<br />

Britta Dieckhoff, 1 Bjçrn Petersen, 2<br />

Wilfried A. Kues, 2 Reinhard<br />

Kurth, 1 Heiner Niemann 2 and<br />

Joachim Denner 1<br />

1 Robert Koch Institute, Berlin, 2 Department of<br />

Biotechnology, Institute for Animal Science,<br />

Mariensee, Neustadt, Germany<br />

Key words: porcine endogenous retrovirus – RNA<br />

interference – shRNA – siRNA – transgenic pig –<br />

xenotransplantation<br />

Address reprint requests to Joachim Denner, PhD,<br />

Robert Koch Institute, Nordufer 20, 13353 Berlin,<br />

Germany (E-mail: dennerj@rki.de)<br />

Received 6 November 2007;<br />

Accepted 12 November 2007<br />

XENOTRANSPLANTATION<br />

significantly enhanced [2]. However, prior to the<br />

clinical use of porcine xenotransplants, three main<br />

hurdles have to be overcome: The immunological<br />

rejections, the physiological incompatibility and<br />

the risk of transmission of porcine pathogens. The<br />

premier immunological hurdle, the hyperacute<br />

rejection, which is due to the stimulation of the<br />

human complement system through the gal alpha<br />

1,3 gal epitopes on the surface of porcine cells can<br />

be overcome in a clinically acceptable manner by<br />

either the production of Gal knock-out animals<br />

[3,4] or animals over-expressing human complement<br />

regulatory factors [5].<br />

Specified pathogen-free breeding of pigs can prevent<br />

transmission of most porcine microbes. However,<br />

porcine endogenous retroviruses (PERVs) are


integrated in the porcine genome [6,7], and can be<br />

released as infectious particles from normal pig<br />

cells [8–10] and can infect human cells in vitro [11–<br />

15]. The three replication-competent subtypes of<br />

PERVs mainly differ in their receptor-binding site<br />

of the envelope protein [6,16]. The subtypes PERV-<br />

A and PERV-B are polytropic and able to infect<br />

human cells in vitro, whereas the ecotropic PERV-<br />

C infects only porcine cells.<br />

Retroviruses are tumorigenic and may cause<br />

immunodeficiencies in infected hosts. The risk of<br />

PERV infection has to be carefully assessed<br />

specifically in view of the fact that the human<br />

immunodeficiency viruses HIV-1 and HIV-2, causing<br />

AIDS in humans, are the result of multiple<br />

trans-species transmissions of retroviruses from<br />

non-human primates [17,18]. However, in several<br />

clinical applications of porcine islet cells or ex vivo<br />

perfusions of pig liver cells such as bridging during<br />

acute liver failure, no virus transmission could be<br />

detected [19–22]. Likewise, in a cohort of 160<br />

patients treated with various porcine tissues, no<br />

PERV transmission was observed [23,24]. Moreover,<br />

PERV transmission was not found after pig<br />

to non-human primate transplantation and PERV<br />

infection experiments [25–27]. However, the risk of<br />

PERV transmission associated with tumors or<br />

immunodeficiencies is still prevalent [28]. Removal<br />

of PERV sequences by knock-out technology<br />

would be the safest strategy, but may prove<br />

difficult due to the presence of numerous replication<br />

competent and defective proviruses capable of<br />

recombination and complementing each other.<br />

Different strategies for the prevention of PERV<br />

transmission have been developed including the<br />

selection of low-producer animals based on sensitive<br />

PERV assays [29–31], the generation of an<br />

antiviral vaccine [32], and the inhibition of PERV<br />

expression by RNA interference (RNAi) [33–35].<br />

In these first experiments efficient short interfering<br />

RNAs (siRNAs) were selected and inhibition of<br />

virus expression was studied in PERV infected<br />

human cells [33] and primary pig cells [36].<br />

The mechanism of RNAi is highly conserved<br />

among different biological systems, including<br />

plants, flies, worms, and mammals [37]. After<br />

processing by the ribonuclease III-like enzyme<br />

Dicer, double stranded RNA is cleaved into small<br />

double stranded fragments (21 to 25 nt) [38]. In<br />

the cytoplasm, these siRNAs interact with the<br />

RNA induced silencing complex, which recognizes<br />

the target mRNA, that is homologous to the<br />

sequence of the siRNA. Nucleases, which are part<br />

of the complex, degrade the target RNA and<br />

prevent protein translation [39,40]. Transfection<br />

of mammalian cells with longer dsRNA (>30 nt)<br />

induced an interferon response associated with an<br />

unspecific degradation of mRNA and finally<br />

apoptosis. The Dicer processing step can be<br />

bypassed by transfecting cells with synthetic<br />

siRNAs [41]. The amount of siRNAs within the<br />

cells is diluted out during cell division, thus<br />

limiting the time of gene silencing activity to<br />

approximately four to eight cell doublings [42,43].<br />

Expression of sh (short hairpin) RNAs, that<br />

contain a short loop sequence linking the forward<br />

and reverse strand, can overcome this limitation.<br />

Defined shRNA without any cap- or polyAstructure<br />

can be expressed by polymerase-IIIdependent<br />

promoters such as the murine or<br />

human U6 snRNA- or the human RNase P<br />

(H1) RNA-promoters [44–47]. In addition, cells<br />

with stable chromosomal integration and permanent<br />

inhibition of target genes can be obtained if<br />

adequate selection markers are used. In addition,<br />

retroviral vectors may be used to integrate the<br />

siRNA producing vector [36]. The efficacy of<br />

RNAi has been investigated in vivo, until now<br />

mainly in mice, rats, and cattles [48–52].<br />

Here, we show for the first time lentiviral vectormediated<br />

RNAi in pigs associated with a significant<br />

inhibition of PERV expression.<br />

Materials and methods<br />

Knockdown of PERV in pigs<br />

Cell culture and lentiviral transfection<br />

Primary pig fibroblasts were transduced with the<br />

lentiviral vector pLVTHM-pol2 (Fig. 1). The vector<br />

expressed a shRNA corresponding to the viral<br />

pol2 sequence able to inhibit PERV expression<br />

efficiently [36]. Transduced cells were grown as<br />

described [53], selected by FACS using the expression<br />

of the reporter gene GFP and used in nuclear<br />

transfer.<br />

Generation of transgenic pig embryos by somatic nuclear transfer<br />

Transgenic pigs were generated by somatic nuclear<br />

transfer cloning using in vitro matured abattoir<br />

oocytes [54]. Briefly, oocytes were enucleated by<br />

removing the first polar body along with adjacent<br />

cytoplasm containing the metaphase plate. The<br />

donor cells were arrested at G0/G1 of the cell cycle<br />

by contact inhibition and serum starvation<br />

[DMEM + 0.5% fetal calf serum (FCS)] for<br />

48 h. Prior to nuclear transfer, cells were treated<br />

for 10 min with trypsin/EDTA and centrifuged<br />

(200 g/3 min) twice in a 10 ml tube containing 5 ml<br />

phosphate-buffered saline and 100 ll FCS. The<br />

supernatant was removed and the cells were resuspended<br />

in Ca 2+ -free TL-HEPES medium. A small<br />

37


Dieckhoff et al.<br />

Fig. 1. Localization of the primers specific for GFP and pol2 in the lentiviral vector pLVTHM-pol2 (LTR long terminal repeat, H1<br />

polymerase III H1-RNA gene promotor, shRNA short hairpin RNA, SIN self-inactivating element, cPPT central polypurine tract,<br />

GFP green fluorescent protein, WPRE post-transcriptional element of woodchuck hepatitis virus, EF-1alfa elongation factor-1a<br />

promoter, tetO tetracycline operator, loxP locus of X-over of P1, recombination site of Cre recombinase for bacteriophage P1, the<br />

size of the amplicons is indicated).<br />

fibroblast from the positive cell clone was placed in<br />

the perivitelline space in close contact with the<br />

oocyte membrane to form a couplet. After cell<br />

transfer, fusion was induced in Ca 2+ -free SOR2<br />

medium [0.25 m Sorbitol, 0.5 mm Mg-acetate,<br />

0.1% bovine serum albumin (BSA)]. A single<br />

electrical pulse of 15 V for 100 ls (Eppendorf<br />

MultiporatorÒ, Eppendorf, Germany) between<br />

two electrodes was employed to fuse the membranes<br />

of the oocyte and the donor cell. The<br />

reconstructed embryos were activated in an electrical<br />

field of 1.0 kV/cm for 45 ls in SOR2 activation<br />

medium (0.25 m Sorbitol, 0.1 mm Ca-acetate,<br />

0.5 mm Mg-acetate, 0.1% BSA) followed by incubation<br />

with 2 mm 6-dimethylaminopurine (DMAP;<br />

Sigma, Steinheim, Germany) in NCSU23 medium<br />

for 3 h prior to transfer to recipients.<br />

Synchronization of recipients and embryo transfer<br />

Peripuberal German Landrace gilts served as<br />

recipients. The gilts were synchronized by feeding<br />

20 mg/pig Altrenogest (Regumate Ò ; Janssen-Cilag,<br />

Neuss, Germany) for 13 days followed by an<br />

injection of 1000 IU Pregnant Mare Serum Gonadotropine<br />

(PMSG; Intergonan Ò , Intervet, Unterschleissheim,<br />

Germany) on the last day of<br />

Altrenogest feeding. Ovulations were induced by<br />

intramuscular injections of 500 IU human Chorion<br />

Gonadotropin (hCG; Ovogest Ò , Intervet) 80 h<br />

after PMSG administration. Nineteen hours later,<br />

115 reconstructed embryos were surgically transferred<br />

into the oviducts of one recipient via mid<br />

ventral laparatomy under general anesthesia<br />

induced and maintained with thiopental sodium<br />

(Trapanal Ò , Altana, Konstanz, Germany). Maintenance<br />

of pregnancies was supported by administration<br />

of 1000 IU PMSG and 500 IU hCG on<br />

days 12 and 15 after surgery. Pregnancy was<br />

confirmed by ultrasound on days 25 and 35.<br />

38<br />

DNA isolation<br />

DNA was isolated from ear biopsies using a salt<br />

chloroform extraction method. Sixty micrograms<br />

tissue were incubated in 500 ll HOM buffer<br />

(160 mm saccharose, 80 mm EDTA, 100 mm<br />

Tris–HCl pH 8.0) (all reagents in this chapter<br />

from Roth, Karlsruhe, Germany) and 20 ll ofa<br />

proteinase K solution (20 mg/ml; Invitrogen, Karlsruhe,<br />

Germany) at 60 °C over night. After<br />

addition of 200 ll of a 4.5 m NaCl solution and<br />

700 ll chloroform/isoamyl alcohol, DNA was<br />

extracted by centrifugation (10 min, 10 000 g,<br />

4 °C). The aqueous phase containing genomic<br />

DNA was mixed with 700 ll isopropanol and<br />

centrifuged as described above. After washing with<br />

70% (v/v) ethanol the precipitated DNA was<br />

resuspended in 70 ll H 2O and stored at )20 °C<br />

until assayed.<br />

Detection of transgene<br />

Integration of the transgene was investigated by<br />

two independent PCR assays, using primers<br />

specific for (i) GFP (GFP for: 5¢-GATCACGAG-<br />

ACTAGCCTCGAGGT, GFP rev: 5¢-CCAGGA-<br />

TGTTGCCGTCCTC) and (ii) the shRNA<br />

expression cassette (pol2 for: 5¢-AACGCT-<br />

GACGTCATCAAC, pol2 rev: 5¢-GGACGCTG-<br />

ACAAATTGAC) under the following temperature<br />

conditions: 95 °C, 10 min; 35 cycles (95 °C, 30 s;<br />

45 °C and 54 °C respectively, 30 s; 72 °C, 1 min);<br />

72 °C, 5 min.<br />

Isolation of siRNA and total RNA<br />

The siRNA and siRNA-depleted total RNA isolation<br />

from frozen tissues was carried out using the<br />

mirVana miRNA isolation Kit (Ambion, Huntingdon,<br />

UK) allowing the enrichment of RNAs


smaller than 200 nucleotides according to the<br />

manufacturer’s instructions and stored at )80 °C<br />

until assayed.<br />

RNA probe construction<br />

For the in liquid hybridization a short 32 P (GE<br />

Healthcare, Freiburg, Germany)-labeled RNA<br />

probe (29 nt) was generated by in vitro transcription<br />

using the mirVana miRNA Probe Construction Kit<br />

(Ambion). A DNA oligonucleotide with the reverse<br />

complement sequence of the target RNA was used as<br />

template for the in vitro transcription (pol2siRNA)<br />

and an additional T7 promotor sequence<br />

(5¢-CCTGTCTC-3¢) on it’s 3¢end was used (5¢-AGT-<br />

CAATTTGTCAGCGTCCcctgtctc-3¢). The transcription<br />

reaction and the gel purification of the<br />

labeled probe were performed according to the<br />

manufacturer’s instructions. Visualization was carried<br />

out by phosphor-imaging (Imagine Plate Type<br />

Bas-III and FLA-2000; Fujifilm, Du¨sseldorf,<br />

Germany) with an exposition time of 30 to 60 s.<br />

Detection of siRNA<br />

Pol2-siRNA was detected with the mirVana miR-<br />

NA Detection Kit (Ambion) according to the<br />

manufacture’s instructions, detection was carried<br />

out by phosphor-imaging with an exposure time of<br />

3 to 7 min.<br />

One-step RT real-time PCR<br />

Quantitative real-time RT PCR was performed<br />

using the SuperScript One-Step RT-qPCR System<br />

with PlatinumÒ Taq DNA polymerase (Invitrogen)<br />

and the MX4000 thermocycler (Stratagene,<br />

La Jolla, CA, USA). A FAM-labeled probe (5¢-<br />

FAM-AGAAGGGACCTTGGCAGACTTTCT-<br />

BHQ1; Sigma) as well as primers specific for PERV<br />

gag, amplifying the viral full length RNA of all<br />

three subtypes (for: TCCAGGGCTCATAATTT-<br />

GTC, rev: TGATGGCCATCCAACATCGA)<br />

were applied. PERV expression was normalized<br />

to the amount of total RNA as well as to the<br />

expression of the house-keeping genes glycerinaldehyd-3-phosphate-dehydrogenase<br />

(GAPDH),<br />

cyclophilin and hypoxanthine-guanine phosphoribosyltransferase<br />

(HPRT). A HEX-labeled probe<br />

and primers specific for porcine GAPDH (5¢-HEX-<br />

CCACCAACCCCAGCAAGAGCACGC-BHQ1,<br />

for: 5¢-ACATGGCCTCCAAGGAGTAAGA, rev:<br />

GATCGAGTTGGGGCTGTGACT) (Operon,<br />

Cologne, Germany), a Cy5-labeled probe and<br />

primers specific for porcine cyclophilin (5¢-Cy5-<br />

TGCCAGGGTGGTGACTTCACACGCC-BHQ2,<br />

for: 5¢-TGCTTTCACAGAATAATTCCAGGA-<br />

TTTA, rev: 5¢-GACTTGCCACCAGTGCCAT-<br />

TA, Operon), as well as a FAM-labeled<br />

probe and primers specific for porcine HPRT<br />

(FAM-ATCGCCCGTTGACTGGTCATTACA-<br />

GTAGCT-BHQ1, for: 5¢-GTGATAGATCCATT-<br />

CCTATGACTGTAGA, rev: 5¢-TGAGAGA<br />

TCATCTCCACCAATTACTT were used [55].<br />

For each reaction, 50 ng total RNA and the<br />

following temperature conditions were used:<br />

50 °C, 15 min; 95 °C, 2 min; 45 cycles (95 °C,<br />

15 s; 54 °C, 30 s). PERV expression in each organ<br />

was normalized to porcine GAPDH, porcine<br />

cyclophilin and total RNA, respectively, and compared<br />

with the expression in the corresponding<br />

organ of the control animal. Samples from control<br />

and experimental animals were always run together.<br />

Data were analyzed using the DDCTmethod<br />

[56].<br />

Western blot analyzes<br />

Western blot analyzes using sera specific for p15E<br />

and p27Gag of PERV were performed as described<br />

[36,57].<br />

Results<br />

Knockdown of PERV in pigs<br />

shRNA-mediated inhibition of PERV expression in primary<br />

fibroblasts<br />

Porcine fetal fibroblasts (primary cell line: P1 F10)<br />

were transduced using the lentiviral vector<br />

pLVTHM-pol2 (Fig. 1) expressing the PERV-specific<br />

shRNA pol2. In order to analyze the shRNAmediated<br />

inhibition of PERV-mRNA expression,<br />

total RNA was isolated from transduced and nontransduced<br />

fibroblasts and one-step RT real-time<br />

PCR was performed. PERV full-length mRNA<br />

expression was inhibited by 94.8% in fetal fibroblasts<br />

P1 F10 transduced with pLVTHM-pol2 as<br />

described previously [36]. The inhibition of PERV<br />

expression in transduced cells was stable and<br />

remained at 95% over months.<br />

Somatic cell nuclear transfer and the production of shRNA<br />

transgenic pigs and control animals<br />

The pLVTHM-pol2-transduced fibroblasts were<br />

cultured in vitro prior to <strong>bei</strong>ng used for the<br />

production of transgenic pigs via somatic cell<br />

nuclear transfer. After activation, 115 cloned<br />

embryos were transferred surgically into the<br />

oviducts of one synchronized recipient. Pregnancy<br />

was confirmed by ultrasound scanning on days 25<br />

and 35 after embryo transfer. After 116 days of<br />

39


Dieckhoff et al.<br />

gestation seven piglets were born; one of them<br />

was stillborn. None of the piglets showed any<br />

malformations. Mean birth weight was 1.3 kg<br />

which is similar to that of non-transgenic piglets<br />

in our pig facility. Ear biopsies were obtained<br />

from six piglets and frozen in liquid nitrogen.<br />

Four piglets died soon after birth due to agalactia<br />

of the sow. Following repeated application of<br />

2 ml oxytocin (Oxytocin 10 IE/ml; Pharma Partner,<br />

Hamburg, Germany) lactation was reinitiated<br />

and two piglets survived. They were killed by<br />

barbiturate overdose at day 3, and samples of<br />

heart, lungs, spleen, liver, kidney, pancreas as well<br />

as muscle were taken and frozen in liquid nitrogen<br />

for further analysis. Control animals were produced<br />

by nuclear transfer using primary fibroblasts<br />

from non-transduced porcine fibroblasts P1<br />

F10. At day 89 of gestation, organs of five<br />

animals were obtained by Caesarian section, the<br />

fetuses had a mean weight of 1.1 kg.<br />

Detection of vector integration<br />

To detect the presence of the transgene, the<br />

lentiviral vector pLVTHM-pol2, in the genome of<br />

the piglets, PCRs were performed with primers<br />

specific for gfp and specific for the shRNA expression<br />

cassette (Fig. 1). All of the six tested piglets<br />

were positive for gfp (piglet 1 was not tested) and<br />

the shRNA expression cassette (all animals were<br />

tested positive), respectively (Fig. 2).<br />

Fig. 2. (A) Integration of the lentiviral vector pLVTHM-pol2<br />

in six piglets demonstrated by PCR analysis using primers<br />

specific for GFP, and (B) primers specific for the shRNA<br />

expression cassette, positive control (+): vector pLVTHMpol2,<br />

negative control ()): DNA of non-transduced porcine<br />

fibroblasts. N.t. not tested.<br />

40<br />

Fig. 3. Expression of the PERV-specific pol2-shRNA in different<br />

tissues of piglets no. 6 and 7. As positive control<br />

pLVTHM-pol2-transduced PK-15 cells and as negative controls<br />

pLVTHM-transduced and non-transduced PK-15 cells<br />

were used. Internal negative controls: no target RNA (T)/<br />

RNase (the single stranded probe was degraded by RNase<br />

treatment), no target RNA/ no RNase (the full-length probe of<br />

29 nt was not digested).<br />

Expression of the pol2-shRNA in vivo<br />

For analysis of the expression of the pol2-shRNA<br />

in different organs, small RNA molecules were<br />

isolated from the organs of piglets no. 6 and 7 and<br />

used for Northern solution hybridization with a<br />

32 P-labeled short RNA probe corresponding the<br />

shRNA sequence. Expression of pol2-shRNA was<br />

detected in all organs, including heart, lung, spleen,<br />

liver, kidney, muscle, and pancreas (Fig. 3).<br />

Inhibition of PERV expression in different tissues<br />

Total RNA was isolated from different organs of<br />

piglets no. 6 and 7 and analyzed using one-step RT<br />

real-time PCR. As control, total RNA isolated<br />

from the corresponding organs taken from five<br />

control, non-transgenic pigs was used. Expression<br />

of PERV in different organs of the shRNAtransgenic<br />

and the non-transgenic control animals<br />

was highest in the spleen and lower expression was<br />

found in the liver. Due to the differences in PERV<br />

expression related to individual organs, PERV<br />

expression was compared in shRNA-transgenic<br />

and non-transgenic animals on a per organ basis.<br />

The expression levels of PERV in each organ of<br />

the control pigs were averaged and set 100%.<br />

PERV-expression in the corresponding organs<br />

from the shRNA-pol2 transgenic piglets was<br />

related to the expression in the control organ.<br />

Data were normalized to the amount of total<br />

RNA or the expression of the house-keeping<br />

genes GAPDH, cyclophilin, and HPRT. In all<br />

organs of the pol2-shRNA transgenic piglets no. 6<br />

and 7 PERV expression was significantly inhibited<br />

by up to 94% (Fig. 4). Irrespective of the type of


normalization (total RNA amount or housekeeping<br />

genes GAPDH, cyclophilin, and HPRT)<br />

similar expression patterns were observed. Noteworthy,<br />

expression of PERV in organs of the<br />

control animals was similar to the expression of<br />

PERV in adult animals. Preliminary data showed<br />

an expression of 2446 copies of PERV/ng in the<br />

hearts from five control animals, 500 and 492 in<br />

the hearts of the transgenic piglets 6 and 7,<br />

respectively. 4564 copies/ng were found in the<br />

control lungs vs. 1337 and 936 in the lungs of the<br />

transgenic piglets 6 and 7. In the control liver<br />

1614 copies/ng were detected vs. 166 and 328 in<br />

the livers of the transgenic piglets, in the spleens<br />

34068 vs. 1908 and 2419, in the kidneys 2532 vs.<br />

975 and 390, showing the efficacy of the RNAi<br />

in vivo.<br />

Undedectibility of PERV protein expression and virus release<br />

When PERV protein expression was studied using<br />

Western blot analyses of homogenates from different<br />

tissues using antisera specific for different<br />

PERV proteins (p15E, p27Gag), no PERV protein<br />

expression was observed, neither in the non-transgenic<br />

controls nor in the transgenic animals (not<br />

shown).<br />

Discussion<br />

Results of the present study demonstrate a strong<br />

and long lasting inhibition of PERV mRNA<br />

expression in primary porcine fibroblasts using a<br />

Knockdown of PERV in pigs<br />

Fig. 4. Inhibition of PERV expression in different tissues of the pol2-shRNA transgenic piglets (p) no. 6 and 7. PERV expression in<br />

each organ of control non-transgenic animals (c) measured by one-step RT real-time PCR was set 100% and compared with PERV<br />

expression in the corresponding organ of the transgenic animals. In case of control animals the standard deviation is based on the<br />

measurement of PERV expression in organs from five different animals, measured in triplicate; in case of transgenic animals the<br />

standard deviation is based on the measurement of one organ measured in triplicate in two assays. Expression was normalized to the<br />

amount of total RNA, measured by one-step RT real-time PCR.<br />

lentiviral vector system expressing a PERV-specific<br />

shRNA and use in somatic cell nuclear transfer.<br />

This is the first report demonstrating shRNAmediated<br />

reduction of PERV expression in<br />

shRNA-transgenic pigs in vivo. The strategy<br />

applied here could lead to microbiological safer<br />

porcine xenotransplants and is therefore of considerable<br />

medical importance.<br />

More than fifty copies of porcine endogenous<br />

retroviruses are integrated in the porcine genome<br />

[6,58,59], rendering conventional gene-knockout<br />

approaches not feasible and making RNAi a<br />

promising alternative. The RNAi approach has<br />

already been used for the in vitro inhibition of HIV<br />

[60–62] and other viral pathogens like hepatitis B<br />

virus (HBV) [63,64], Coxsackie virus B3 [65], West<br />

Nile virus [66], and transmissible gastroenteritis<br />

virus [51]. Furthermore, RNAi-mediated reduced<br />

HBV expression has been demonstrated in mice in<br />

vivo [67,68] by murine hydronamic tail vene<br />

injection. Long-term inhibition of HBV was<br />

achieved by adeno-associated vector-mediated<br />

RNAi [69]. RNAi was also induced by lentiviral<br />

gene transfer in transgenic mice with a long-term<br />

inhibition of the target gene [70]. In contrast, HIV-<br />

1 escapes from RNAi-mediated inhibition not only<br />

through nucleotide substitutions or deletions in the<br />

siRNA target sequence, but also through mutations<br />

that alter the secondary RNA structure [71].<br />

Resistance and/or escape from the inhibitory<br />

effects of RNAi are impossible in the case of<br />

PERV, since its genome behaves like a cellular<br />

gene.<br />

41


Dieckhoff et al.<br />

In future experiments, the number of integrated<br />

shRNA vectors as well as the influence on the<br />

expression of PERV during life time and the<br />

persistence of the shRNA expression in the next<br />

generation will be analyzed. The physiological role<br />

of endogenous retroviruses is unknown. Our<br />

results indicate that the siRNA expression did<br />

not compromise the fetal development because the<br />

siRNA transgenic piglets were born after a normal<br />

gestation period, with a normal birth weight and<br />

without malformations. Due to the sequence specificity<br />

of the RNAi mechanism and the fact that<br />

PERV is a retroviral sequence, off-target effects<br />

resulting in gene silencing of vital cellular proteins<br />

are highly unlikely. Some RNAi off-target effects<br />

have recently been reported when interfering with<br />

the expression of cellular target genes [72–75]. Due<br />

to the early death of the pol2-shRNA-transgenic<br />

piglets caused by agalactia of the sow, long-term<br />

observations of the inhibitory effect could not yet<br />

be made. In transgenic mice expression of the Neil1<br />

gene, a DNA N-glycosylase, was inhibited by<br />

shRNA over an extended period of time in all<br />

organs [76].<br />

The detection of the inhibitory effect of the pol2shRNA<br />

on PERV expression in vivo requires a<br />

suitable control, especially in the presence of a very<br />

low expression of the PERV in the non-transgenic<br />

founder pigs. In addition, PERV expression varies<br />

in different pig strains [9,31] as well as in different<br />

tissues and organs of one individuum [77,78]. Thus<br />

detailed analysis of non-transgenic pigs of the same<br />

strain provides only limited insight into the RNAi<br />

effects. To provide an appropriate control, pigs<br />

were produced by somatic cell nuclear transfer<br />

using the non-transfected counterparts of the<br />

identical primary cell isolate. This allowed us to<br />

compare PERV expression between animals and<br />

organs isolated from different donors and this<br />

comparison revealed a significant reduction of<br />

PERV expression in the siRNA transgenic pigs.<br />

Expression data were normalized both to the<br />

amount of total RNA and to the expression of<br />

different house-keeping genes such as GAPDH,<br />

cyclophilin, and HPRT. HPRT is considered to be<br />

a suitable reference gene for the analyses of gene<br />

expression in different porcine tissues and immune<br />

cells [79,80]. Numerous studies on GAPDH expression<br />

in humans and rodents showed changes<br />

related to different experimental conditions. In<br />

addition, the existence of pseudogenes for HPRT<br />

[81] and GAPDH [82] complicates the use of these<br />

genes as standards in RT-PCR reactions. However,<br />

although the expression of all house-keeping genes<br />

used here was different in different organs, normalization<br />

of PERV expression according to the<br />

42<br />

expression of each of these house-keeping genes<br />

showed a similar inhibitory effect of PERV expression<br />

up to 95%.<br />

Based on the low expression of viral RNA also a<br />

low expression of PERV proteins was expected. In<br />

Western blot assays using sera specific for p15E<br />

and p27Gag of PERV, no protein expression was<br />

observed, neither in the non-transgenic controls<br />

nor in the transgenic animals. Using the same<br />

assays PERV protein expression was also not<br />

detected in primary PBMCs from different pig<br />

strains such as German landrace, Duroc, Schwa¨bisch<br />

Ha¨llisch, German large white as well as<br />

crossbreeds. In PK-15 cells, producing detectable<br />

amounts of infectious virus particles, expression of<br />

p27Gag was detectable, but still very low and this<br />

expression was reduced by RNAi nearly to 100%<br />

[36]. Higher expression of p27Gag protein was<br />

observed in human 293 cells infected with a high<br />

titer strain of PERV, which was only partially<br />

inhibited by RNAi [33]. A higher expression of<br />

PERV proteins was also observed in pig melanomas<br />

and in melanoma cell lines [57].<br />

The absence of viral proteins in non-transgenic<br />

and transgenic pigs suggested the absence of PERV<br />

release. However, the low sensitivity of the Western<br />

blot assay (detection limit: 50 ng for p15E and<br />

100 ng for p27Gag [57] left questions open.<br />

Improving the assays using monoclonal antibodies,<br />

developing a sensitive immunohistochemistry analysis<br />

or using a sensitive assay to detect RT activity<br />

in the plasma, such as the PERT (productenhanced<br />

reverse transcriptase) assay [83] may<br />

help to compare protein expression and virus<br />

release in transgenic and non-transgenic animals<br />

more correctly.<br />

To increase the safety of porcine xenotransplants,<br />

multiple shRNAs corresponding to different<br />

regions of the PERV genome shall be used. The<br />

advantage of the RNAi strategy is, that the<br />

expression of all PERVs will be inhibited and the<br />

likelihood of recombinations or complementation<br />

is further decreased. Selection of PERV lowproducer<br />

pigs using methods described earlier<br />

[9,31] together with the production of shRNAtransgenic<br />

animals may reduce PERV expression in<br />

cells and organs and thus increase the viral safety<br />

of porcine xenografts.<br />

Acknowledgments<br />

The authors gratefully acknowledge funding by<br />

DFG within the TransRegio research group 535.<br />

The competent assistance of the staff of the<br />

Department of Biotechnology in the production<br />

of cloned pigs is gratefully acknowledged.


References<br />

1. Cooper DK. Clinical xenotransplantion – how close are<br />

we? Lancet 2003; 362: 557–559.<br />

2. Kues WA, Niemann H. The contribution of farm animals<br />

to human health. Trends Biotechnol 2004; 22: 286–294.<br />

3. Dai Y, Vaught TD, Boone J et al. Targeted disruption of<br />

the alpha1,3-galactosyltransferase gene in cloned pigs. Nat<br />

Biotechnol 2002; 20: 251–255.<br />

4. Lai L, Prather RS. Progress in producing knockout<br />

models for xenotransplantation by nuclear transfer. Ann<br />

Med 2002; 34: 501–506.<br />

5. Niemann H. Current status and perspectives for the generation<br />

of transgenic pigs for xenotransplantation. Ann<br />

Transplant 2001; 6: 6–9.<br />

6. Patience C, Switzer WM, Takeuchi Y et al. Multiple<br />

groups of novel retroviral genomes in pigs and related<br />

species. J Virol 2001; 75: 2771–2775.<br />

7. Ericsson T, Oldmixon B, Blomberg J, Rosa M,<br />

Patience C, Andersson G. Identification of novel porcine<br />

endogenous betaretrovirus sequences in miniature<br />

swine. J Virol 2001; 75: 2765–2770.<br />

8. Martin U, Kiessig V, Blusch JH et al. Expression of pig<br />

endogenous retrovirus by primary porcine endothelial cells<br />

and infection of human cells. Lancet 1998; 352: 692–694.<br />

9. Tacke S, Specke V, Stephan O, Seibold E, Bodusch K,<br />

Denner J. Porcine endogenous retroviruses: diagnostic<br />

assays and evidence for immunosuppressive properties.<br />

Transplant Proc 2000; 32: 1166.<br />

10. McIntyre MC, Kannan B, Solano-Aguilar GI, Wilson<br />

CA, Bloom ET. Detection of porcine endogenous<br />

retrovirus in cultures of freshly isolated porcine bone<br />

marrow cells. Xenotransplantation 2003; 10: 337–342.<br />

11. Patience C, Takeuchi Y, Weiss RA. Infection of human<br />

cells by an endogenous retrovirus of pigs. Nat Med 1997;<br />

3: 282–286.<br />

12. Takeuchi Y, Patience C, Magre S et al. Host range and<br />

interference studies of three classes of pig endogenous<br />

retrovirus. J Virol 1998; 72: 9986–9991.<br />

13. Martin U, Winkler ME, Id M et al. Productive infection<br />

of primary human endothelial cells by pig endogenous<br />

retrovirus (PERV). Xenotransplantation 2000; 7: 138–142.<br />

14. Wilson CA, Wong S, VanBrocklin M, Federspiel MJ.<br />

Extended analysis of the in vitro tropism of porcine<br />

endogenous retrovirus. J Virol 2000; 74: 49–56.<br />

15. Specke V, Rubant S, Denner J. Productive infection of<br />

human primary cells and cell lines with porcine endogenous<br />

retroviruses. Virology 2001; 285: 177–180.<br />

16. Le Tissier P, Stoye JP, Takeuchi Y, Patience C, Weiss<br />

RA. Two sets of human-tropic pig retrovirus. Nature<br />

1997; 389: 681–682.<br />

17. Gao F, Yue L, Robertson DL et al. Genetic diversity of<br />

human immunodeficiency virus type 2: evidence for distinct<br />

sequence subtypes with differences in virus biology.<br />

J Virol 1994; 68: 7433–7447.<br />

18. Gao F, Bailes E, Robertson DL et al. Origin of HIV-1<br />

in the chimpanzee Pan troglodytes troglodytes. Nature<br />

1999; 397: 436–441.<br />

19. Heneine W, Tibell A, Switzer WM et al. No evidence<br />

of infection with porcine endogenous retrovirus in<br />

recipients of porcine islet-cell xenografts. Lancet 1998;<br />

352: 695–699.<br />

20. Garkavenko O, Croxson MC, Irgang M, Karlas A,<br />

Denner J, Elliott RB. Monitoring for presence of<br />

potentially xenotic viruses in recipients of pig islet xenotransplantation.<br />

J Clin Microbiol 2004; 42: 5353–5356.<br />

Knockdown of PERV in pigs<br />

21. Irgang M, Sauer IM, Karlas A et al. Porcine endogenous<br />

retroviruses: no infection in patients treated with a<br />

bioreactor based on porcine liver cells. J Clin Virol 2003;<br />

28: 141–154.<br />

22. Xu H, Sharma A, Okabe J et al. Serologic analysis of<br />

anti-porcine endogenous retroviruses immune responses in<br />

humans after ex vivo transgenic pig liver perfusion.<br />

ASAIO J 2003; 49: 407–416.<br />

23. Paradis K, Langford G, Long Z et al. Search for crossspecies<br />

transmission of porcine endogenous retrovirus in<br />

patients treated with living pig tissue. The XEN 111 Study<br />

Group. Science 1999; 285: 1236–1241.<br />

24. Tacke SJ, Bodusch K, Berg A, Denner J. Sensitive and<br />

specific immunological detection methods for porcine<br />

endogenous retroviruses applicable to experimental and<br />

clinical xenotransplantation. Xenotransplantation 2001;<br />

8: 125–135.<br />

25. Loss M, Arends H, Winkler M et al. Analysis of<br />

potential porcine endogenous retrovirus (PERV) transmission<br />

in a whole-organ xenotransplantation model<br />

without interfering microchimerism. Transpl Int 2001; 14:<br />

31–37.<br />

26. Winkler ME, Winkler M, Burian R et al. Analysis of<br />

pig-to-human porcine endogenous retrovirus transmission<br />

in a triple-species kidney xenotransplantation model.<br />

Transpl Int 2005; 17: 848–858. Epub 2005 Apr.<br />

27. Specke V, Schuurman HJ, Plesker R et al. Virus safety<br />

in xenotransplantation: first exploratory in vivo studies in<br />

small laboratory animals and non-human primates.<br />

Transpl Immunol 2002; 9: 281–288.<br />

28. Denner J. Immunosuppression by retroviruses: implications<br />

for xenotransplantation. Ann N Y Acad Sci 1998;<br />

862: 75–86.<br />

29. Tacke SJ, Kurth R, Denner J. Porcine endogenous<br />

retroviruses inhibit human immune cell function: risk for<br />

xenotransplantation? Virology 2000; 268: 87–93.<br />

30. Oldmixon BA, Wood JC, Ericsson TA et al. Porcine<br />

endogenous retrovirus transmission characteristics of an<br />

inbred herd of miniature swine. J Virol 2002; 76: 3045–3048.<br />

31. Tacke SJ, Specke V, Denner J. Differences in release and<br />

determination of subtype of porcine endogenous retroviruses<br />

produced by stimulated normal pig blood cells.<br />

Intervirology 2003; 46: 17–24.<br />

32. Fiebig U, Stephan O, Kurth R, Denner J. Neutralizing<br />

antibodies against conserved domains of p15E of porcine<br />

endogenous retroviruses: basis for a vaccine for xenotransplantation?<br />

Virology 2003; 307: 406–413.<br />

33. Karlas A, Kurth R, Denner J. Inhibition of porcine<br />

endogenous retroviruses by RNA interference: increasing<br />

the safety of xenotransplantation. Virology 2004; 325: 18–<br />

23.<br />

34. Miyagawa S, Nakatsu S, Nakagawa T et al. Prevention<br />

of PERV infections in pig to human xenotransplantation<br />

by the RNA interference silences gene. J Biochem (Tokyo)<br />

2005; 137: 503–508.<br />

35. Niemann H, Kues WA. Transgenic farm animals – an<br />

update. Reprod Fertil Dev 2007; 19: 762–770.<br />

36. Dieckhoff B, Karlas A, Hofmann A et al. Inhibition of<br />

porcine endogenous retroviruses (PERVs) in primary<br />

porcine cells by RNA interference using lentiviral vectors.<br />

Arch Virol 2007; 152: 629–634. Epub 2006 Nov.<br />

37. Caplen NJ, Parrish S, Imani F, Fire A, Morgan RA.<br />

Specific inhibition of gene expression by small doublestranded<br />

RNAs in invertebrate and vertebrate systems.<br />

Proc Natl Acad Sci U S A 2001; 98: 9742–9747. Epub 2001<br />

Jul.<br />

43


Dieckhoff et al.<br />

38. Hammond SM, Bernstein E, Beach D, Hannon GJ. An<br />

RNA-directed nuclease mediates post-transcriptional gene<br />

silencing in Drosophila cells. Nature 2000; 404: 293–296.<br />

39. Billy E, Brondani V, Zhang H, Muller U, Filipowicz<br />

W. Specific interference with gene expression induced by<br />

long, double-stranded RNA in mouse embryonal teratocarcinoma<br />

cell lines. Proc Natl Acad Sci U S A 2001; 98:<br />

14428–14433. Epub 2001 Nov.<br />

40. Yang D, Lu H, Erickson JW. Evidence that processed<br />

small dsRNAs may mediate sequence-specific mRNA<br />

degradation during RNAi in Drosophila embryos. Curr<br />

Biol 2000; 10: 1191–1200.<br />

41. Elbashir SM, Lendeckel W, Tuschl T. RNA interference<br />

is mediated by 21- and 22-nucleotide RNAs. Genes<br />

Dev 2001; 15: 188–200.<br />

42. McManus MT, Petersen CP, Haines BB, Chen J,<br />

Sharp PA. Gene silencing using micro-RNA designed<br />

hairpins. RNA 2002; 8: 842–850.<br />

43. Stein P, Svoboda P, Anger M, Schultz RM. RNAi:<br />

mammalian oocytes do it without RNA-dependent RNA<br />

polymerase. RNA 2003; 9: 187–192.<br />

44. Brummelkamp TR, Bernards R, Agami R. A system for<br />

stable expression of short interfering RNAs in mammalian<br />

cells. Science 2002; 296: 550–553. Epub 2002 Mar 2021.<br />

45. Paddison PJ, Caudy AA, Hannon GJ. Stable suppression<br />

of gene expression by RNAi in mammalian ells. Proc<br />

Natl Acad Sci U S A 2002; 99: 1443–1448. Epub 2002 Jan.<br />

46. Paul CP, Good PD, Winer I, Engelke DR. Effective<br />

expression of small interfering RNA in human cells. Nat<br />

Biotechnol 2002; 20: 505–508.<br />

47. Sui G, Soohoo C, Affar el B et al. A DNA vector-based<br />

RNAi technology to suppress gene expression in mammalian<br />

cells. Proc Natl Acad Sci U S A 2002; 99: 5515–<br />

5520.<br />

48. Grimm D, Streetz KL, Jopling CL et al. Fatality in mice<br />

due to oversaturation of cellular microRNA/short hairpin<br />

RNA pathways. Nature 2006; 441: 537–541.<br />

49. Rao MK, Wilkinson MF. Tissue-specific and cell typespecific<br />

RNA interference in vivo. Nat Protoc 2006; 1:<br />

1494–1501.<br />

50. Tsuchiya M, Yoshida T, Taniguchi S et al. In vivo<br />

suppression of mafA mRNA with siRNA and analysis of<br />

the resulting alteration of the gene expression profile in<br />

mouse pancreas by the microarray method. Biochem<br />

Biophys Res Commun 2007; 356: 129–135. Epub 2007<br />

Feb.<br />

51. Zhou JF, Hua XG, Cui L et al. Effective inhibition of<br />

porcine transmissible gastroenteritis virus replication in ST<br />

cells by shRNAs targeting RNA-dependent RNA polymerase<br />

gene. Antiviral Res 2007; 74: 36–42. Epub 2007<br />

Jan.<br />

52. Golding MC, Long CR, Carmell MA, Hannon GJ,<br />

Westhusin ME. Suppression of prion protein in livestock<br />

by RNA interference. Proc Natl Acad Sci U S A 2006; 103:<br />

5285–5290. Epub 2006 Mar.<br />

53. Kues W, Anger M, Motlik J, Carnwath JW, Niemann<br />

H, Paul D. Cell cycle synchronization of porcine primary<br />

fibroblasts: effects of serum deprivation and reversible cell<br />

cycle inhibitors. Biol Reprod 2000; 62: 412–419.<br />

54. Holker M, Petersen B, Hassel P et al. Duration of in<br />

vitro maturation of recipient oocytes affects blastocyst<br />

development of cloned porcine embryos. Cloning Stem<br />

Cells 2005; 7: 35–44.<br />

55. Duvigneau JC, Hartl RT, Groiss S, Gemeiner M.<br />

Quantitative simultaneous multiplex real-time PCR for the<br />

detection of porcine cytokines. J Immunol Methods 2005;<br />

306: 16–27. Epub 2005 Sep.<br />

44<br />

56. Livak KJ, Schmittgen TD. Analysis of relative gene<br />

expression data using real-time quantitative PCR and the<br />

2(-Delta Delta C(T)) Method. Methods 2001; 25: 402–408.<br />

57. Dieckhoff B, Puhlmann J, Büscher K et al. Expression<br />

of porcine endogenous retroviruses (PERVs) in melanomas<br />

of MMS Troll pigs. Vet Microbiol 2007; 123: 53–68.<br />

58. Bosch S, Arnauld C, Jestin A. Study of full-length<br />

porcine endogenous retrovirus genomes with envelope<br />

gene polymorphism in a specific-pathogen-free large white<br />

swine herd. J Virol 2000; 74: 8575–8581.<br />

59. Herring C, Quinn G, Bower R et al. Mapping fulllength<br />

porcine endogenous retroviruses in a large white<br />

pig. J Virol 2001; 75: 12252–12265.<br />

60. Coburn GA, Cullen BR. Potent and specific inhibition<br />

of human immunodeficiency virus type 1 replication by<br />

RNA interference. J Virol 2002; 76: 9225–9231.<br />

61. Banerjea A, Li MJ, Bauer G et al. Inhibition of HIV-1<br />

by lentiviral vector-transduced siRNAs in T lymphocytes<br />

differentiated in SCID-hu mice and CD34+ progenitor<br />

cell-derived macrophages. Mol Ther 2003; 8: 62–71.<br />

62. Christensen HS, Daher A, Soye KJ et al. Small interfering<br />

RNAs against the TAR RNA binding protein,<br />

TRBP, a Dicer cofactor, inhibit human immunodeficiency<br />

virus type 1 long terminal repeat expression and viral<br />

production. J Virol 2007; 81: 5121–5131. Epub 2007 Mar.<br />

63. Radhakrishnan SK, Layden TJ, Gartel AL. RNA<br />

interference as a new strategy against viral hepatitis.<br />

Virology 2004; 323: 173–181.<br />

64. Kayhan H, Karatayli E, Turkyilmaz AR, Sahin F,<br />

Yurdaydin C, Bozdayi AM. Inhibition of hepatitis B<br />

virus replication by shRNAs in stably HBV expressed<br />

HEPG2 2.2.15 cell lines. Arch Virol 2007; 152: 871–879.<br />

Epub 2007 Jan.<br />

65. Fechner H, Pinkert S, Wang X et al. Coxsackievirus B3<br />

and adenovirus infections of cardiac cells are efficiently<br />

inhibited by vector-mediated RNA interference targeting<br />

their common receptor. Gene Ther 2007; 22: 22.<br />

66. Kachko AV, Ivanova AV, Protopopova EV, Netesov<br />

V, Loktev VB. Inhibition of West Nile virus replication<br />

by short interfering RNAs. Dokl Biochem Biophys 2006;<br />

410: 260–262.<br />

67. Ying RS, Zhu C, Fan XG et al. Hepatitis B virus is<br />

inhibited by RNA interference in cell culture and in mice.<br />

Antiviral Res 2007; 73: 24–30.<br />

68. Weinberg MS, Ely A, Barichievy S et al. Specific inhibition<br />

of HBV replication in vitro and in vivo with expressed<br />

long hairpin RNA. Mol Ther 2007; 15: 534–541.<br />

69. Chen CC, Ko TM, Ma HI et al. Long-term inhibition of<br />

hepatitis B virus in transgenic mice by double-stranded<br />

adeno-associated virus 8-delivered short hairpin RNA.<br />

Gene Ther 2007; 14: 11–19. Epub 2006 Aug.<br />

70. Rubinson DA, Dillon CP, Kwiatkowski AV et al. A<br />

lentivirus-based system to functionally silence genes in<br />

primary mammalian cells, stem cells and transgenic mice<br />

by RNA interference. Nat Genet 2003; 33: 401–406. Epub<br />

2003 Feb.<br />

71. Westerhout EM, Ooms M, Vink M, Das AT, Berkhout<br />

B. HIV-1 can escape from RNA interference by<br />

evolving an alternative structure in its RNA genome.<br />

Nucleic Acids Res 2005; 33: 796–804.<br />

72. Jackson AL, Bartz SR, Schelter J et al. Expression<br />

profiling reveals off-target gene regulation by RNAi. Nat<br />

Biotechnol 2003; 21: 635–637. Epub 2003 May.<br />

73. Jackson AL, Linsley PS. Noise amidst the silence: offtarget<br />

effects of siRNAs? Trends Genet 2004; 20: 521–524.<br />

74. Persengiev SP, Zhu X, Green MR. Nonspecific, concentration-dependent<br />

stimulation and repression of


mammalian gene expression by small interfering RNAs<br />

(siRNAs). RNA 2004; 10: 12–18.<br />

75. Scacheri PC, Rozenblatt-Rosen O, Caplen NJ et al.<br />

Short interfering RNAs can induce unexpected and<br />

divergent changes in the levels of untargeted proteins in<br />

mammalian cells. Proc Natl Acad Sci U S A 2004; 101:<br />

1892–1897.<br />

76. Carmell MA, Zhang L, Conklin DS, Hannon GJ,<br />

Rosenquist TA. Germline transmission of RNAi in mice.<br />

Nat Struct Biol 2003; 10: 91–92.<br />

77. Akiyoshi DE, Denaro M, Zhu H, Greenstein JL,<br />

Banerjee P, Fishman JA. Identification of a full-length<br />

cDNA for an endogenous retrovirus of miniature swine.<br />

J Virol 1998; 72: 4503–4507.<br />

78. Clemenceau B, Lalain S, Martignat L, Sai P. Porcine<br />

endogenous retroviral mRNAs in pancreas and a panel of<br />

tissues from specific pathogen-free pigs. Diabetes Metab<br />

1999; 25: 518–525.<br />

79. Foss DL, Baarsch MJ, Murtaugh MP. Regulation of<br />

hypoxanthine phosphoribosyltransferase, glyceraldehyde-<br />

Knockdown of PERV in pigs<br />

3-phosphate dehydrogenase and beta-actin mRNA<br />

expression in porcine immune cells and tissues. Anim<br />

Biotechnol 1998; 9: 67–78.<br />

80. Ledger TN, Pinton P, Bourges D, Roumi P, Salmon<br />

H, Oswald IP. Development of a macroarray to specifically<br />

analyze immunological gene expression in swine. Clin<br />

Diagn Lab Immunol 2004; 11: 691–698.<br />

81. Sellner LN, Turbett GR. The presence of a pseudogene<br />

may affect the use of HPRT as an endogenous mRNA<br />

control in RT-PCR. Mol Cell Probes 1996; 10: 481–483.<br />

82. Piechaczyk M, Blanchard JM, Riaad-El Sabouty S,<br />

Dani C, Marty L, Jeanteur P. Unusual abundance of<br />

vertebrate 3-phosphate dehydrogenase pseudogenes. Nature<br />

1984; 312: 469–471.<br />

83. Bisset LR, Böni J, Lutz H, Schüpbach J. Lack of evidence<br />

for PERV expression after apoptosis-mediated<br />

horizontal gene transfer between porcine and human cells.<br />

Xenotransplantation 2007; 14: 13–24.<br />

45

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!