24.11.2013 Views

Stefan Wirtz Vom Fachbereich VI (Geographie/Geowissenschaften ...

Stefan Wirtz Vom Fachbereich VI (Geographie/Geowissenschaften ...

Stefan Wirtz Vom Fachbereich VI (Geographie/Geowissenschaften ...

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.

<strong>Stefan</strong> <strong>Wirtz</strong><br />

<strong>Vom</strong> <strong>Fachbereich</strong> <strong>VI</strong><br />

(<strong>Geographie</strong>/<strong>Geowissenschaften</strong>)<br />

der Universität Trier<br />

zur Verleihung des akademischen Grades<br />

Doktor der Naturwissenschaften (Dr. rer. nat.)<br />

genehmigte Dissertation<br />

Experimentelle Rinnenerosionsforschung vs.<br />

Modellkonzepte – Quantifizierung der<br />

hydraulischen und erosiven Wirksamkeit von<br />

Rinnen<br />

Betreuende:<br />

Prof. Dr. Johannes B. Ries<br />

Prof. Dr. Jean F. Wagner<br />

Datum der wissenschaftlichen Aussprache:<br />

28. Januar 2013<br />

Trier, 2013


<strong>Stefan</strong> <strong>Wirtz</strong><br />

<strong>Vom</strong> <strong>Fachbereich</strong> <strong>VI</strong><br />

(<strong>Geographie</strong>/<strong>Geowissenschaften</strong>)<br />

der Universität Trier<br />

zur Verleihung des akademischen Grades<br />

Doktor der Naturwissenschaften (Dr. rer. nat.)<br />

genehmigte Dissertation<br />

Experimentelle Rinnenerosionsforschung<br />

vs. Modellkonzepte – Quantifizierung der<br />

hydraulischen und erosiven Wirksamkeit<br />

von Rinnen<br />

Betreuende:<br />

Prof. Dr. Johannes B. Ries<br />

Prof. Dr. Jean F. Wagner<br />

Datum der wissenschaftlichen Aussprache:<br />

28. Januar 2013<br />

Trier, 2013


Danksagung<br />

Danksagung<br />

Die vorliegende Arbeit wäre ohne die Unterstützung von vielen mehr oder weniger<br />

freiwilligen Helfern so nicht zustande gekommen. Daher möchte ich mich bei allen bedanken,<br />

die zum Erfolg (hoffentlich…) des Projektes „Dr. rer. nat.“ beigetragen haben. Das<br />

Endergebnis lag bei Redaktionsschluss noch nicht vor.<br />

Herr Prof. Dr. J.B. Ries: Er hat nicht nur die Betreuung der Arbeit übernommen, sondern<br />

das Interesse für das Thema „Bodenerosion“ bei mir überhaupt erst geweckt. Außerdem<br />

ermöglichte er mir die Teilnahme an vielen interessanten Geländeaufenthalten. Zusätzlich hat<br />

er alles Menschen(un)mögliche unternommen, um Geldtöpfe aufzutreiben, die es eigentlich<br />

gar nicht gab. Dies ermöglichte 1) die Anschaffung von Materialien und Geräten für die<br />

Versuche und 2) dem Autor dieser Arbeit die regelmäßige Nahrungsaufnahme.<br />

Herr Prof. Dr. J.F. Wagner: Er hat sich nach der Diplomarbeit sofort bereiterklärt, auch die<br />

Zweitbetreuung der Dissertation zu übernehmen, obwohl das Thema nicht direkt etwas mit<br />

seinem Arbeitsgebiet zu tun hat. Trotzdem verfolgte er die Fortschritte der Arbeit immer mit<br />

großem Interesse und war zudem an der Erstellung von mehreren Artikeln direkt beteiligt.<br />

Dr. M. Seeger: Er hat den „Rinnenerosions-Stein“ 2005 mit einem denkwürdigen Zitat ins<br />

Rollen gebracht und trotz seiner zwischenzeitlichen Versetzung nach Holland immer noch die<br />

Zeit gefunden, zu diskutieren, Artikel mitzuschreiben und mich auf seine diplomatische Art<br />

notfalls wieder in die richtige Bahn zu treten. Ach ja, das denkwürdige Zitat: „Wir schütten<br />

oben mal Wasser rein und schauen, was so passiert!“ Hier, das ist passiert! Ausgeartet in eine<br />

Doktorarbeit.<br />

Alex, Sabrina, René, Sarah, Thomas, Miriam, Verena, Gilles, Chrissi, Bastienne, Britta:<br />

Sie haben viel Zeit und Geld investiert, um bei einer, zwei, drei oder sogar allen vier<br />

Geländephasen dabei zu sein und den Versuchsaufbau immer besser, schneller und noch<br />

besser zu machen, die Verpflegung sicherzustellen und Poster und/oder Artikel mit zu<br />

verfassen. Leider war nie genug Geld da, um alle Ideen umzusetzen. Bis jetzt…<br />

Die Studentinnen und Studenten der Lehrforschungsprojekte: Es waren diese (nicht<br />

immer) freiwilligen Helfer, die die Versuche im Rahmen von Lehrforschungsprojekten<br />

ermöglicht haben. Ohne sie wären wahrscheinlich nur wenige Experimente gelaufen. Sie<br />

I


Danksagung<br />

mussten mich und meine diplomatische Art jeweils 14 Tage auf dem Campingplatz in Freila<br />

ertragen, manche sogar zweimal. Einige reden sogar noch mit mir.<br />

Herr Prof. Dr. C. Wagner und Dr. A. Zell von der Abteilung für Technische Physik an der<br />

Universität in Saarbrücken. Sie stellten mir die Laboreinrichtungen kostenfrei zur Verfügung<br />

und unterstützten mich bei den Versuchen. Als Belohnung gab es jeweils Co-Autorenschaften<br />

auf einem Artikel und mehreren Postern.<br />

Die Mitarbeiter der Physischen <strong>Geographie</strong>: Sie haben durch technische Unterstützung,<br />

Hilfe bei bürokratischen Problemen oder durch wichtige und gute Vorschläge und Ideen die<br />

Arbeit unterstützt. Grüße an Harry Willger.<br />

Alle Korrekturleser und Reviewer: Neben schon oben genannten Personen seien hier noch<br />

zusätzlich Olli Gronz, Dr. Christoph Müller, Dr. Marco Hümann, Christoph Paulußen, Dr.<br />

Birgit Kausch und Frank <strong>Wirtz</strong> erwähnt. Die Kommentare und Kritiken haben die Qualität der<br />

Arbeit deutlich verbessert.<br />

Meine Eltern: Sie unterstützten mich während des gesamten Studiums. Außerdem mussten<br />

sie immer wieder als „Probe-Auditorium“ für Vorträge herhalten. Jede Hausarbeit sowie die<br />

Diplomarbeit und diese Dissertation wurden auf das Sorgfältigste korrekturgelesen.<br />

Dr. A. Keller und Alexander Weber: Ohne die Trainingseinheiten wäre die Arbeit nie fertig<br />

geworden. Klingt komisch, ist aber so. Sportverrückte wissen, was ich meine. Ach ja, da<br />

waren auch noch die Bildungsreisen nach Namibia, Jordanien und in die DomRep (Ja, auch<br />

das war eine Bildungsreise!!), die den Motivations-Akku regelmäßig aufgeladen haben.<br />

Moncho aus Aisa: Er hat dafür gesorgt, dass die Gamaschengang im Arnástal nicht erfroren<br />

und/oder verhungert ist. Die ca. 100.000 Kalorien pro Person und Tag kredenzt in der<br />

praktischen „Rohrbombenform“ waren knapp aber ausreichend.<br />

Die Liste erhebt keinen Anspruch auf Vollständigkeit, eine Nicht-Erwähnung ist aber nicht<br />

böse gemeint!<br />

II


Inhaltsverzeichnis<br />

Inhaltsverzeichnis<br />

Kurzfassung 1<br />

1. Experimentelle Rinnenerosionsforschung vs. Modellkonzepte –<br />

Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

3<br />

1.1 Einleitung 3<br />

1.2 Experimente zur Rinnenerosion 8<br />

1.3 Rinnen als Abflusssammler 12<br />

1.4 Rinnen als Sedimentquelle 13<br />

1.5 Prozesse in der Rinne 15<br />

1.6 Messen oder Modellieren 19<br />

1.7 Weitere Anwendungen 24<br />

1.8 Zusammenfassung und Ausblick 26<br />

Quellenangaben 32<br />

2. <strong>Wirtz</strong>, S., Iserloh, T., Rock, G., Hansen, R., Marzen, M., Seeger, M., Betz,<br />

S., Remke, A., Wengel, R., Butzen, V., Ries, J.B. (2012b): Soil erosion on<br />

abandoned land in Andalusia – a comparison of interrill– and rill erosion<br />

rates. ISRN Soil Science, doi: 10.5402/2012/730870.<br />

38<br />

3. <strong>Wirtz</strong>, S., Seeger, M., Ries, J.B. (2010): The rill experiment as a method to<br />

approach a quantification of rill erosion process activity. Zeitschrift für<br />

Geomorphologie 54 (1), 47-64.<br />

70<br />

4. <strong>Wirtz</strong>, S., Seeger, M., Ries, J.B. (2012a): Field experiments for<br />

understanding and quantification of rill erosion processes. Catena 91, 21-34.<br />

96<br />

5. <strong>Wirtz</strong>, S., Seeger, M., Remke, A., Wengel, R., Wagner, J.-F., Ries, J.B.<br />

(2013): Do deterministic sediment detachment and transport equations<br />

adequately represent process-interactions in eroding rills? An experimental<br />

field study. Catena 101, 61-78<br />

131<br />

III


Inhaltsverzeichnis<br />

6. <strong>Wirtz</strong>, S., Seeger, M., Zell, A., Wagner, C., Wagner, J.-F., Ries, J.B. (subm.<br />

2012): Applicability of different hydraulic parameters to describe soil<br />

detachment in eroding rills. PLoSOne.<br />

178<br />

7. Ries, J.B., Andres, K., <strong>Wirtz</strong>, S., Tumbrink, J., Wilms, T., Peter, K.D.,<br />

Burczyk, M., Butzen, V., Seeger, M. (subm. 2011): Sheep and goat erosion –<br />

experimental geomorphology as an approach for the quantification of<br />

underestimated processes. Zeitschrift für Geomorphologie.<br />

224<br />

Wissenschaftlicher Werdegang 257<br />

Lebenslauf 258<br />

IV


Inhaltsverzeichnis<br />

Abbildungsverzeichnis<br />

Abbildung 1: Zusammenhänge zwischen den einzelnen Artikeln 7<br />

Abbildung 2: Veränderungen in der Methode „Spülversuch“ 12<br />

Abbildung 3: Idealtypischer Verlauf der Sedimentkonzentration bei Transport von 17<br />

losem Material<br />

Abbildung 4: Idealtypischer Verlauf der Sedimentkonzentration bei Einschneiden in 17<br />

die Rinnensohle<br />

Abbildung 5: Idealtypischer Verlauf der Sedimentkonzentration bei<br />

18<br />

Massenbewegungen an Stufen und Seitenwänden<br />

Abbildung 6: Quotient aus transport rate und transport capacity 20<br />

Abbildung 7: Quotient aus transport rate und transport capacity 20<br />

Abbildung 8: Relativer Messfehler verschiedener Parameter 22<br />

Abbildung 9: Empirischer Variationskoeffizient verschiedener Parameter 22<br />

Abbildung 10: Runoff-Koeffizienten auf Waldwegen 25<br />

Abbildung 11: Abtragswerte auf Waldwegen 26<br />

Abbildung 12: Drehkrankonstruktion 29<br />

Tabellenverzeichnis<br />

Tabelle 1: Englische Begriffe im Text 8<br />

Tabelle 2: Entwicklung der Spülversuche 11<br />

V


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Kurzfassung<br />

In der Bodenerosion zählen Rinnen zu den wirksamsten Formen: Sie stellen bevorzugte<br />

Fließwege für den Oberflächenabfluss dar und werden dadurch zu den ergiebigsten<br />

Sedimentquellen innerhalb eines Einzugsgebietes. Allerdings ist umstritten, wie hoch ihr<br />

Anteil an der Abtragsdynamik eines Gebietes im Vergleich zu anderen<br />

Bodenerosionsprozessen tatsächlich ist. Voraussetzung für die Bearbeitung dieser Thematik<br />

ist eine Vereinheitlichung der Messmethodik zur Quantifizierung der Erosion in Rinnen. Nur<br />

mittels einer reproduzierbaren Messung können Ergebnisse aus verschiedenen Studien<br />

miteinander verglichen und zu einem Gesamtergebnis synthetisiert werden. Im Bereich der<br />

Rinnenerosion ist eine solche messmethodische Standardisierung bisher nicht gegeben.<br />

Aus diesem Grund besteht das erste Ziel der vorliegenden Arbeit darin, einen<br />

Versuchsaufbau zu entwerfen, der es ermöglicht, unter reproduzierbaren Bedingungen<br />

vergleichbare Daten zur Prozessdynamik in Rinnen zu gewinnen. Die schrittweise<br />

Weiterentwicklung der Methode „Spülversuch“ spiegelt sich in den verschiedenen Versionen<br />

wider. Dabei werden jeweils Fehler und Probleme der vorherigen Version beseitigt, sodass<br />

eine stetige Optimierung stattfindet. Mithilfe der Spülversuche sowie zusätzlicher<br />

Beregnungssimulationen und geomorphologischer Kartierungen wird als zweites Ziel<br />

überprüft, wie hoch die Abflusseffektivität der Rinnen in einem Einzugsgebiet und als drittes<br />

Ziel wie hoch der Anteil der Rinnenerosion an der Gesamt-Erosionsdynamik ist. Hierbei zeigt<br />

sich eine hohe Variabilität: So weisen in einem Kleineinzugsgebiet in den spanischen<br />

Pyrenäen die Proben aus den untersuchten Rinnen zwar nur geringe Sedimentkonzentrationen<br />

auf, verglichen mit dem gesamten Einzugsgebiet zeigt sich jedoch die hohe Effektivität, mit<br />

der trotz Mangel an erodierbarem Material das wenige in den Versuchen eingeleitete Wasser<br />

in den Rinnen wirkt. In einem Gully-Einzugsgebiet in Andalusien weisen Rinnen eine 20-60-<br />

fach höhere Abtragsrate als die interrill-Bereiche auf. Die 0,25 % Rinnenfläche entwässert<br />

20 % dieses Gebietes. Deutlich erkennbar ist, dass nicht ein einzelner Prozess das<br />

Sedimentmaterial bereitstellt, vielmehr laufen in der Rinne verschiedene Prozesse mit großer<br />

räumlicher und zeitlicher Variabilität ab. Die Identifizierung und Quantifizierung der<br />

verschiedenen Prozesse kann als viertes Ziel der vorliegenden Arbeit angesehen werden. Es<br />

ist zu erkennen, dass bis zu 60 % der Prozesse, die Sediment mobilisieren, nicht durch<br />

hydraulische Parameter bestimmt sondern eher gravitativ gesteuert sind. Neben der<br />

hydraulisch gesteuerten Einschneidung in die Rinnensohle wird Material durch den<br />

Zusammenbruch von Seitenwänden und durch die rückschreitende Erosion an Stufen<br />

1


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

bereitgestellt. Ebenso wird lose in der Rinne liegendes Material abtransportiert. Die im Zuge<br />

dieser Arbeit ermittelten Werte beruhen auf experimentell im Gelände erhobenen Daten.<br />

Viele Publikationen veröffentlichen dagegen Abtrags- und Abflusswerte, die mit Hilfe von<br />

Modellen erzeugt worden sind. Den Bearbeitern fehlt oftmals ein Hintergrundwissen zu den<br />

spezifischen Gebietseigenschaften, welches nur durch vor Ort erworbene Geländekenntnisse<br />

erreicht werden kann. Die Ergebnisse solcher Modellrechnungen weisen oftmals eine<br />

unbefriedigende Qualität auf. Vergleiche von Modellergebnissen mit experimentell erhobenen<br />

Daten zeigen, dass die berechneten Werte deutlich von den Messwerten abweichen. Die<br />

verwendeten Modelle basieren auf bestimmten physikalischen Grundannahmen. Aufgrund der<br />

oftmals schlechten Ergebnisqualität von Bodenerosionsmodellen werden als fünftes Ziel<br />

diese physikalischen Grundannahmen mithilfe der Spülversuche überprüft. Dabei zeigt sich<br />

deutlich, dass ganz grundlegende Probleme auftreten können. So weisen viele der während<br />

der Spülversuche entnommenen Proben eine höhere Materialmenge auf, als dies aufgrund der<br />

hydraulischen Gegebenheiten nach gängigen Modellannahmen möglich sein sollte. Auch der<br />

in Modellen angenommene lineare Zusammenhang zwischen hydraulischen Parametern und<br />

Abtragswerten ist in den Ergebnissen der Geländeexperimente nicht zu erkennen. Es stellt<br />

sich die Frage, woran dies liegen könnte. Durch die Interaktion verschiedener Prozesse ist es<br />

nicht möglich, diese mithilfe eines einzelnen Parameters sinnvoll zu beschreiben. Als<br />

problematisch erweist sich auch, dass die Berechnung einzelner hydraulischer Parameter sehr<br />

uneinheitlich gehandhabt wird: Je nach Literaturangabe werden unterschiedliche Faktoren zur<br />

Berechnung herangezogen. Grundlage der Herleitung vieler hydraulischer Parameter ist dabei<br />

die Navier-Stokes-Gleichung, deren Gültigkeit allerdings bis heute weder bewiesen noch<br />

widerlegt ist. Als Ziel Sechs der Arbeit wird geprüft, in wieweit die Methode „Spülversuch“<br />

zur Bearbeitung weiterer Fragestellungen im Bereich der Bodenerosion eingesetzt werden<br />

kann. Insgesamt zeigen die Ergebnisse der Spülversuche, wie wichtig es ist, experimentelle<br />

Daten im Gelände, also unter möglichst natürlichen Bedingungen, zu erheben. Nur auf diese<br />

Weise kann das Grundlagenwissen so weit verbessert werden, dass die Erstellung eines in der<br />

Praxis anwendbaren Bodenerosionsmodells überhaupt in den Bereich des Möglichen rückt.<br />

Dabei müssen jedoch die räumliche und zeitliche Variabilität sowie die verschiedenen<br />

Kombinationen der einzelnen Prozesse berücksichtigt werden. Die Abbildung natürlicher<br />

Prozesse mittels einfacher statischer mathematischer Formeln erscheint nach den Ergebnissen<br />

der im Rahmen dieser Arbeit durchgeführten Messungen nicht möglich.<br />

2


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Kapitel 1<br />

Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der<br />

hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Probleme kann man niemals mit derselben<br />

Denkweise lösen, durch die sie entstanden sind.<br />

Albert Einstein<br />

1.1 Einleitung<br />

Im Bereich der Bodenerosionsforschung wird im Allgemeinen zwischen flächenhafter<br />

Erosion (interrill erosion), Rillen- und Rinnenerosion (rill erosion) sowie Gully-Erosion<br />

unterschieden (z. B. Richter, 2001; Ries, 2011). Die Arbeitsgruppe der Physischen<br />

<strong>Geographie</strong> unter Leitung von Herrn Prof. Ries an der Universität Trier befasst sich seit vielen<br />

Jahren mit den Prozessen und Formen der Bodenerosion, insbesondere mit der Erforschung<br />

der räumlichen und zeitlichen Entwicklung von Gullys (z. B. Marzolff et al., 2011; Ries &<br />

Marzolff, 2003). Da diese Erosionsformen ein sehr großes Schadenspotential aufweisen<br />

(Avni, 2005), ist es nicht empfehlenswert, die zur Entwicklung der Formen notwendigen<br />

Prozesse experimentell zu aktivieren. Daher wird durch Monitoring mit Hilfe verschiedener<br />

Kameraträger wie Heißluftzeppelin, Starkwinddrachen und unbemannter Drohne das<br />

Wachstum über mehrere Jahre festgestellt und ein Abtragswert ermittelt. Neben der Erfassung<br />

des Gullywachstums wird mit experimentellen Methoden wie der Beregnungssimulation<br />

mittels mobiler Kleinberegnungsanlage das Abfluss- und Abtragsverhalten im Einzugsgebiet<br />

eines Gullys quantifiziert. Hierbei wird der flächenhafte Abfluss und Abtrag ermittelt. Bisher<br />

ist jedoch nicht berücksichtigt worden, dass in vielen Fällen das Wasser aus dem<br />

Einzugsgebiet nicht über den flächenhaften sheet flow dem Gully zugeführt wird, sondern in<br />

konzentrierter Form über Rinnensysteme. Dabei laufen deutlich andere Prozesse als in den<br />

Bereichen zwischen den Rinnen (interrill area) ab: Der Abtrag im Zuge der interrill erosion<br />

wird durch den Einschlag der Regentropfen ausgelöst und verstärkt (Beuselinck et al., 2002).<br />

Neben dem Einfluss der wesentlichen Bodeneigenschaften (Kuhn & Bryan, 2004; Kuhn et al.,<br />

2003; Le Bissonnais et al., 2005) wird angenommen, dass die interrill erosion hauptsächlich<br />

von der Niederschlagsintensität abhängt (Brodie & Rosewell, 2007; Bryan, 2000).<br />

Rill erosion dagegen wird durch den konzentrierten Abfluss von Wasser ausgelöst (Bryan,<br />

2000; Govers et al., 2007; Knapen et al., 2007) und gilt als der wichtigste Prozess der<br />

Sedimentproduktion und damit des Bodenverlustes (Cerdan et al., 2002; Poesen, 1987).<br />

3


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Allgemein wird Rinnenerosion als ein Prozess angesehen, der eintritt, wenn fließendes<br />

Wasser bestimmte, die Erodibilität betreffende Kennwerte übersteigt (Knapen et al., 2007).<br />

Die Sonderstellung linearer Erosionsprozesse gegenüber flächenhaftem Abtrag aus<br />

geomorphologischer, hydrologischer und ökonomischer Sicht ergibt sich aus der Tatsache,<br />

dass abfließendes Wasser seine maximale Wirkung sowohl als Erosions- als auch als<br />

Transportmedium erreicht, wenn es in Rinnen konzentriert wird (Merz, 1993). Abflusstiefe<br />

und Fließgeschwindigkeit erreichen hier höhere Werte als beim flächenhaften sheet flow,<br />

sodass die auf einen Bodenpartikel wirkenden Kräfte um ein Vielfaches höher sind als bei<br />

flächenhaftem Abfluss (Poesen, 1987). Dadurch steht mehr Energie für die Mobilisierung und<br />

den Transport von Bodenmaterial zur Verfügung (Merz, 1993), was zu einer deutlichen<br />

Erhöhung der Erosionsraten führt.<br />

Übergeordnetes Ziel der Bodenerosionsforschung ist es, Gegenmaßnahmen zu finden, um den<br />

Verlust von fruchtbarem Boden zu verhindern. Empfehlungen basieren häufig auf<br />

Ergebnissen von Modellberechnungen. Werden in diesen Modellen jedoch keine Rinnen<br />

berücksichtigt, führt dies zu falschen Ergebnissen und Schlussfolgerungen und endet<br />

schließlich in unwirksamen Maßnahmen zur Erosionsbekämpfung. Allerdings ist der<br />

Forschungsstand zur Rinnenerosion der Bereich der Bodenerosionsforschung, der - verglichen<br />

mit dem zum flächenhaften Abtrag - noch viele offene Fragen aufweist. Es ist noch wenig<br />

über die Unterschiede im Verhalten von Rinnen in verschiedenen Einzugsgebieten bekannt.<br />

Aus dem aktuellen Forschungsstand lassen sich verschiedene Fragestellungen ableiten,<br />

welche im Rahmen dieser Arbeit bearbeitet werden.<br />

1) Experimente zur Rinnenerosion<br />

Um Rinnenerosionsprozesse verstehen und in der Folge in Modellen auch nur annähernd<br />

korrekt abbilden zu können, sind reproduzierbare Messmethoden notwendig. Viele<br />

Untersuchungen zur Rinnenerosion haben bisher unter Laborbedingungen stattgefunden,<br />

wobei Böden verschiedener Texturen und natürliche oder künstliche Niederschläge verwendet<br />

worden sind (Brunton & Bryan, 2000; Bryan & Poesen, 1989; Gilley et al., 1990; Huang et<br />

al., 1996; Mancilla et al., 2005). Giménez & Govers (2002) zeigen jedoch, dass die meisten<br />

Daten, die aus Rinnenmodellen mit glatter Sohle und Gerinnewandung abgeleitet werden,<br />

nicht auf natürlich entwickelte Rinnen mit rauem Bett übertragen werden können.<br />

Prozessbasierte Untersuchungen müssen also in natürlichen Rinnen durchgeführt werden oder<br />

zumindest sollte eine Kombination aus prozessorientierten Laborversuchen und<br />

Geländeuntersuchungen angestrebt werden. Oftmals werden hydraulische Parameter mittels<br />

4


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Formeln berechnet, die ursprünglich für die Beschreibung des Fließverhaltens in Flüssen<br />

entwickelt worden sind. Untersuchungen von Govers (1992) und Govers et al. (2007) zeigen<br />

jedoch, dass diese Parameter nicht ohne Weiteres auf das Fließverhalten in Rinnen übertragen<br />

werden können. Wie lassen sich also unter möglichst natürlichen Bedingungen experimentelle<br />

Daten zur Rinnenerosion gewinnen?<br />

2) Rinnen als Abflusssammler<br />

Durch Rinnensysteme wird das Wasser des Einzugsgebietes schneller einem Gully zugeführt<br />

als über den sheet flow der interrill-Bereiche. Dabei erreicht das Wasser den headcut in<br />

konzentrierter Form, sodass dort verstärkte Erosionsprozesse ablaufen können. Die Frage ist<br />

also: Welcher Flächenanteil eines Einzugsgebietes wird über Rinnennetzwerke entwässert?<br />

Siehe: <strong>Wirtz</strong> et al., 2010, 2012b<br />

3) Rinnen als Sedimentquelle<br />

Das Wasser, das am headcut wirkt, transportiert meist bereits eine gewisse Sedimentmenge.<br />

Dieses Sediment stammt aus dem Einzugsgebiet, d. h., zusätzlich zur Aufzehrung des<br />

Einzugsgebietes durch den Gully führt auch die Wirkung des Wassers auf der Fläche und in<br />

den Rinnen zu Bodenabtrag. Die Frage ist in diesem Fall: Wie groß ist der Anteil der<br />

Rinnenerosion an der Gesamt-Sedimentdynamik? Siehe: <strong>Wirtz</strong> et al., 2010, 2012b<br />

4) Prozesse in der Rinne<br />

Um Maßnahmen gegen Erosionsprozesse aufzeigen zu können, müssen zunächst die Prozesse<br />

identifiziert und verstanden werden, die für den Bodenverlust verantwortlich sind. Während<br />

die Erosionsprozesse, die auf den interrill-Flächen stattfinden, relativ bekannt sind, gibt es<br />

noch Defizite beim Prozessverständnis zur Rinnenerosion. Die zu beantwortenden Fragen<br />

sind demnach: Welche Prozesse laufen in der Rinne ab? Wie sind die unterschiedlichen<br />

Prozesse anteilig vertreten? Siehe: <strong>Wirtz</strong> et al., 2012a, 2013<br />

5) Messen oder Modellieren?<br />

Da es im Bereich der angewandten Erosionsforschung nicht immer möglich ist, vor der<br />

Beurteilung des Erosionsverhaltens eines konkreten Gebietes eine Vielzahl von Experimenten<br />

durchzuführen, wird auf bestehende Bodenerosionsmodelle zurückgegriffen. Die Ergebnisse<br />

dieser Modellrechnungen stimmen jedoch oftmals nicht mit tatsächlich gemessenen Werten<br />

überein. Die Modelle weisen bestimmte Grundannahmen auf, die sich anhand von<br />

5


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Experimenten überprüfen lassen. Die Fragen sind: Stimmen diese Grundannahmen? Und, für<br />

den Fall, dass sie es nicht tun: Wo liegen die Probleme der Modellannahmen? Siehe: <strong>Wirtz</strong> et<br />

al., subm. 2012, 2013<br />

6) Weitere Anwendungen<br />

Im Zuge dieser Arbeit ist die Methode „Spülversuch“ entwickelt worden, um experimentelle<br />

Daten zu verschiedenen Bereichen der Rinnenerosionsforschung v. a. in semiariden Gebieten<br />

zu liefern. Auf dieses Ziel hin ist die Entwicklung des Versuchsaufbaus abgestimmt. Neben<br />

der eigentlichen Funktion im Rahmen der Rinnenerosionsforschung stellt sich jedoch immer<br />

die Frage: Kann eine bereits erprobte Methode auch in Bereichen anderer Fragestellungen<br />

sinnvoll eingesetzt werden? Siehe: Ries et al., subm. 2011<br />

Jede der sechs Fragestellungen wird in einem oder mehreren Artikeln behandelt. In diesem<br />

zusammenfassenden Kapitel 1 sollen die einzelnen Veröffentlichungen, die als nachfolgende<br />

Kapitel 2-7 vorliegen, in einen wissenschaftlichen Zusammenhang gebracht werden.<br />

Abbildung 1 zeigt eine grafische Darstellung der Struktur der Arbeit. Die Fragestellungen und<br />

ihre Antworten werden in den folgenden Unterkapiteln kurz vorgestellt und die<br />

Zusammenhänge zwischen den einzelnen Artikeln deutlich herausgestellt. In einem<br />

abschließenden Ausblick werden v.a. die weiterführenden Änderungen der Methoden<br />

erläutert, die sich aktuell in der Erprobungsphase befinden.<br />

Um Unklarheiten, Ungenauigkeiten oder sperrige Begriffe durch Übersetzung zu vermeiden,<br />

werden die in Tabelle 1 aufgelisteten Begriffe im englischen Originalterminus beibehalten.<br />

Sie sind im Text kursiv gedruckt.<br />

Die Schreibweise der physikalischen Einheiten erfolgt gemäß der PTB Mitteilungen (2007).<br />

6


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Abbildung 1: Struktur der Arbeit<br />

7


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Begriff<br />

Black box<br />

Bulldozer-effect<br />

Cloud<br />

Connectivity<br />

Critical shear stress<br />

Detachement capacity<br />

Detachment rate<br />

Event<br />

Feature points<br />

Goat erosion<br />

Headcut<br />

High flow experiment<br />

Interrill<br />

Knickpoint<br />

Low flow experiment<br />

Open source<br />

Plunge pool<br />

Probability density functions<br />

Rill erosion<br />

Runoff<br />

Shear stres<br />

Sheet flow<br />

Stream power<br />

Transport capacity<br />

Transport rate<br />

Unit length shear force<br />

Water repellency<br />

Tabelle1: Englische Begriffe im Text<br />

Erklärung<br />

System, von dem in geg. Kontext nur äußeres Verhalten betrachtet wird<br />

Prozess, bei dem Substrat vor der Wasserfront hergeschoben wird<br />

IT-Infrastruktur, die über ein Netzwerk zur Verfügung gestellt wird<br />

Hydraulische Verbindung zwischen z.B. Rinne und Gully<br />

Substratabhängiger Parameter, Gegenpart zum shear stress [Pa]<br />

Max. möglich aufzunehmende Materialmenge pro Zeit und Fläche<br />

Aktuell aufgenommene Materialmenge pro Zeit und Fläche<br />

Singuläres Erosionsereignis z.B. Zusammenbruch einer Seitenwand<br />

Punkte ähnlicher Eigenschaften<br />

Durch Ziegentritt ausgelöste Bodenerosion<br />

Oberes Ende eines Gullys oder einer Rinne<br />

Spülversuch mit hoher Einleitintensität (≥ 250 L/min)<br />

Fläche zwischen zwei Rinnen<br />

Stufe innerhalb eines Rinnenlängsprofils<br />

Spülversuch mit geringer Einleitintensität (9 L/min)<br />

Öffentlich zugängliche Softwareform<br />

Vertiefung unterhalb eines knickpoints oder headcuts<br />

Wahrscheinlichkeits-Dichtefunktion<br />

Zusammenfassung für Rillen- und Rinnenerosion<br />

Oberflächenabfluss<br />

Hydraulischer Parameter [Pa]<br />

Flächenhafter Abfluss, Gegenteil: konzentrierter Abfluss<br />

Hydraulischer Parameter, versch. Varianten mit versch. Einheiten<br />

Max. möglich zu transportierende Materialmenge pro Zeit<br />

Aktuell transportierte Materialmenge pro Zeit<br />

Hydraulischer Parameter [N]<br />

„Wasserabweisungsvermögen“<br />

1.2 Experimente zur Rinnenerosion<br />

Es sind bei Weitem noch nicht alle Faktoren bekannt, die Entstehung und Verhalten von<br />

Rinnen beeinflussen, und auch das Wissen um die Wechselwirkungen zwischen den einzelnen<br />

Parametern weist noch deutliche Lücken auf. Knapen et al. (2007) führen dies auf den Mangel<br />

vergleichbarer Daten zurück, die nötig wären, um eine vergleichende Analyse von<br />

verschiedenen Studien zur Rinnenerosion durchzuführen. Das Fehlen einer einheitlichen<br />

Methode führt zu großen Problemen in der Quantifizierung von Bodenerosionsprozessen und<br />

8


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

macht Messungen schwer vergleichbar (Auerswald et al., 2009; Knapen et al., 2007;<br />

Stroosnijder, 2005). Die Wichtigkeit standardisierter Methoden wird auch von Casali et al.<br />

(2006) aufgezeigt. Sie testen drei häufig verwendete Methoden, um die Intensität von<br />

Rinnenerosionsprozessen quantifizieren zu können und zeigen dabei, dass verschiedene<br />

Methoden stark abweichende Ergebnisse liefern können. Abhängig von Methode, Form der<br />

untersuchten Rinne, Position und Distanz zwischen den aufgenommenen Querschnitten liegt<br />

der Messfehler bei ≤ 40 %. Daher ist ein wichtiges Ziel dieser Arbeit, eine reproduzierbare<br />

Methode zur Quantifizierung von Rinnenerosionsprozessen zu entwickeln. Im Verlauf der<br />

Arbeit wird die Methode des Spülversuches aufgrund von praktischen Geländeerfahrungen<br />

immer wieder verbessert und gezielt erweitert. Die verschiedenen Versionen des<br />

Versuchsaufbaus sind jeweils in den zugehörigen Artikeln, die in den Kapiteln 2 bis 7<br />

vorliegen, beschrieben. An dieser Stelle soll die Entwicklung in ihrer Gesamtheit dargestellt<br />

werden.<br />

Die konstanten methodischen Parameter betreffen die Aufteilung eines Spülversuches in zwei<br />

Durchgänge, um den Einfluss einer erhöhten Bodenfeuchte berücksichtigen zu können. Der<br />

Einfluss der Ausgangsbodenfeuchte auf den Sedimenttransport wird von Govers et al. (1990)<br />

und Govers (1991) beschrieben. In ihren Experimenten unter trockenen Bedingungen erreicht<br />

die Sedimentkonzentration nahezu die transport capacity des Abflusses, während unter<br />

höherer Ausgangsbodenfeuchte die Sedimentkonzentrationen geringer ausfallen: Höhere<br />

Bodenfeuchte reduziert die Infiltrationsrate, was zu einer höheren Wassermenge führt, die<br />

durch den zu untersuchenden Rinnenabschnitt fließt. Gleichzeitig wird jedoch nicht<br />

entscheidend mehr Material transportiert, die Sedimentkonzentration sinkt. Dabei ist zu<br />

differenzieren, ob eine Befeuchtung und damit eine erhöhte Kohäsion oder eine<br />

Wassersättigung und damit eine Verringerung des Zusammenhaltes zwischen den einzelnen<br />

Bodenteilchen vorliegt. Eine höhere Bodenfeuchte an der Oberfläche reduziert zudem die<br />

water repellency des Bodens. Dies kann für die Abflussdynamik wichtig sein, besonders unter<br />

sonst trockenen Bedingungen.<br />

Des Weiteren werden in allen Versionen der Spülversuche pro Rinne drei Messstellen<br />

gewählt, an denen jeweils vier Proben nach dem Zeitschema 0, 30, 90, 150 Sekunden nach<br />

Eintreffen der Wasserfront entnommen werden. Auch die Sohlenneigung wird in allen<br />

Versionen aufgenommen, die Länge der Segmente wechselt jedoch: Je nach verfügbarem<br />

Messinstrument weisen die Segmente entweder eine Länge von 1 m oder von 2,5 m auf.<br />

Stufen in der Rinne werden dabei nicht aufgenommen, ihre Lage und Höhe wird getrennt<br />

vermerkt. Die Scherfestigkeit des Substrates wird mit Hilfe eines Flügelschergerätes zwischen<br />

9


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

der letzten Probennahmestelle und dem Ende des untersuchten Rinnenabschnittes bestimmt.<br />

Unterschiede betreffen die verwendete Pumpe, die eingeleitete Wassermenge, die<br />

Einleitintensität, die Erfassung des Wasserstandes an den Messstellen, den Einleitzeitpunkt<br />

von Farbtracern (um Veränderungen in der Fließgeschwindigkeit festzustellen) und die<br />

Erfassung des Rinnenquerschnittes.<br />

In der ersten im Rahmen dieser Arbeit zum Einsatz gekommenen Version der<br />

Versuchsanordnung (V1) des Spülversuchs wird mithilfe einer benzingetriebenen<br />

Motorpumpe (Fitosa-Spritze mit Bertolini-Pumpe), wie sie auch in<br />

Kleinberegnungsversuchen Verwendung findet, eine konstante Einleitintensität von 9 L min -1<br />

über acht Minuten aufrechterhalten (insgesamt also 72 L). Eine Mobilisierung von Material an<br />

der Einleitstelle wird durch eine spezielle Abbremsplatte verhindert. Die Bewertung der<br />

Abflussintensität erfolgt durch Messung des Abflusses nach einer vorher festgelegten<br />

Fließstrecke mittels Drucksensoren. Die Veränderung der Fließgeschwindigkeit wird mittels<br />

Farbtracern bestimmt. Allerdings können mit der Versuchsanordnung dieser Version noch<br />

keine Informationen zu hydraulischen Parametern, also Werten, die sich aus<br />

Rinnenquerschnitt und Wasserstand ableiten ließen, aufgenommen werden.<br />

Mit Einführung von V2 erfolgt an den drei Messstellen nun auch eine Erfassung des<br />

Rinnenquerschnittes mit Hilfe von Metallstäben. Diese werden durch Löcher in einer mobilen<br />

Messbrücke bis zur Rinnensohle geführt. Der Abstand zwischen Messbrückenunterkante auf<br />

Bodenniveau und Stabende auf Sohlenniveau wird gemessen und dadurch ein Rinnenprofil<br />

am Messpunkt erstellt. Der Wasserstand kann aufgrund von problembehafteter Benutzung der<br />

verwendeten Zollstöcke zunächst nur geschätzt werden: Durch das Eintauchen des Zollstocks<br />

entsteht ein Wasserwirbel, der das Ablesen erschwert, des Weiteren ist es nicht möglich,<br />

immer an derselben Position in der Rinne zu messen.<br />

Dieses Problem kann jedoch durch Aufstellen einer Messbrücke in V3 gelöst werden. Diese<br />

dient als fester vertikaler Referenzpunkt. So ist sichergestellt, dass immer an derselben Stelle<br />

der Rinne der Abstand zwischen Messbrücke und Wasseroberfläche mit einem Zollstock<br />

gemessen wird. Bei bekanntem Abstand zwischen Rinnensohle und Brücke lässt sich der<br />

Wasserstand berechnen ohne dass Verwirbelungen durch das Eintauchen des Zollstocks<br />

entstehen. Die bis zu dieser Version verwendete Pumpe ist in ihrer Leistung begrenzt, sodass<br />

ein realistischer Abfluss in Rinnen mit einem größeren Einzugsgebiet damit nicht zu<br />

simulieren ist.<br />

Als wichtige Neuerung wird in V4 die Fitosa-Spritze durch eine weit leistungsfähigere GMP-<br />

150-Pumpe der Firma Güde ersetzt, sodass nun pro Durchgang eine Wassermenge von<br />

10


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

1000 L in die Rinnen eingeleitet werden kann. Die Intensität schwankt je nach Anordnung<br />

von Wassertank, Pumpe und Einleitstelle im Hang zwischen 250 und 330 L min -1 . Die<br />

Einführung der leistungsfähigeren Pumpe macht es zudem erforderlich, die Einleitzeiten der<br />

Farbtracer von 3 min und 6 min auf 1 min und 2 min vorzuziehen sowie den Durchmesser des<br />

Abflussmessrohrs zu vergrößern. Ebenso wird eine veränderte Version der Abbremsplatte bei<br />

der Wassereinleitung eingesetzt. Die jetzt verwendete Pumpe stellt eine höhere Wassermenge<br />

bei geringerem Wasserdruck bereit, die Abbremsplatte muss diesen neuen Anforderungen<br />

angepasst werden. Die Erfassung des Rinnenquerschnittes kann durch das Ersetzen der<br />

Metallstangen durch den digitalen Laser-Entfernungsmesser Bosch PLR 30 genauer und<br />

schneller gestaltet werden. Die Zollstöcke zur Wasserstandsmessung werden in dieser Version<br />

durch Ultraschallsensoren ersetzt.<br />

Die Entwicklung der Spülversuche von V1 bis V4 ist in Tabelle 2 zusammengefasst und in<br />

Abbildung 2 grafisch dargestellt.<br />

Tabelle 2: Entwicklung der Spülversuche: Nicht erfasste Werte sind durch „X“ gekennzeichnet. In der<br />

Zeile „Artikel“ gilt, soweit kein anderer Autor angegeben ist, der Vorsatz „<strong>Wirtz</strong> et al.“<br />

Version V1 V2 V3 V4<br />

Pumpe Fitosa Fitosa Fitosa Güde<br />

Intensität [L min -1 ] 9 9 9 250-330<br />

Menge [L] 72 72 72 1000<br />

Strecke bis Abflussmessung [m] 15-25 15-25 15-25 15-25<br />

Wasserstand X X Zollstock Ultraschall<br />

Farbtracer [min] 3 und 6 3 und 6 3 und 6 1 und 2<br />

Querschnitt X Metallstäbe Metallstäbe Laser<br />

Artikel 2010 2012a<br />

2012b<br />

2013 subm. 2012<br />

2012b<br />

Ries et al.,<br />

subm. 2011<br />

11


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Abbildung 2: Veränderungen in der Methode „Spülversuch“. Zu beachten: Es sind nur maßgebliche<br />

Veränderungen dargestellt, die die verwendeten Gerätschaften betreffen. Veränderungen der<br />

Messmethodik (z.B. Veränderung der Einleitzeiten der Farbtracer) sind hier nicht abgebildet.<br />

1.3 Rinnen als Abflusssammler<br />

Ein wichtiger Punkt im Rahmen der Rinnenerosionsforschung ist die Frage nach der<br />

Effektivität linearer Erosionsformen bezüglich der Konzentration des Oberflächenabflusses.<br />

Über Rinnen kann Wasser schneller z. B. einem Gully zugeführt werden, dadurch gelangt<br />

12


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

mehr Wasser gleichzeitig an einen bestimmten Punkt des headcuts und kann dort konzentriert<br />

rückschreitende Erosion leisten. Allerdings ist ein Einzugsgebiet niemals vollständig mit<br />

Rinnen bedeckt, sondern die Rinnennetzwerke werden durch unterschiedlich große interrill-<br />

Bereiche getrennt. Die Frage ist also, wie effektiv das Wasser durch den geringen<br />

Rinnenanteil eines Gebietes ab- und dem Gully zugeführt werden kann. Dies wird im Rahmen<br />

dieser Arbeit in zwei Artikeln diskutiert. In dem Gully-Einzugsgebiet Freila in Andalusien<br />

vergleichen <strong>Wirtz</strong> et al. (2012b) die Anteile von Rinnenerosion und flächenhaftem Abtrag an<br />

der Gesamterosion. Das etwa 10 ha große Einzugsgebiet besteht zu 0,25 % aus Rinnen, die<br />

aber mit den an die Rinnen angeschlossenen Flächen 20 % dieses Gebietes entwässern. Die<br />

runoff-Intensitäten liegen in den interrill-Bereichen um den Faktor 5 höher als in den Rinnen.<br />

Dies ist jedoch auf die größere interrill-Gesamtfläche zurückzuführen. Wird die runoff-<br />

Intensität pro Quadratmeter betrachtet, liegen die Werte in den Rinnen um den Faktor 50-60<br />

höher als auf den interrill-Bereichen. Im Arnástal in den spanischen Pyrenäen vergleichen<br />

<strong>Wirtz</strong> et al. (2010) gemessene Abflusswerte aus Rinnen mit Abflusswerten, die Seeger et al.<br />

(2005) im Rio Arnás gemessen haben. Der Arnás erreicht Werte zwischen 48 und 708 L s -1<br />

mit Abflussspitzen von bis zu 2.347 L s -1 . Bezogen auf das Gesamteinzugsgebiet von 284 ha<br />

bedeutet dies einen spezifischen Abfluss zwischen 1,7x10 -5 und 2,5x10 -4 L s -1 m -2 . In den<br />

untersuchten Rinnen werden spezifische Abflüsse von 7x10 -5 und 8x10 -5 L s -1 m -2 erreicht.<br />

Die Werte in den Rinnen liegen somit innerhalb der Spanne der im Rio Arnás gemessenen<br />

Werte. Dabei ist zu beachten, dass die experimentell verwendete Wassermenge als sehr gering<br />

eingestuft werden muss. Beregnungsversuche in den Rinneneinzugsgebieten zeigen, dass bei<br />

realen Niederschlägen ein bis 30-fach höherer Abfluss in den Rinnen erwartet werden kann<br />

(<strong>Wirtz</strong> et al., 2010). Es würden also in den Rinnen deutlich höhere spezifische runoff-<br />

Intensitäten auftreten als im Vorfluter Arnás. Damit zeigt sich die hohe Effektivität der<br />

Rinnen als Abflusssammler. Allerdings muss bedacht werden, dass die Rinnennetzwerke<br />

nicht zwangsläufig den Gully oder den Vorfluter erreichen. Es ist durchaus möglich, dass die<br />

Rinne ausläuft und der Abfluss wieder flächenhaft weitergeleitet wird (connectivity). Neben<br />

der schnellen und effektiven Zuführung zu einem Gully oder einem Vorfluter bewirkt der<br />

konzentrierte Abfluss auch eine erhöhte Erosionsleistung. Das bedeutet, der zweite wichtige<br />

Punkt bezüglich der Rinneneffektivität bezieht sich auf die Wirksamkeit als Sedimentquelle.<br />

1.4 Rinnen als Sedimentquelle<br />

In einem Einzugsgebiet kann Abtrag sowohl auf den interrill-Bereichen als auch in den<br />

Rinnen stattfinden. Die Frage dabei ist: Welcher Bereich liefert in einem Einzugsgebiet mehr<br />

13


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Sediment? Die wenigen Rinnen oder der große interrill-Bereich? Bezüglich der Effektivität<br />

von Rinnen finden sich in der Literatur verschiedene Größenordnungen, gemeinsam ist ihnen<br />

jedoch, dass es sich meist um sehr hohe Werte handelt. Die Spanne reicht von<br />

„Verdreifachung“ der Erosionsraten nach der Entwicklung von Rinnen (Meyer et al., 1975)<br />

über „9-21-mal mehr“ als flächenhafter Abtrag (Morgan et al., 1987) bis hin zu „90 %“ des<br />

Bodenverlustes nach einzelnen Niederschlägen (Cerdan et al., 2002). Gemessene<br />

Sedimentkonzentrationen reichen von 10-130 g L -1 (Polyakov & Nearing, 2003) bis 210 g L -1<br />

(Loch, 2000). Gullys als größere Erosionsform weisen eine noch höhere Kapazität als<br />

Sedimentquelle auf. Allerdings sehen Faust & Schmidt (2009) aufgrund der seltenen Aktivität<br />

der Gullys die geomorphologische Effektivität dieser Großformen als eher gering an. Sie<br />

weisen den Gullys eine Aktivitätsfrequenz von 1/20 a -1 , den Rinnen 4 a -1 zu. Die im Rahmen<br />

dieser Arbeit durchgeführten Spülversuche decken eine Sedimentkonzentrationsspanne von<br />

0 g L -1 (<strong>Wirtz</strong> et al., 2010) bis 438 g L -1 (<strong>Wirtz</strong> et al., 2012a) ab. Dabei ist zu beachten, dass<br />

immer unterschiedliche Ausgangssituationen gegeben sind. Bei der Einordnung dieser Werte<br />

ist zu berücksichtigen, dass in einem Einzugsgebiet in den Pyrenäen, in dem sich die Rinnen<br />

im anstehenden Gestein gebildet haben bzw. unter starker Mitwirkung anderer Prozesse (wie<br />

Frostsprengung), die Werte deutlich geringer ausfallen als in einem schluffig-lehmigen<br />

Substrat ohne Vegetationsbedeckung in Andalusien. Daher ist die wichtigere Frage: Wie<br />

verhält sich die Rinnenerosion im Vergleich zur flächenhaften Erosion in einem bestimmten<br />

Testgebiet? Die Daten bezüglich des flächenhaften Abtrags können mittels<br />

Kleinberegnungsanlagen ermittelt werden. Das größere Problem besteht darin, die gegebenen<br />

Punktdaten sinnvoll und wissenschaftlich korrekt auf eine Fläche wie z. B. ein ganzes<br />

Einzugsgebiet zu extrapolieren. In <strong>Wirtz</strong> et al. (2010) wird im Arnástal in den spanischen<br />

Pyrenäen eine maximale Sedimentkonzentration von 6,3 g L -1 gemessen. Am Gebietsauslass<br />

des Arnástals ermitteln Seeger et al. (2005) maximale Sedimentkonzentrationen von 4,8 g L -1 ,<br />

ähnliche Werte zeigen auch Beregnungsversuche (Seeger, 2007). Dies deutet zunächst darauf<br />

hin, dass die Rinnen keine deutlich höheren Abtragswerte als andere Prozesse liefern und<br />

damit auch keinen auffallend hohen Anteil am Sedimentaustrag des Gebietes aufweisen.<br />

Dabei gilt es jedoch, folgende Punkte zu bedenken: Die Beregnungsversuche werden mit<br />

einer Intensität von 40 mm h -1 durchgeführt, die Abflussintensität in den Spülversuchen liegt<br />

bei 9 L min -1 . Wird die Größe der Rinneneinzugsgebiete und deren runoff-Koeffizienten<br />

berücksichtigt, so ist klar zu erkennen, dass der Abfluss in den Rinnen unter einem solchen<br />

Niederschlagsereignis bis zu 30-mal höher ausfällt als der simulierte. Daraus folgt, dass die<br />

Rinnen trotz der schlechten Voraussetzungen (anstehendes Gestein, bestenfalls grob<br />

14


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

verwittert, Pflanzenwachstum im wenigen feineren Material in den Rinnen, viel zu wenig<br />

Wasser) das wenige Wasser aus den Experimenten sehr effektiv für den Abtrag nutzen<br />

können. Vermutlich ist das Material, das am Gebietsauslass gemessen wird, nicht allein durch<br />

den Oberflächenabfluss unter Niederschlagsereignissen am Messpunkt angekommen, sondern<br />

auch durch Massenbewegungen und Erosion im Flussbett selbst.<br />

In dem Gully-Einzugsgebiet Freila in Andalusien (<strong>Wirtz</strong> et al., 2012b) werden verschiedene<br />

Methoden (Beregnungen, Spülversuche, Luftbildaufnahmen, Kartierung) kombiniert, um die<br />

Anteile des flächenhaften und des linienhaften Abtrages in diesem Gesamtgebiet zu<br />

vergleichen. Die Gesamtmenge des abgetragenen Materials ist auf den interrill-Bereichen 5-<br />

15-mal höher als in den Rinnen. Dies ist nicht verwunderlich, da der Flächenanteil der<br />

interrill-Fläche 99,75 %, der Anteil der Rinnen nur 0,25 % beträgt. Werden jedoch die<br />

abgetragenen Mengen auf die Flächenanteile bezogen, zeigen die Rinnen eine 20-60-fach<br />

höhere Rate als die interrill-Flächen. In anderen Testgebieten in Andalusien können noch<br />

deutlich höhere Werte in Rinnen gemessen werden, allerdings fehlen bisher die Daten, um<br />

auch für diese Gebiete eine Hochrechnung auf das Gesamtgebiet durchführen zu können. Die<br />

Versuche zeigen eindeutig, dass Rinnen effektive Sedimentquellen sind, selbst unter für<br />

Erosionsprozesse erschwerten Bedingungen. Während Abflussereignissen laufen in Rinnen<br />

andere Prozesse als auf den interrill-Flächen ab und/oder ähnliche Prozesse laufen mit einer<br />

völlig anderen Intensität ab. Im Rahmen dieser Arbeit wird geprüft, welche Prozesse in den<br />

Rinnen ablaufen.<br />

1.5 Prozesse in der Rinne<br />

Es ist also zu klären, welche Prozesse die Rinnen zu einer solch effektiven Sedimentquelle<br />

machen. In Andalusien und in den Bardenas Reales (<strong>Wirtz</strong> et al., 2012a, 2013) wird versucht,<br />

die Prozesse anhand der Werte der einzelnen Proben und deren zeitliche Anordnung an den<br />

Messstellen zu identifizieren. Insgesamt lassen sich anhand der Ergebnisse der Spülversuche<br />

und der Beobachtungen während der Experimente mehrere parallel ablaufende Prozesse<br />

ausweisen. Zunächst wird das lose in der Rinne liegende Material aufgenommen und<br />

abtransportiert. Dies ist an der oft hohen Sedimentkonzentration der relativ langsam<br />

fließenden Wasserfront zu erkennen. Je nach Materialmenge und Abflusswirksamkeit<br />

geschieht dies direkt an der Wasserfront, oder der Prozess zieht sich über einen gewissen<br />

Zeitraum von bis zu mehreren Minuten hin, sodass die Auswirkungen in mehreren Proben an<br />

hohen Sedimentkonzentrationen zu erkennen sind. Dieser Prozess wird als bulldozer-effect<br />

bezeichnet (Seeger et al., 2004; Regüés et al., 2000), da ein großer Teil des losen Materials<br />

15


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

vor der Wasserfront hergeschoben wird. Ist dies der einzige Prozess der abläuft, fällt die<br />

Sedimentkonzentration von der ersten zur letzten Probe kontinuierlich ab. Im zweiten<br />

Durchgang wiederholt sich der Vorgang auf einem etwas geringeren Niveau, da nach<br />

Beendigung des ersten Durchgangs wieder eine gewisse Menge Material in der Rinne<br />

verbleibt. Der idealtypische Verlauf dieses Prozesses ist in Abbildung 3 dargestellt.<br />

Neben dem Abtransport losen Materials kann sich das Wasser durch Tiefenerosion in die<br />

Rinnensohle einschneiden, d. h. die Rinnensohle wird eingetieft. Dieser Prozess wird<br />

hauptsächlich über die hydraulischen Parameter wie shear stress gesteuert. Allerdings scheint<br />

es, dass während der Experimente nur in Ausnahmefällen die verwendete Wassermenge und<br />

Einleitintensität ausreichen, diesen Prozess auf der gesamten Rinnenlänge zu aktivieren. Dies<br />

wäre an einer gleichbleibenden oder sogar ansteigenden Sedimentkonzentration an einzelnen<br />

Messstellen zu erkennen. Im zweiten Durchgang verliefe die Sedimentkonzentration auf<br />

einem höheren Niveau, da das bereits vollständig wassergesättigte Material leichter zu<br />

erodieren wäre. Der idealtypische Verlauf dieses Prozesses ist in Abbildung 4 dargestellt.<br />

Des Weiteren treten Prozesse auf, die nur indirekt durch den Einfluss des fließenden Wassers<br />

kontrolliert werden. Der Abfluss unterschneidet sowohl die Seitenwände der Rinne als auch in<br />

Form von rückschreitender Erosion die Stufen (knickpoints) innerhalb des Rinnenbettes.<br />

Dabei werden noch vergleichsweise geringe Sedimentmengen mobilisiert. Der höhere Anteil<br />

an Material gelangt erst durch das Nachbrechen der unterspülten Wand- und Stufenbereiche<br />

gravitativ in die Rinne und wird, wie das vor Beginn des Experimentes in der Rinne<br />

vorhandene lose Material, vom fließenden Wasser aufgenommen und abtransportiert. Auch<br />

die plunge pool-Dynamik kann innerhalb kurzer Zeit relativ große Sedimentmengen<br />

bereitstellen: Zunächst wird die Senke unterhalb einer Stufe mit Sediment angefüllt. Sobald<br />

der pool jedoch gefüllt ist, kann das darüber fließende Wasser diesen innerhalb kurzer Zeit<br />

wieder freispülen, sodass das über einen längeren Zeitraum dort gespeicherte Material sehr<br />

schnell abtransportiert wird. Erkennbar sind diese events an deutlich hervorstechenden<br />

Sedimentkonzentrationen im Probenverlauf einer Messstelle. Der idealtypische Verlauf dieses<br />

Prozesses ist in Abbildung 5 dargestellt.<br />

16


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Abbildung 3: Idealtypischer Verlauf der Sedimentkonzentration bei Transport von losem Material: a =<br />

erster Durchgang, b = zweiter Durchgang<br />

Abbildung 4: Idealtypischer Verlauf der Sedimentkonzentration bei Einschneiden in die Rinnensohle:<br />

a = erster Durchgang, b = zweiter Durchgang<br />

17


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Abbildung 5: Idealtypischer Verlauf der Sedimentkonzentration bei Massenbewegungen an Stufen und<br />

Seitenwänden (durch Oval markiert): a = erster Durchgang, b = zweiter Durchgang<br />

Der idealtypische Verlauf wird durch Überlappung verschiedener Prozesse verwischt, sodass<br />

die direkten Beobachtungen der ablaufenden Prozesse in der Rinne während des Versuches<br />

wichtige Hinweise für die Interpretation liefern. Die verschiedenen Prozesse weisen eine hohe<br />

räumliche und zeitliche Variabilität auf, sodass eine allgemeingültige Aussage bezüglich der<br />

relativen Anteile nur sehr schwer zu treffen ist. In Untersuchungen von Govers (1987) liegt<br />

der Anteil der „Massenbewegungen“ an den Seitenwänden bei 37 %, an den Stufen und<br />

knickpoints bei 12 %. Das Einschneiden in die Rinnensohle kann nur während dreier extremer<br />

Ereignisse mit Spitzenabflüssen zwischen 70 und 90 L s -1 beobachtet werden, liefert dann<br />

aber den höchsten Anteil am Gesamtaustrag (Govers, 1987). Meist ist der Abfluss jedoch zu<br />

gering, um den Prozess des Einschneidens zu aktivieren. In den vorliegenden Untersuchungen<br />

muss zur Bestimmung des Anteils der einzelnen Prozesse ein anderer Weg beschritten<br />

werden. Eine wichtige Annahme in verschiedenen Bodenerosionsmodellen ist, dass die<br />

transport rate die transport capacity nicht überschreiten kann, da in diesem Falle<br />

Sedimentation einsetzt (Scherer, 2008). Die Modelle berücksichtigen jedoch nur den durch<br />

die hydraulischen Parameter gesteuerten Prozess des Einschneidens in die Rinnensohle. Das<br />

bedeutet, dass zumindest die Differenz zwischen transport rate und transport capacity durch<br />

andere, nicht durch hydraulische Parameter gesteuerte Prozesse bereitgestellt wird. In <strong>Wirtz</strong> et<br />

al. (subm. 2012) liegt in 82 von 144 Proben die transport rate oberhalb der transport<br />

18


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

capacity. Im Mittel wird die transport capacity um 41,5 % überschritten, d. h. mindestens<br />

41,5 % des transportierten Materials wird von Prozessen bereitgestellt, die nicht hydraulisch<br />

kontrolliert sind, also Transport von losem Material, rückschreitende Erosion an Stufen,<br />

Überlaufen von plunge pools und Zusammenbrüche an Seitenwänden. In Wirklichkeit wird<br />

der Anteil höher sein, da meist die transport capacity nicht allein durch das Einschneiden<br />

ausgeschöpft wird, insbesondere bei den in den vorliegenden Versuchen verwendeten<br />

Wassermengen und Intensitäten. Der Anteil der nicht hydraulisch kontrollierten Prozesse ist<br />

jedoch nicht gleichmäßig über die Versuche verteilt. In den drei Versuchen mit den geringsten<br />

Sedimentkonzentrationen liegt der Mittelwert nur bei 24,3 % nicht hydraulisch kontrollierter<br />

Prozesse, während in den drei Versuchen mit den höchsten Sedimentkonzentrationen ein Wert<br />

von 58,7 % ermittelt wird. Das wiederum bedeutet, dass der effektivste Prozess eben nicht die<br />

hydraulisch kontrollierte Erosion in die Rinnensohle sein kann, sondern die gravitativ<br />

gesteuerten Prozesse. Diese werden jedoch in gegebenen Bodenerosionsmodellen nicht<br />

berücksichtigt, die Bodenerosion wird hier als rein von hydraulischen Parametern gesteuerter<br />

Prozess angesehen. Daher erscheint es sinnvoll, einige Grundannahmen aus gängigen<br />

Bodenerosionsmodellen genauer zu betrachten und zu bewerten.<br />

1.6 Messen oder Modellieren?<br />

Die erste Modellannahme, dass die transport rate die transport capacity nicht übersteigen<br />

kann, wird in drei Artikeln überprüft (<strong>Wirtz</strong> et al., 2012a, subm. 2012, 2013). Es zeigt sich,<br />

dass in 75 % der Fälle (<strong>Wirtz</strong> et al., 2013) bzw. in 59 % der Fälle (<strong>Wirtz</strong> et al., subm. 2012)<br />

die transport rate die transport capacity um einen Faktor > 500 übersteigen kann. In <strong>Wirtz</strong> et<br />

al. (2012a) ist zudem deutlich zu erkennen, dass besonders in den Fällen, in denen hohe<br />

Sedimentkonzentrationen gemessen werden, die transport rate die transport capacity<br />

übersteigt. In diesem Fall ist der Quotient aus transport rate und transport capacity > 1 (siehe<br />

Abbildung 6 und 7).<br />

19


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Abbildung 6: Quotient aus transport rate und transport capacity (<strong>Wirtz</strong> et al., 2013)<br />

Abbildung 7: Quotient aus transport rate und transport capacity (<strong>Wirtz</strong> et al., subm. 2012)<br />

In deterministischen Bodenerosionsmodellen werden die Erosionsprozesse als rein<br />

hydraulisch kontrolliert angesehen. Sie gehen von einem linearen Zusammenhang zwischen<br />

einem Abtragswert und einem hydraulischen Parameter aus. Abtragswerte sind z. B.<br />

20


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Sedimentkonzentration, transport rate oder detachment rate, hydraulische Parameter shear<br />

stress, unit length shear force oder verschiedene Varianten der stream power. Knapen et al.<br />

(2007) berechnen in ihrer Studie die Korrelationen zwischen verschiedenen hydraulischen<br />

Parametern und der detachment capacity verschiedener WEPP-Datensätze (Water Erosion<br />

Prediction Project). Die beste Korrelation zur detachment capacity weist die stream power mit<br />

einem mittleren Bestimmtheitsmaß von R² = 0,59 auf. Im WEPP-Modell wird der shear stress<br />

verwendet, der allerdings in keinem der getesteten Datensätze einen hohen R²-Wert aufweist.<br />

In <strong>Wirtz</strong> et al. (subm. 2012) werden an eigenen Daten die Korrelationen zwischen der<br />

detachment rate und verschiedenen hydraulischen Parametern überprüft. Dabei wird die<br />

gesamte Breite von 0 ≤ R² ≤ 0,99 abgedeckt, Trendlinien sind steigend, fallend oder auch<br />

nahezu konstant, es ist letztendlich nicht möglich, irgendeine Regelmäßigkeit zu<br />

identifizieren. Nur 40 von 252 Korrelationen (16 %) zeigen eine steigende Trendlinie mit<br />

einem Bestimmtheitsmaß von R² ≥ 0,7. Auch hier ist wieder deutlich zu erkennen, dass die<br />

Versuche mit geringen Sedimentkonzentrationen höhere R²-Werte aufweisen als die Versuche<br />

mit hohen Sedimentkonzentrationen, ganz gleich ob alle Korrelationen oder nur diejenigen<br />

mit steigender Trendlinie berücksichtigt werden.<br />

In <strong>Wirtz</strong> et al. (2013) werden auf einer homogenen Ackerfläche in vier Rinnen sowohl die<br />

Eingangsparameter für die Berechnung des shear stress als auch Erosionswerte direkt<br />

gemessen, d. h. es stehen sowohl die Eingangsdaten als auch die „korrekten“ Ergebnisse eines<br />

theoretischen Modelldurchlaufes zur Verfügung. Die Eingangsdaten für die Berechnung des<br />

shear stress und damit auch die daraus resultierenden shear stress Werte weisen für die vier<br />

beprobten Rinnen relativ geringe Schwankungen (relative Messfehler und empirische<br />

Variationskoeffizienten) auf. Träfe die Modellannahme „linearer Zusammenhang zwischen<br />

hydraulischem Parameter und Erosionsparameter“ zu, müssten auch die gemessenen<br />

Erosionswerte eine geringe Schwankung aufweisen, bei gleichen Eingangsdaten liefert ein<br />

deterministisches Modell auch immer dieselben Ergebnisse. Die Erosionsparameter weisen<br />

jedoch hohe relative Messfehler (Abbildung 8) und empirische Variationskoeffizienten<br />

(Abbildung 9) auf.<br />

21


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Abbildung 8: Relativer Messfehler verschiedener Parameter. Shear stress 1 berücksichtigt in der<br />

Probendichte Sedimentkonzentration und Korndichte, shear stress 2 nimmt einen konstanten Wert von<br />

1 an (<strong>Wirtz</strong> et al. 2013)<br />

Abbildung 9: Empirischer Variationskoeffizient verschiedener Parameter. Shear stress 1<br />

berücksichtigt in der Probendichte Sedimentkonzentration und Korndichte, shear stress 2 nimmt einen<br />

konstanten Wert von 1 an (<strong>Wirtz</strong> et al. 2013)<br />

22


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Für das Nicht-Zutreffen der Modellannahmen können mehrere Gründe aufgeführt werden:<br />

Zum einen stellt sich die Vielzahl unterschiedlicher Prozesse mit großer räumlicher und<br />

zeitlicher Variabilität als kaum zu überwindendes Problem heraus. Der Ansatz, diese Prozesse<br />

mithilfe eines einzelnen hydraulischen Parameters abzubilden, erweist sich als unbrauchbar.<br />

Zum anderen liegt ein weiteres Problem in der Verwendung des hydraulischen Parameter an<br />

sich: In den meisten Modellen wird der shear stress als hydraulischer Parameter verwendet,<br />

aus diesem lassen sich andere Parameter wie unit length shear force und verschiedene stream<br />

power-Varianten ableiten. In einem Review-Teil in <strong>Wirtz</strong> et al. (subm. 2012) und <strong>Wirtz</strong> et al.<br />

(2013) wird gezeigt, dass die Berechnung von shear stress nicht eindeutig ist. Je nach<br />

Literaturangabe werden ganz unterschiedliche Faktoren verwendet, um den shear stress zu<br />

berechnen. Zudem wird nicht deutlich zwischen shear stress - einem hydraulischen Parameter<br />

- und critical shear stress - einem Substratparameter - unterschieden. Auch die Berechnung<br />

der transport capacity ändert sich je nach Quelle. Als Folge dieser wechselnden<br />

Eingangsparameter ergibt sich, dass der shear stress nicht in allen Fällen in Pascal und die<br />

transport capacity nicht immer in Kilogramm pro Sekunde angegeben ist.<br />

Basis der shear stress-Formel in der Hydraulik ist die Navier-Stokes-Gleichung. Diese<br />

beschreibt die Bewegung von Flüssigkeiten, wobei Newtons zweites Gesetz, das<br />

Aktionsprinzip, auf Flüssigkeiten angewendet wird: Die Änderung der Bewegung einer Masse<br />

ist zur Einwirkung der bewegenden Kraft proportional und geschieht nach der Richtung<br />

derjenigen geraden Linie, nach welcher jene Kraft wirkt. Kombiniert wird dies mit der<br />

Annahme, dass die Flüssigkeitsspannung als Summe aus einem Viskositätsterm und einem<br />

Druckterm darzustellen ist. Durch Verwendung der Navier-Stokes-Gleichung kann ein<br />

inkompressibles Fluid komplett beschrieben werden, hydrodynamische Fragen werden auf ein<br />

mathematisches Problem reduziert. Dieses Problem besteht jedoch aus einem System von<br />

nichtlinearen partiellen Differentialgleichungen zweiter Ordnung. Nur die einfachsten Fälle<br />

können mit den leistungsfähigsten Computern numerisch gelöst werden. Für den allgemeinen<br />

drei-dimensionalen Fall sind Existenz, Eindeutigkeit und Richtigkeit noch nicht bewiesen.<br />

Die Dringlichkeit dieses Problems spiegelt sich in der Tatsache wider, dass der Beweis oder<br />

ein Gegenbeweis der Navier-Stokes-Gleichung als eines der sieben wichtigsten Probleme der<br />

Mathematik angesehen wird, auf dessen Lösung vom Clay Mathematics Insitute eine<br />

Belohnung von 1 Mio. $ ausgesetzt ist (Constantin, 2001; Fefferman, 2006; Schneider, 2008;<br />

Seiler, 2002; Temam, 2000; Wiegner, 1999).<br />

Ein weiteres Problem ist, dass der Einfluss von Turbulenzen in Modellen nicht berücksichtigt<br />

wird. Nearing & Parker (1994) untersuchen den Einfluss der Turbulenz auf den shear stress.<br />

23


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Sie zeigen, dass unter turbulenten Fließbedingungen dieselben shear stress Werte deutlich<br />

höhere Abtragsraten verursachen. Die Unterschiede zwischen den Abtragsraten bei<br />

turbulentem und laminarem Fließen vergrößern sich mit ansteigenden shear stress-Werten.<br />

Das bedeutet, wenn die gegebenen hydraulischen Bedingungen hohe shear stress-Werte<br />

verursachen, ist der Einfluss der Turbulenz auf die Bodenerosion höher als in Bereichen mit<br />

niedrigeren shear stress-Werten. Dies kann auch der Grund sein, warum<br />

Bodenerosionsmodelle geringe Bodenabträge überschätzen und große Bodenverluste<br />

unterschätzen. Nearing (1998) erklärt dies durch das Vorhandensein von natürlichen<br />

Variationen in den Modell-Daten, also Variationen, die das Modell nicht abbilden kann. Der<br />

Einfluss von Turbulenzen auf den Bodenabtrag wird auch von Nearing et al. (1991) gezeigt.<br />

In einer Laboruntersuchung messen sie shear stress-Werte zwischen 0,5 Pa und 2 Pa,<br />

während die Scherfestigkeit des Substrates zwischen 1 kPa und 2 kPa liegt, ein<br />

Größenunterschied um den Faktor 1000. Trotz dieses Ungleichgewichtes werden deutliche<br />

Abtragsraten gemessen. Nearing et al. (1991) erklären dieses Ergebnis mit plötzlich<br />

auftretenden Turbulenzen, die viel stärker sind als der mittlere shear stress. Die<br />

beschriebenen Probleme und Lücken zeigen deutlich die Notwendigkeit, auf experimentelle<br />

Art Daten zu gewinnen.<br />

1.7 Weitere Anwendungen<br />

Im Rahmen dieser Arbeit werden die Spülversuche in einer Studie eingesetzt, die nur indirekt<br />

die Problematik der Rinnenerosion mit einschließt. Der Spülversuch findet Verwendung in<br />

einer Studie zur goat erosion, also Erosion, die durch Viehtritt von Ziegen auf einer<br />

vegetationsarmen Fläche entsteht (Ries et al., subm. 2011). In dieser Studie wird der Einfluss<br />

von Viehtritt auf den Bodenabtrag ermittelt. Dazu wird eine Rinne zunächst ohne Viehtritt<br />

„bespült“ und beprobt, danach in einem zweiten Versuch noch mal nach dem Queren von 600<br />

Tieren. Die Ergebnisse zeigen, dass die Ziegen eine relativ große Menge an Material lockern,<br />

welches jedoch sehr schnell aus der Rinne ausgespült wird (bulldozer-effect). Unterschiede<br />

zwischen den beiden Versuchen sind nur in den ersten Durchgängen zu erkennen, die<br />

Unterschiede sind zudem auf die Wasserfront und die 30-Sekunden-Probe beschränkt. Die<br />

Sedimentkonzentrationen erreichen in den Versuchen mit vorherigem Viehtritt hier deutlich<br />

höhere Werte, in den nachfolgenden Proben weisen die Werte mit und ohne<br />

Ziegentritteinfluss wieder ähnliche Werte auf (Ries et al., subm. 2011).<br />

24


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Auch wenn die Spülversuche nicht auf diese Fragestellung zugeschnitten sind, können sie hier<br />

dennoch gewinnbringend eingesetzt werden. Diese Tatsache zeigt den Wert der Methode<br />

„Spülversuch“ im Rahmen einer übergeordneten Erosionsforschung.<br />

Weitere Anwendungsbereiche sind z. B. die Quantifizierung von Abfluss und Abtrag aus<br />

Fahrspuren auf Waldwegen und Rückegassen. Dazu liegen bereits erste Ergebnisse aus<br />

verschiedenen Testgebieten in Luxemburg und Deutschland vor, diese sind bisher jedoch<br />

nicht publiziert. Die Wirksamkeit linearer Strukturen wie Waldwege und Rückegassen als<br />

Abfluss- und Sedimentquelle in Wäldern ist allgemein bekannt und akzeptiert (z.B. Anderson<br />

& Macdonald, 1998; Appelboom et al., 2002; Eastaugh et al., 2008; Grayson et al., 1993;<br />

Motha et al., 2003). Auf diesen linearen Strukturen befinden sich jedoch oftmals zusätzlich<br />

mehr oder weniger deutlich ausgeprägte Fahrspuren. Die Frage ist also, welchen Einfluss<br />

diese kleineren linearen Formen haben. Um diese Frage zu beantworten, werden auf den<br />

Wegen Beregnungssimulationen, in den Fahrspuren Spülversuche durchgeführt. Die runoff-<br />

Koeffizienten auf den Wegen und in den Fahrspuren erweisen sich als ähnlich, sie liegen<br />

meist zwischen 60 und fast 100 % (siehe Abb. 10), d. h. die Wirksamkeit des Abflusses ist in<br />

beiden Formen als vergleichbar anzusehen.<br />

Abbildung 10: Runoff-Koeffizienten in Fahrspuren (Spülversuche) und auf Waldwegen<br />

(Beregnungsversuche). LFE = low flow experiment (Einleitintensität 9 L min -1 ), HFE = high flow<br />

experiment (Einleitintensität 250 L min -1 )<br />

25


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Die Erosions-Effizienz zeigt ein etwas komplexeres Bild. In den Fahrspuren sind sowohl sehr<br />

hohe als auch sehr niedrige Werte zu verzeichnen. Insgesamt verursacht der Abfluss in den<br />

Fahrspuren bis zu 100-mal höhere Abtragswerte als der Abfluss auf dem restlichen Teil der<br />

Wege und Rückegassen, 432 g m -2 min -1 in den Fahrspuren gegenüber 4 g m -2 min -1 (siehe<br />

Abbildung 11).<br />

Abbildung 11: Abtragswerte in Fahrspuren (Spülversuche) und auf Waldwegen<br />

(Beregnungsversuche). LFE = low flow experiment (Einleitintensität 9 L min -1 ), HFE = high flow<br />

experiment (Einleitintensität 250 L min -1 )<br />

Des Weiteren wäre der Einsatz auf Ackerflächen möglich und besonders im Bereich des<br />

Vorgewendes am Ackerrand sinnvoll. In diesem Bereich entwickeln sich sehr schnell deutlich<br />

ausgebildete Rinnen in Gefällsrichtung, da die normale Bearbeitungsrichtung meist parallel zu<br />

den Höhenlinien verläuft, im Vorgewende im rechten Winkel dazu.<br />

1.8 Zusammenfassung und Ausblick<br />

Die vorliegende Arbeit setzt sich mit verschiedenen Fragen und Problemen bezüglich der<br />

Rinnenerosion auseinander. Dabei kann zunächst gezeigt werden, dass Rinnen sehr effektive<br />

Abflusssammler sind. Es wird ein vergleichsweise großer Anteil der Gesamtfläche eines<br />

Einzugsgebietes über vorhandene Rinnennetzwerke entwässert, selbst wenn nur ein kleiner<br />

Teil eines betrachteten Gebietes (z. B. Gully-Einzugsgebiet) tatsächlich von Rinnen bedeckt<br />

26


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

ist. Die Abflussintensitäten sind insgesamt in den interrill-Bereichen höher, da diese Bereiche<br />

einen weit größeren Teil einnehmen. Bezogen auf eine bestimmte Grundfläche weisen die<br />

Rinnen jedoch eine deutlich höhere Abflussintensität auf. Dadurch kann zum einen das<br />

Wasser eines Einzugsgebietes einem vorhandenen Gully schnell zugeführt werden, zum<br />

anderen kann durch den konzentrierten Abfluss bereits in der Rinne eine starke Erosionsarbeit<br />

geleistet werden. Durch ihre im Vergleich zu Gullys deutlich höhere Aktivitätsfrequenz<br />

können in bestimmten Einzugsgebieten die Rinnen die effektivsten Sedimentquellen sein. In<br />

den vorliegenden Untersuchungen liegen die Abtragsraten in den Rinnen bis um den Faktor<br />

60 über den Abtragsraten auf der Fläche. Die Frage ist nun, welche Prozesse für diese<br />

effektive Sedimentbereitstellung verantwortlich sind. Es können mehrere parallel ablaufende<br />

Prozesse identifiziert werden, die mit einer großen räumlichen und zeitlichen Variabilität<br />

ablaufen. Diese Prozesskombination besteht zunächst aus dem Transport des losen Materials,<br />

welches in der Rinne vorhanden ist, zusätzlich wird durch Einschneiden die Rinnensohle<br />

tiefergelegt. Die vermutlich wichtigsten Sedimentquellen sind jedoch die rückschreitende<br />

Erosion an Stufen und headcuts, das Nachbrechen von Seitenwänden als Folge der seitlichen<br />

Unterschneidung und das Überlaufen von plunge pools unterhalb von Stufen. Diese<br />

Überlagerung unterschiedlicher Prozesse bereiten aktuell verwendeten, physikalisch basierten<br />

Bodenerosionsmodellen große Probleme. Diese Modelle verwenden meist hydraulische<br />

Parameter, die theoretisch in einem linearen Verhältnis zur Abtragsrate stehen. Allerdings<br />

wird nur das Einschneiden in die Rinnensohle hydraulisch kontrolliert, die anderen Prozesse<br />

haben nur bedingt etwas mit dem fließenden Wasser zu tun. Daher ist auch die Annahme, dass<br />

die transport rate die transport capacity nicht übersteigen könne, nicht zu bestätigen,<br />

zumindest dann nicht, wenn die allgemein üblichen Gleichungen zur Berechnung der<br />

transport capacity verwendet werden. Als weitere Probleme der Bodenerosionsmodelle<br />

können Unklarheiten bei der Definition und der Herleitung bestimmter Parameter ausgemacht<br />

werden. Da die gegebenen Modelle keine zufriedenstellenden Ergebnisse liefern, ist es<br />

wichtig, weiterhin experimentelle Daten zu erheben. Dazu ist eine reproduzierbare Methode<br />

erforderlich, welche mit den Spülversuchen entwickelt und im Laufe der Zeit immer weiter<br />

verbessert worden ist, um ein möglichst breites Spektrum an Informationen generieren zu<br />

können. Neben Abtragswerten können auch hydraulische Parameter erfasst bzw. aus erfassten<br />

Daten abgeleitet werden. Die letzte hier vorgestellte Version des Versuchsaufbaus ist ein<br />

Etappenziel, weitere Verbesserungen und Erweiterungen sind bereits in der Erprobungsphase.<br />

Die Methode kann zudem innerhalb anderer Fragestellungen im Bereich Oberflächenabfluss<br />

27


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

und Bodenabtrag erfolgreich eingesetzt werden, Untersuchungen zur Ziegenerosion<br />

profitieren bereits von der Methode „Spülversuch“.<br />

Eine Schwachstelle der aktuellen Versuchsversion liegt in der fehlenden Information<br />

bezüglich der tatsächlichen Quelle des bewegten Sedimentes. Bisher können diese Quellen<br />

nur durch Beobachtungen während des Versuches erfasst werden. Dabei ist jedoch eine<br />

Quantifizierung nicht möglich, des Weiteren werden auf diese Weise nur sehr große und<br />

damit deutlich sichtbare Ereignisse berücksichtigt. Aktuell wird überprüft, ob die Erstellung<br />

von Differenzbildern eines DGMs (digitales Geländemodell) aus Stereoluftbildpaaren der<br />

Aufnahmehöhe 1 m über Grund geeignet ist, Aussagen bezüglich der Menge des beim<br />

Spülversuch bewegten/abgetragenen Sediments zu machen. Dabei handelt es sich um eine<br />

Anpassung der bereits erprobten Methode des großmaßstäbigen Luftbildmonitorings, welches<br />

seit Jahren in der Arbeitsgruppe der Physischen <strong>Geographie</strong> an der Universität Trier unter<br />

Leitung von Herrn Prof. Ries verwendet wird, um die Entwicklung von Gullys zu erfassen.<br />

Statt einer fliegenden Kameraplattform wird hier ein terrestrisches Trägersystem verwendet.<br />

Die Versuchsanordnung soll sicherstellen, dass beide Kameras mit gleichbleibendem Abstand<br />

über die Geländeoberfläche geführt werden und in Abständen von ca. 1 m Bildpaare von der<br />

Versuchsfläche aufnehmen. Die Berechnung der Flughöhe über Grund erfolgt überschlägig<br />

anhand der gültigen Formeln für Abbildungsmaßstab und technischer Daten des<br />

Kameraherstellers. Zur Volumenberechnung anhand der Differenzbilder ist die zu bespülende<br />

Geländeoberfläche vor und nach der Durchführung des Spülversuchs mit einer Serie von<br />

Stereoluftbildpaaren unter Beachtung der für die Luftbilderstellung üblichen Überdeckung<br />

von ca. 60 % lückenlos abzudecken. Nach Durchführung erster Tests ist eine Konstruktion<br />

vergleichbar einem Turmdrehkran auf Basis eines Vermessungsstatives hergestellt worden.<br />

Diese Stereoluftbildpaare sollen mit dieser in Abbildung 12 dargestellten Versuchsanordnung<br />

aufgenommen werden, wobei durch die niedrige Aufnahmehöhe von ca. 1 m über Grund<br />

sichergestellt wird, dass auch kleine Mengen des umgelagerten Sediments erfasst werden<br />

können.<br />

28


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Abbildung 12: Drehkrankonstruktion: A: Gesamtgestell, B: Kameraaufhängung, C: maximale<br />

Reichweite, D: minimale Reichweite<br />

Zu behebende Probleme liegen zum einen in der Tatsache, dass hohe Temperaturen 1) die<br />

Akkulaufzeit auf ca. 10 min begrenzen und 2) sich die Kameragehäuse verziehen. Das<br />

Akkuproblem kann provisorisch über eine modifizierte Stromversorgung (Autobatterie <br />

Umformer 12 V auf 220 V Handyladegerät Akkudummy) gelöst werden. Eine<br />

Abschattung der Rinne verringert die Hitze und verbessert die Licht- und damit auch die<br />

Bildqualität. Zum anderen verhindert die Ausführung mit nur zwei Kameras die Aufnahme<br />

von Überhängen in den Rinnenwänden. Dazu sollen in Zukunft zwei weitere, schräg<br />

angebrachte Kameras an der Kameraschiene montiert werden. Die Auswertung der Bilder<br />

erfolgt über das Standardprogramm Leica Photogrammetry Suite.<br />

Ein weiterer Ansatzpunkt für die Auswertung der erstellten Fotos liegt in der Verwendung<br />

einer Kombination von kommerziellen und open source - Programmen zur 3-D Visualisierung<br />

von Rinnen und zur automatischen Berechnung von Umlagerungsvolumina. Das erste zu<br />

verwendende Programm ist MS-Photosynth©. Dieses Programm stellt aus einer Vielzahl von<br />

Einzelfotos ein Gesamtbild zusammen. Dabei werden nicht mehr, wie bisher üblich,<br />

29


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

bestimmte Passpunkte benötigt, die in jedem der Fotos vorhanden sein müssen, das Programm<br />

erkennt vielmehr selbstständig Punktcluster (Übereinstimmung bezüglich Helligkeitswerten,<br />

Farbwerten, Kontrastwerten), die in überlappenden Fotopaaren vorhanden sind. Anhand<br />

dieser feature-points, also Punkthaufen, die in den Fotos zu finden sind, werden räumliche<br />

Markierungen gesetzt, anhand derer die Einzelbilder in Übereinstimmung gebracht und soweit<br />

entzerrt werden, dass ein passgenaues Panorama entsteht. Dadurch können Fotos kombiniert<br />

werden, die aus sehr unterschiedlichen Winkeln aufgenommen worden sind. Als nächster<br />

Schritt wird die Punktwolke aus MS-Photosynth© extrahiert. Jeder Punkt der Punktwolke<br />

weist zu diesem Zeitpunkt relative x-, y-, und z- Koordinaten auf. Dieses Zwischenergebnis<br />

wird in open source - Programmen weiterverarbeitet. In PMVS2 (Patch-Based Multi-View<br />

Stereo) und CMVS (Clustering for Multi-View Stereo) wird die Punktwolke mit den<br />

vorhandenen Bildinformationen überlagert. Mesh Lab trianguliert zwischen den jeweils<br />

benachbarten Punkten der Punktwolke, es entsteht ein dreidimensionales Gitternetz.<br />

Probleme liegen v.a. in der unbekannten Genauigkeit und dem Fehlen eines Maßstabes dieser<br />

in MS-Photosynth© erstellten Höhenmodelle. Weiterhin lassen die mit MS-Photosynth©<br />

erstellten Punktwolken immer wieder Bereiche ohne Werte erkennen. Diese Bereiche führen<br />

in der Volumendifferenzberechnung zu großen Fehlern. Das Ziel der Versuche liegt jedoch<br />

darin, auch möglichst kleine Mengen von verlagertem Substrat zu erfassen, die Gesamt-<br />

Fehlergröße liegt jedoch geschätzt um mehrere Größenordnungen über der theoretisch durch<br />

das Aufnahmesystem möglichen Genauigkeit. Des Weiteren muss MS-Photosynth© als<br />

black-box-Modell angesehen werden, die verwendeten Algorithmen sind nicht bekannt und<br />

dadurch fehlt die Kontrolle über Fehlergrößen, -mengen und -quellen. Angedacht für die<br />

Zukunft ist zunächst die Verwendung eines nicht cloud-basierten Programmes als Alternative<br />

zu MS-Photosynth©. Dadurch soll das Auftreten von unkontrollier- und unkorrigierbaren<br />

Fehlern schon im ersten Verarbeitungsschritt ausgeschlossen werden. Nach Herstellung der<br />

endgültigen Einsatzbereitschaft wird diese Methode eine kostengünstige Variante zum sehr<br />

viel kostenintensiveren Laserscanning darstellen.<br />

Des Weiteren soll zukünftig die Fließgeschwindigkeit im Bereich von einem Meter ober- und<br />

unterhalb der Messstellen mit Hilfe von einfachen Webcams gemessen werden. Dadurch kann<br />

Personal eingespart und Messkampagnen selbst mit kleineren Teams realisiert werden. Die<br />

Auswertung des Filmmaterials und die Ermittlung der Fließgeschwindigkeit kann im<br />

Anschluss an die Geländearbeit durchgeführt werden.<br />

Ein wünschenswertes Ziel wäre es, mithilfe der gewonnenen Daten und Beobachtungen ein<br />

funktionierendes Bodenerosionsmodell bzw. Submodell rill erosion zu entwickeln. Fraglich<br />

30


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

bleibt jedoch, ob die Vielzahl verschiedener Prozesse - jeder mit hoher räumlicher und<br />

zeitlicher Variabilität - mittels empirischer Formeln ausreichend genau beschrieben werden<br />

kann. Ein neuer Ansatz besteht in der Verwendung von Wahrscheinlichkeits-<br />

Dichtefunktionen. Die erfolgversprechendste, aber noch nicht einsatzbereite Version ist von<br />

Sidorchuk (Sidorchuk, 2002; Sidorchuk et al., 2004; Sidorchuk, 2005 a, b; Sidorchuk et al.,<br />

2008; Sidorchuk, 2009 a, b) entwickelt worden. Diese stochastischen Modelle reduzieren die<br />

Anzahl der empirischen Komponenten, wodurch das theoretische und das Prozesswissen<br />

erhöht und der black-box-Charakter der empirischen Modelle verhindert werden kann. Es<br />

wird nicht mehr ein einzelner Wert als Eingangsgröße sondern Wahrscheinlichkeits-<br />

Dichtefunktionen (probability density functions, pdf’s) des gemessenen Parameters<br />

verwendet. Das Problem dieser Modelle der dritten Generation (Sidorchuk, 2009 b) ist, dass<br />

der Typ der Wahrscheinlichkeits-Dichtefunktion noch nicht bekannt ist. Es gibt eine große<br />

Anzahl von Boden- und Umweltparametern, welche die Erosionsresistenz beeinflussen, und<br />

es ist nicht klar, ob dieses komplexe Konzept durch Wahrscheinlichkeits-Dichtefunktionen<br />

beschrieben werden kann. Aufgrund der benötigten Anzahl von Wahrscheinlichkeits-<br />

Dichtefunktionen und der benötigten Computerleistung scheint dieser Ansatz laut Knapen et<br />

al. (2007) zu kompliziert zu sein, sodass bis auf Weiteres nicht auf experimentelle<br />

Bodenerosionsforschung verzichtet werden kann; zum einen zur Verbesserung des<br />

Grundlagenwissens, zum anderen zur Ermittlung der pdf’s in der stochastischen<br />

Bodenerosionsmodellierung.<br />

Zwei Dinge sind zu unserer Arbeit nötig: Unermüdliche<br />

Ausdauer und die Bereitschaft, etwas, in das man viel<br />

Zeit und Arbeit gesteckt hat, wieder wegzuwerfen.<br />

Albert Einstein<br />

31


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Quellenangaben<br />

Anderson, D.M., Macdonald, L.H. (1998): Modelling road surface sediment production using<br />

a vector geographic information system. Earth Surface Processes and Landforms 23,<br />

95–107.<br />

Appelboom, T.W., Chescheir, G.M., Skaggs, R.W., Hesterberg, D.L. (2002): Management<br />

practices for sediment reduction from forest roads in the coastal plains. Transactions<br />

of ASAE 45(2), 337–344.<br />

Auerswald, K., Fiener, P., Dikau, R. (2009): Rates of sheet and rill erosion in Germany - A<br />

meta-analysis. Geomorphology 111(3-4), 182-193.<br />

Avni, Y. (2005): Gully incision as a key factor in desertification in an arid environment, the<br />

Negev highlands, Israel. Catena 63, 185-220.<br />

Beuselinck, L., Govers, G., Hairsine, P., Sander, G., Breynaert M. (2002): The influence of<br />

rainfall on sediment transport by overland flow over areas of net deposition. Journal of<br />

Hydrology 257, 145-163.<br />

Brodie, I., Rosewell C. (2007): Theoretical relationships between rainfall intensity and kinetic<br />

energy variants associated with stormwater particle washoff. Journal of Hydrology,<br />

340(1-2), 40-47.<br />

Brunton, D.A., Bryan, R.B. (2000): Rill network development and sediment budgets. Earth<br />

Surface Processes and Landforms 25, 783-800.<br />

Bryan, R.B., Poesen, J. (1989): Laboratory experiment on the influence of slope length on<br />

runoff, percolation and rill development. Earth surface processes and Landforms 14,<br />

211-231.<br />

Bryan, R.B. (2000): Soil erodibility and processes of water erosion on hillslope.<br />

Geomorphology 32(3-4), 385-415.<br />

Casali, J., Loizu, J., Campo, M., De Santisteban, L., Alvarez-Mozos, J. (2006): Accuracy of<br />

methods for field assessment of rill and ephemeral gully erosion. Catena 67(2), 128-<br />

138.<br />

Cerdan, O., Le Bissonnais, Y., Couturier, A., Bourennane, H., Souchère, V. (2002): Rill<br />

erosion on cultivated hillslopes during two extreme rainfall events in Normandy,<br />

France. Soil and Tillage Research 67(1), 99-108.<br />

Constantin, P. (2001): Some open problems and research directions in the mathematical study<br />

of fluid dynamics. Mathematics unlimited-2001 and beyond, 353-360.<br />

32


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Eastaugh, C. S., Rustomji, P. K., Hairsine, P. B. (2008): Quantifying the altered hydrologic<br />

connectivity of forest 23 roads resulting from decommissioning and relocation.<br />

Hydrological Processes 22, 2438–2448.<br />

Faust, D., Schmidt, M. (2009): Soil erosion processes and sediment fluxes in a Mediterranean<br />

marl landscape, Campina de Cádiz, SW Spain. Z.Geomorph. N.F. 52 (2), 247-265.<br />

Fefferman, C.L. (2006): Existance and smoothness of the Navier-Stokes equation, The<br />

millennium prize problems. Clay Math. Inst., Cambridge, MA, 57-67.<br />

Gilley, J.E., Kottwitz, E.R., Simanton, J.R. (1990): Hydraulic characteristics of rill.<br />

Transactions of the ASAE 33, 1900-1906.<br />

Giménez, R., Govers, G. (2002): Flow Detachment by Concentrated Flow on Smooth and<br />

Irregular Beds. Soil Sci Soc Am J 66(5), 1475-1483.<br />

Govers, G. (1987): Spatial and temporal variability in rill development processes at the<br />

Huldenberg experimental site. Catena Supplement 8, 17-34.<br />

Govers, G., Everaert, W., Poesen, J., De Ploey, J., Lautridou, J.P. (1990): A Long flume study<br />

of dynamic factors affecting the resistance of loamy soil to concentrated flow erosion.<br />

Earth surface processes and landforms 15, 313-328.<br />

Govers, G. (1991): Time-dependency of runoff velocity and erosion: the effect of the initial<br />

soil moisture profile. Earth Surface Processes and Landforms 16, 713-729.<br />

Govers, G. (1992): Evaluation of transporting capacity formulae. In: Parsons, A.J., Abrahams,<br />

A.D. (Eds.): Overland flow – Hydraulics and erosion mechanics. UCL-Press,<br />

Guildford, 243-273.<br />

Govers, G., Giménez, R., Van Oost, K. (2007): Rill erosion: Exploring the relationship<br />

between experiments, modelling and field observations. Earth-Science Reviews 84(3-<br />

4), 87-102.<br />

Grayson, R.B., Haydon, S.R., Jayasuriya, M.D.A., Finlayson, B.L. (1993): Water quality in<br />

mountain ash forests—separating the impacts of roads from those of logging<br />

operations. Journal of Hydrology 150: 459–480.<br />

Huang, C., Bradford, J.M., Laflen, J.M. (1996): Evaluation of the detachment-transport<br />

coupling concept in the WEPP rill erosion equation. Soil Science Society American<br />

Journal 60, 734-739.<br />

Knapen, A., Poesen, J., Govers, G., Gyssels, G., Nachtergaele, J. (2007): Resistance of soils<br />

to concentrated flow erosion: A review. Earth-Science Reviews 80(1-2), 75-109.<br />

Kuhn, N., Bryan, R.B., Navar, J. (2003): Seal formation and interrill erosion on a smectiterich<br />

Kastanozem from NE-Mexico. Catena 52, 149-169.<br />

33


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Kuhn, N., Bryan, R. (2004): Drying, soil surface condition and interrill erosion on two<br />

Ontario soils. Catena 57, 113-133.<br />

Le Bissonnais, Y., Cerdan, O., Lecomte, V., Benkhadra, H., Souchère, V., Martin, P. (2005):<br />

Variability of soil surface characteristics influencing runoff and interrill erosion.<br />

Catena 62, 111-124.<br />

Loch, R.J. (2000): Using Rainfall simulation to guide planning and management of<br />

rehabilitated areas: Part I. Experimental methods and results from a study at the<br />

northparkes mine, Australia. Land Degradation and Development 11, 221-240.<br />

Mancilla, G.A., Chen, S., McCool, D.K. (2005): Rill density prediction and flow velocity<br />

distribution on agricultural areas in the Pacific Northwest. Soil and Tillage Research<br />

84, 54-66.<br />

Marzolff, I., Ries, J.B., Poesen, J. (2011): Short-term vs. medium-term monitoring for<br />

detecting gully-erosion variability in a Mediterranean environment. Earth Surface<br />

Processes and Landforms 36, 1604–1623.<br />

Merz, W. (1993): Experimentelle Untersuchungen zur Rillenerosion auf landwirtschaftlich<br />

genutzten Böden in Kanada und der Volksrepublik China. Freiburger Geographische<br />

Hefte, Heft 40, Selbstverlag des Institutes für Physische <strong>Geographie</strong> der Albert-<br />

Ludwig-Universität Freiburg i.Br., Freiburg, 147 pp.<br />

Meyer, L.D., Forster, G.R., Römkens, M.J.M. (1975): Source of soil eroded by water from<br />

upland slopes. Proc. 1972 Sediment Yield Workshop, US Dept. of Agric. Sediment.<br />

Lab., Oxford, Mississippi: 177-189.<br />

Morgan, R.P.C., Martin, L., Noble, C.A. (1987): Soil erosion in the United Kingdom: a Case<br />

Study from Mid-Bedfordshire, England. Occasional Paper No.14.<br />

Motha JA, Wallbrink PJ, Hairsine PB, Grayson RB. (2003): Determining the sources of<br />

suspended sediment in a forested catchment in southeastern Australia. Water<br />

Resources Research 39(3): 1056. DOI: 10.1029/2001WR000794.<br />

Nearing, M.A., Bradford, J.M., Parker, S.C. (1991): Soil detachment by shallow flow at low<br />

slopes. Soil Science Society of America Journal 55 (2), 339-344.<br />

Nearing, M.A., Parker, S.C. (1994): Detachment of soil by flowing water under turbulent and<br />

laminar conditions. Soil Sci. Soc. Am. J. 58 (6), 1612-1614.<br />

Nearing, M.A. (1998): Why soil erosion models over-predict small soil losses and underpredict<br />

large soil losses. Catena 32, 15-22.<br />

Poesen, J. (1987): Transport of rock fragments by rill flow – a field study. In: Bryan, R.B.<br />

(Ed.): Rill erosion - process and significance. Catena Supplement 8, 35-54.<br />

34


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Polyakov, V.O., Nearing, M.A. (2003): Sediment transport in rill flow under deposition and<br />

detachment conditions. Catena 51, 33-43.<br />

PTB Mitteilungen (2007): Das Internationale Einheitensystem (SI). Amts- und<br />

Mitteilungsblatt der Physikalisch-Technischen Bundesanstalt, 117. Jahrgang. Heft 2,<br />

Braunschweig und Berlin.<br />

Regüés, D., Balasch, J.C., Castelltort, X., Soler, M., Gallart, F. (2000): Relación entre las<br />

tendencias temporales de producción y transporte de sedimentos y las condiciones<br />

climáticas en una pequeña cuenca de montaña mediterránea (Vallcebre, Pirineos<br />

Orientales). Cuadernos de Investigación Geográfica 26, 41-65.<br />

Richter, G. (2001): Bodenerosion – Analyse und Bilanz eines Umweltproblems.<br />

Wissenschaftliche Buchgesellschaft, 264 Seiten.<br />

Ries, J.B., Marzolff, I. (2003): Monitoring of gully erosion in the Central Ebro Basin by large<br />

scale aerial photography taken from a remotely controlled blimp. Catena 50, 309-328.<br />

Ries, J.B. (2011): Bodenerosion. In: Gebhardt, H., Glaser, R., Radtke, U., Reuber, P. (Hrsg.):<br />

<strong>Geographie</strong> – Physische <strong>Geographie</strong> und Humangeographie. 2. Auflage, Spektrum<br />

Akademischer Verlag, 506-517.<br />

Ries, J.B., Andres, K., <strong>Wirtz</strong>, S., Tumbrink, J., Wilms, T., Peter, K.D., Burczyk, M., Butzen,<br />

V., Seeger, M. (subm. 2011) Sheep and goat erosion – experimental geomorphology<br />

as an approach for the quantification of underestimated processes. Zeitschrift für<br />

Geomorphologie.<br />

Scherer, U. (2008): Prozessbasierte Modellierung der Bodenerosion in einer Lösslandschaft.<br />

Dissertation, Fakultät für Bauingenieur-, Geo- und Umweltwissenschaften, Universität<br />

Fridericiana zu Karlsruhe (TH), Karlsruhe, Germany.<br />

Schneider, G. (2008): Ein Millenniumsproblem: Die globale Existenz und Eindeutigkeit<br />

glatter Lösungen der dreidimensionalen Navier-Stokes-Gleichungen. Mathematik-<br />

Online: Beiträge zu berühmten, gelösten und ungelösten Problemen, Nummer 2.<br />

Seeger, M., Errea, M.P., Beguería, S., Arnáez, J., Martí, C., García-Ruiz, J.M. (2004):<br />

Catchment soil moistureand rainfall characteristics as determined factors for discharge<br />

/ suspended sediment hysteretic loops in a small headwater catchment in the Spanish<br />

Pyrenees. Journal of Hydrology 288, 299-311.<br />

Seeger, M., Errea-Abad, M.-P., Lana-Renault, N. (2005): Spatial Distribution of Soils and<br />

their Properties as Indicators of Degradation/Regradation Processes in a Highly<br />

Disturbed Mediterranean Mountain Catchment. Journal of Mediterranean Ecology 6<br />

(1), 53-59.<br />

35


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Seeger, M. (2007): Uncertainity of factors determining runoff and erosion processes as<br />

quantified by rainfall simulations. Catena 71 (1), 56-67.<br />

Seiler, R. (2002): Die Navier-Stokes-Gleichung. Elemente der Mathematik 57, 109-114.<br />

Sidorchuk, A. (2002): Stochastic Modelling of soil erosion and deposition. Proceedings of the<br />

12 th ISCO Conference, Beijing, China, 26-31 May 2002, 136-142.<br />

Sidorchuk, A., Smith, A., Nikora, V. (2004): Probability distribution function approach in<br />

stochastic modelling of soil erosion. Sediment transfer trough the fluvial system.<br />

IHAS Publ. 288, 345-353.<br />

Sidorchuk, A. (2005a): Stochastic components in the gully erosion modelling. Catena 63,<br />

299-317.<br />

Sidorchuk, A. (2005b): Stochastic modelling of erosion and deposition in cohesive soils.<br />

Hydrol. Process. 19, 1399-1417.<br />

Sidorchuk, A., Schmidt, J., Cooper, G. (2008): Variability of shallow overland flow velocity<br />

and soil aggregate transport observed with digital videography. Hydrological<br />

Processes 22, 4035-4048.<br />

Sidorchuk, A. (2009a): High-Frequency variability of aggregate transport under water erosion<br />

of well-structured soils. Eurasian Soil Sci. 42 (5), 543-552.<br />

Sidorchuk, A. (2009 b): A third generation erosion model: The combination of probabilistic<br />

and deterministic components. Geomorphology 110 (1-2), 2-10.<br />

Stroosnijder, L. (2005): Measurement of erosion: Is it possible? Catena 64, 162-173.<br />

Temam, R. (2000): Some developments on Navier-Stokes equations in the second half of the<br />

20 th century. Development of mathematics 1950-2000, Birkhäuser, Basel, Switzerland,<br />

1049-1106.<br />

Wiegner, M. (1999): The Navier-Stokes equations – a neverending challenge? Jahresbericht<br />

Deutsch. Math.-Verein 101 (1), 1-25.<br />

<strong>Wirtz</strong>, S., Seeger, M., Ries, J.B. (2010): The rill experiment as a method to approach a<br />

quantification of rill erosion process activity. Zeitschrift für Geomorphologie 54 (1),<br />

47-64.<br />

<strong>Wirtz</strong>, S., Seeger, M., Ries, J.B. (2012a): Field experiments for understanding and<br />

quantification of rill erosion processes. Catena 91, 21-34.<br />

<strong>Wirtz</strong>, S., Iserloh, T., Rock, G., Hansen, R., Marzen, M., Seeger, M., Betz, S., Remke, A.,<br />

Wengel, R., Butzen, V., Ries, J.B. (2012b): Soil erosion on abandoned land in<br />

Andalusia – a comparison of interrill – and rill erosion rates. ISRN Soil Science,<br />

doi:10.5402/2012/730870.<br />

36


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

<strong>Wirtz</strong>, S., Seeger, M., Zell, A., Wagner, C., Wagner, J.-F., Ries, J.B. (subm. 2012):<br />

Applicability of different hydraulic parameters to describe soil detachment in eroding<br />

rills. PLoSOne.<br />

<strong>Wirtz</strong>, S., Seeger, M., Remke, A., Wengel, R., Wagner, J.-F., Ries, J.B. (2013): Do<br />

deterministic sediment detachment and transport equations adequately represent<br />

process-interactions in eroding rills? An experimental field study. Catena 101, 61-78.<br />

37


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Kapitel 2<br />

<strong>Wirtz</strong> et al. (2012b): Soil erosion on abandoned land in Andalusia – a comparison<br />

of interrill– and rill erosion rates. ISRN Soil Science, doi: 10.5402/2012/730870.<br />

38


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Soil erosion on abandoned land in Andalusia – a comparison of interrill- and rill erosion<br />

rates<br />

<strong>Wirtz</strong>, S. (1); Iserloh, T. (1); Rock, G. (2); Hansen, R. (1); Marzen, M. (1); Seeger, M. (1);<br />

Betz, S. (1); Remke, A. (1); Wengel, R. (1); Butzen, V. (1); Ries, J.B. (1)<br />

(1): Trier University, Dep. of Physical Geography, Behringstraße, 54286 Trier, Germany<br />

(2): Trier University, Remote Sensing Department, Behringstraße, 54286 Trier, Germany<br />

Abstract<br />

The present paper is based on several field investigations (monitoring soil and rill erosion by<br />

aerial photography, rainfall simulations with portable rainfall simulators and manmade rill<br />

flooding) in southern Spain. Experiments lead now to a closer understanding of the dynamics<br />

and power of different soil erosion processes in a gully catchment area.<br />

The test site Freila (Andalusia, SE Spain) covers an area of 10.01 ha with a rill density of 169<br />

m ha -1 , corresponding to a total rill length of 1694 m. Assuming an average rill width of 0.15<br />

m, the total rill surface can be calculated at 250 m² (0.025 ha). This given, the surface covered<br />

by rills makes up only 0.25 % of the total test site. Since the rill network drains 1.98 ha, 20 %<br />

of the total runoff comes from rills. The rills’ sediment erosion was measured and the total<br />

soil loss was then calculated for detachment rates between 1685 g m -2 and 3018 g m -2 . The<br />

interrill areas (99.75 % of the test site) show values between 29-143 g m -2 . This suggests an<br />

important role of rill erosion concerning runoff and soil detachment.<br />

Keywords<br />

Soil erosion, water erosion, rill erosion, interrill erosion, rainfall simulation<br />

1 Introduction<br />

Soil erosion by water involves different physical processes at variable spatial and temporal<br />

scales. The two main processes are interrill and rill erosion by runoff water, the mechanisms<br />

of these two processes are completely different. The soil detachment in interrill erosion is<br />

induced and enhanced by splash and shallow overland flow [1]. In addition, it is influenced by<br />

the intrinsic characteristics of the soils [2, 3, 4] and rainfall intensity [5, 6]. In contrast, rill<br />

erosion is caused by a concentrated overland flow [6, 7, 8]. This is considered to be the most<br />

important process of sediment erosion and soil loss [9, 10]. As a result, the new rills can<br />

39


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

become persistent and form gullies, potentially constraining further land use [11, 12, 13, 14].<br />

Especially on fallow and shrub land, rills can develop without being disturbed by land<br />

management measures like ploughing. Since huge areas in the Mediterranean are covered by<br />

fallow and shrub land [15], rill erosion can be assumed to be a major process of soil erosion in<br />

the Mediterranean.<br />

The outstanding importance of linear erosion for all intents and purposes (geomorphology,<br />

hydrology and economy) can be explained by the amount of kinetic energy of water, running<br />

as a concentrated runoff in channels. When concentrated, water reaches its maximum impact<br />

concerning erosion and transport [16]. The percentage of interrill-, rill- and gully erosion on<br />

the total soil loss in a catchment area is difficult to determine, and results of many research<br />

groups differ considerably in comparing the effectiveness of the different soil erosion<br />

processes.<br />

Results concerning the proportion of gully erosion on total soil erosion vary between 10 %<br />

and more than 90 % [17]. Gullies need high-intensity rainfall events to be activated. Most<br />

thunderstorms do not activate the gully in a catchment, but are able to generate or reactivate<br />

smaller forms, the rills. Faust & Schmidt [18] considered the geomorphologic importance of<br />

gullies as quite low, corresponding to the rare activity of the gullies. They state an activity<br />

frequency of one single event in 20 years, compared with an assumed activity of the rills of<br />

about four times per year. On that account, the rills deliver smaller sediment quantities, but<br />

this can occur several times within one year. Meyer et al. [19] suggested a triplication of<br />

erosion rates due to rill development. Cornfields in Bedfordshire (Great Britain) lost as much<br />

as 9 to 21 times more sediment by rill erosion, compared with erosion only affected by<br />

interrill erosion [20]. Cerdan et al. [9] reported the soil loss by rill erosion in Normandy<br />

during only a few heavy rainfall-events to be up to 90 % of the total soil loss.<br />

During the last decades, several approaches have been developed to describe and predict soil<br />

detachment and sediment transport in rills in a reliable way [7, 21, 22]. In contrast, the factors<br />

influencing the development and the behaviour of rills have not been comprehensively<br />

assessed yet. The relationships between factors influencing these processes remain unclear as<br />

well. Knapen et al. [8] attribute this to the lack of comparable data that could be used for a<br />

meta-analysis on rill erosion. Furthermore, there is still little known about the function of rills<br />

in specific environments.<br />

In this paper, the results of a set of experimental methods are used to draw conclusions from<br />

the differences between interrill- and rill erosion dynamics on a gully catchment in Andalusia<br />

(Spain). Three main questions were to be answered:<br />

40


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

1. Which proportion of the test site area is drained by rill networks?<br />

2. Which runoff intensities appear in the rills, which on the interrill-areas?<br />

3. How effective is the rill erosion compared to the interrill erosion in the catchment?<br />

2 Material and Methods<br />

2.1 Study Area<br />

The study area Freila is situated in the Hoya de Baza sedimentary basin (see figure 1) in<br />

Andalusia (Southeast Spain). The bedrock mostly consists of Pliocene sedimentary rocks, i.e.<br />

marls and fine grained sandstones. At the surface the marls and sandstones are weathered to<br />

calcareous loamy to sandy lithosols. The semi-arid climate is characterised by a mean annual<br />

temperature of about 14.2 °C. The annual precipitation is down to only 368 mm, with a high<br />

inter-annual variability. The vegetation is dominated by low shrub-land of Thymus vulgaris<br />

and Stipa tenacissima grassland. The land use on the southern lake-side of the Negratin<br />

reservoir is dominated by abandoned cereal fields, which are extensively grazed by flocks of<br />

sheep. Agricultural land use comprises mainly cereal dry-farming [23]. At the top of the<br />

catchment, a gully has developed.<br />

Figure 1: Test site in Andalusia (UTM: 509864 E – 4154369 N)<br />

41


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

2.2 Methods<br />

During the field work, we combined the experimental methods rill experiment and rainfall<br />

simulation with field mapping. As a geodetic control of the maps we used large scale aerial<br />

photographs, taken with our own equipment.<br />

2.2.1 Rill experiment<br />

During the rill experiments, water was pumped into an existing erosion rill using a motor<br />

driven pump. For low flow experiments (LFEs) we used a constant discharge at the inlet of 9<br />

L min -1 for 8 minutes. The total amount of water was 72 L. In the high flow experiments<br />

(HFEs) we used a pump with a capacity between 250 and 330 L min -1 . During 3 to 4 minutes<br />

we reached a water discharge of about 1000 L. The mobilisation of sediments due to<br />

turbulences of water at the inlet of the measuring channel was suppressed by a special inlet<br />

construction.<br />

Each rill experiment consisted of two runs. In the first run the rill was tested under field (dry)<br />

conditions and in a second run, about 15 min later, the same rill was tested under wet<br />

conditions. The flow velocity within the rill was measured by recording the travel times of the<br />

waterfront and of two applied colour tracers. In the LFEs, the colour tracers were used after<br />

the flow stabilisation (3 and 6 minutes) and in the HFEs after 1 and 2 minutes. Accordingly,<br />

three velocity curves were recorded and changes in flow dynamics can be detected. The used<br />

colour tracers were the food colourings E 124 (red) and E 13 (blue).<br />

The rill's slope was measured using a spring bow of 1 m range with a digital spirit level. It has<br />

to be considered, that slope measuring provided only average slopes for each meter of the rill<br />

length, so steps or knick-points in the rill were not captured. However, the position of each<br />

step or knick-point and also the heights were separately recorded.<br />

At each of the three measuring points (MP) along the rill, four water samples were taken: the<br />

first directly when the waterfront has reached the sampling point, the second 30 sec later, the<br />

third after 90 sec and the fourth 150 sec after the arrival of the waterfront. The sediment<br />

concentration of each sample was determined by filtration in the laboratory. [24, 25]<br />

At each MP, the rill cross section was measured by means of a yardstick (LFEs) or a laser<br />

rangefinder (HFEs). The distance between ground level and sensor respectively, and rill<br />

bottom was measured in 0.02 m steps. This method allowed an accurate calculation of the<br />

rill’s cross section area and also an estimation of the rills’ volume at the MP.<br />

The water height was measured simultaneously with the time of sampling by means of a<br />

yardstick in the LFEs or continuously by ultrasonic sensors (HFEs) at each measuring point.<br />

42


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

2.2.2 Rainfall Simulations<br />

The test plot was circular, with a diameter of 60 cm and an area of 0.28 m 2 . It was delimited<br />

by a steel ring of 7 cm height, which was introduced into the soil for at least 3 cm. The outlet<br />

was V-shaped and was placed at the deepest point of the plot at surface level. The commercial<br />

full cone nozzle (Lechler 460.608) was fed with a pressure of about 40 kPa (0.4 bars) at a<br />

height of 2 m. The water flow was regulated by a flow metre (Type KSK-1200HIG100)<br />

positioned on a separate pole in 1.5 m height. The resulting rainfall intensity was maintained<br />

throughout the experiments at around 40 mm h -1 , which is considered to be mean<br />

thunderstorm rainfall intensity with a return period of about 2 years [26]. The complete runoff<br />

was collected and runoff and soil detachment was determined gravimetrically. [23, 27, 28, 29,<br />

30]. The standard deviation of the rainfall intensity is < 5 % [28].<br />

Plot surface parameters were rock fragment cover, separated in embedded and overlying,<br />

vegetation cover, bare soil area, separated in crusted and uncrusted and finally the slope of the<br />

plot. These parameters were optically estimated in the field and validated again with the help<br />

of digital photographs taken from every plot, and preferably taken at an angle of 90° to the<br />

plot surface. The slope was measured by a clinometer.<br />

For comparing the variability of the different parameters, we calculated the average of the<br />

relative measurement errors (RME) following the DIN 1319-1 [31]. This error is defined as<br />

xa<br />

xr<br />

f = 100<br />

Eq. (1)<br />

x<br />

r<br />

where x a are the measured values and x r are “correct” values; we used the mean of the<br />

measured values as x r . This value describes the deviation of each single experiment from the<br />

mean of all experiments.<br />

2.2.3 Small Format Aerial Photography (SFAP)<br />

The photos were taken using three different camera carriers depending on the wind<br />

conditions, requested flight level (photographed area) and available manpower: a hot air<br />

blimp, a kite or an unmanned aerial vehicle (UAV). A detailed description of the camera<br />

carriers and aerial photography in general is to find in Aber et al. [32]; the hot air blimp is<br />

also described in Ries & Marzolff [33].<br />

The first used camera was a Canon 350D with a Canon EF-S 20 mm objective, the maximum<br />

solution of the photos is 8 MP, the resulting stereoscopic images show a ground resolution<br />

between 0.5 and 11 cm, depending on the flying height.<br />

43


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

The second used camera was a Nikon Coolpix S6000 digital compact camera with a<br />

maximum solution of 14.48 MP. The zoom-objective has an aperture angle dynamic between<br />

28 and 144 mm and a light intensity between 2.8 and 5.6.<br />

Aerial photographs were used during field work as a basis for the mapping and the<br />

classification in “vegetation areas” and “no vegetation areas”.<br />

2.2.4 Field mapping and GIS Analysis<br />

The field mapping based on the self-made aerial photographs should deliver information<br />

about the spatial distribution of linear erosion forms, the catchment areas of these forms and a<br />

total rill length on the test site. With this information, different parameters can be calculated.<br />

The total rill area is calculated as follows:<br />

A =W L<br />

(Eq. 2)<br />

R<br />

R<br />

R<br />

Where A R = total rill area [m²], W R = estimated average rill width [m] and L R = total rill<br />

length in the test site [m].<br />

The rill density is calculated using the following equation (eq. 3):<br />

D R<br />

= L R<br />

A T<br />

(Eq. 3)<br />

Where D R = rill density [m ha -1 ], L R = total rill length in the test site [m] and A T = test site<br />

area [ha].<br />

The rill drainage index is calculated as follows:<br />

I R<br />

= C R<br />

A T<br />

(Eq. 4)<br />

Where I R = rill drainage index [ha ha -1 ], C R = total area of all rill catchments [ha] and A T =<br />

test site area [ha].<br />

2.2.5 Classification of the aerial photographs<br />

In order to obtain spatial information about the different land cover types and distribution, a<br />

classification of digital UAV imagery has been performed. First, an unsupervised<br />

classification was done for identifying the classes that can be statistically distinguished. As a<br />

result, only three classes could be distinguished, i.e. “vegetation”, “no vegetation/soil” and<br />

“shadow”. A drawback of classifying UAV imagery was the fact, that for such a large area,<br />

multiple flights were needed to cover the whole site. This led to a change of the illumination<br />

conditions for the different flights’ images. Considering these facts, every tile of the mosaic<br />

needed to be classified individually. Using manually digitised training areas, the maximum<br />

44


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

likelihood classifier was chosen. After classification of every single tile, a mosaic was<br />

calculated.<br />

3 Results and Interpretation<br />

3.1 Rill experiments<br />

The rill experiments were accomplished in five different rills, one rill was tested three times,<br />

once in a LFE and two times in a HFE. The catchment areas of the tested rills were between<br />

190 and 1600 m², the tested rill parts showed lengths between 15 and 21 m and widths<br />

between 0.4 and 2.1 m, resulting in rill areas between 6 and 44 m². These values are<br />

summarised in table 1.<br />

Table 1: Setup parameters of the tested rills<br />

Experiment rill Inflow intensity Inflow Catchment Tested rill Estimated Rill area<br />

[L min -1 ] duration [min] area [m²] length [m] rill width [m] [m²]<br />

RE1_2008 1 9 8 415 15 0.4 6<br />

RE2_2008 2 9 8 203 21 0.4 8.4<br />

RE3_2008 3 9 8 579 15 0.5 7.5<br />

RE4_2009 1 250 4 415 16 0.4 6.4<br />

RE5_2009 4 250 4 187 17 0.4 6.8<br />

RE6_2009 5 330 3 1641 21 2.1 44.1<br />

RE7_2009 1 250 4 415 16 0.4 6.4<br />

Figure 2 shows the results of the rill experiments. The sediment concentrations in the rill<br />

experiments with an inflow intensity of 9 L min -1 showed a data range between 0.8<br />

(RE1_2008b) and 4.5 g L -1 (RE2_2008a), the average sediment concentration was 2.3 g L -1 .<br />

The rill experiments with an inflow intensity higher than 250 L min -1 showed sediment<br />

concentrations between 5 (RE4_2009b, RE5_2009b and RE7_2009b) and 20 g L -1<br />

(RE5_2009a), the average sediment concentration in these experiments was 8.6 g L -1 , that<br />

means about 3.5 times higher than in the experiments with an inflow intensity of 9 L min -1 .<br />

The runoff coefficients RCs showed a data range between 11 and around 95 %. In the<br />

experiments with an inflow quantity of 1000 L per run, nearly the complete water amount<br />

reached the end of the tested rill part. Using the RC and the flow length of the LFEs, we could<br />

calculate an average infiltration rate in L m -1 . This value was used to calculate the RC of the<br />

HFEs. The percentages of infiltrated water in the HFEs were between 3 % and 6 % so we<br />

used RCs between 94 % and 97 %. These values are summarised in table 2. The RMEs<br />

describe the variations between the single experiments and an average value. Most RMEs in<br />

45


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

the 9 L min -1 – experiments ranged between 67.5 (75 % quartile) and 42.8 % (25 % quartile),<br />

in the 250 L min -1 – experiments most RMEs were between 51.3 and 29.9 %. The highest<br />

RMEs were 93.1 and 131.4 %, the lowest values were 9.5 and 26.7 %; the medians were 55.8<br />

and 41.3 %. The RCs showed lower variations than the sediment concentrations, most RMEs<br />

were between 48.3 and 13.9 %, the maximum value was 75.4 %, the minimum 0.3 % and the<br />

median 20.8 %. The reason for the higher variations in sediment concentrations were “special<br />

events” such as bank failure and headcut retreat. Such processes increased the sediment<br />

concentration for a certain time period while the runoff stayed about constant.<br />

Figure 2: Results of the rill experiments: Sediment concentrations of the 6 runs of the 3 rill<br />

experiments with an inflow intensity of 9 L min -1 (A), sediment concentrations of the 8 runs<br />

of the 4 rill experiments with an inflow intensity higher than 250 L min -1 (B), runoff<br />

coefficients of all rill experiments (C), relative measurements errors of the sediment<br />

concentrations (separated in experiments with an inflow intensity of 9 L min -1 and an inflow<br />

intensity higher than 250 L min -1 ) and the runoff coefficients (D).<br />

46


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Table 2: Sediment concentrations and runoff coefficients of the rill experiments<br />

Experiment rill sediment runoff<br />

concentration [g L -1 ] coefficient [%]<br />

RE1_2008a<br />

2.10 11.27<br />

1<br />

RE1_2008b 0.82 37.75<br />

RE2_2008a<br />

4.48 45.61<br />

2<br />

RE2_2008b 1.23 56.78<br />

RE3_2008a<br />

3.91 51.56<br />

3<br />

RE3_2008b 1.36 71.54<br />

RE4_2009a<br />

6.30 94.00<br />

1<br />

RE4_2009b 5.00 94.00<br />

RE5_2009a<br />

19.90 97.00<br />

4<br />

RE5_2009b 5.10 97.00<br />

RE6_2009a<br />

6.00 96.00<br />

5<br />

RE6_2009b 6.10 96.00<br />

RE7_2009a<br />

15.50 94.00<br />

1<br />

RE7_2009b 5.00 94.00<br />

3.2 Rainfall Simulations<br />

In this study, the results of 33 rainfall experiments were used; the plots of all simulations are<br />

shown in figure 3. In most cases, the rock fragments were embedded into the soil surface and<br />

the bare soil areas were crusted. Only in some special cases, e.g. after a mechanical<br />

disturbance (goat trampling) and the resulting destruction of the crusts, the rock fragments<br />

were overlying. Most test plots showed a sparse vegetation cover and showed either a crusted<br />

soil surface or a cover with embedded rock fragments. Rainfall experiment 23 showed a dense<br />

vegetation cover, plots with mechanical treatment are 5-7 and 12-15. The influence of these<br />

parameters on runoff and erosion are discussed in Ries et al. [27] and Ries & Marzolff [33]<br />

for example. The percentages of the different surface parameters and the slope values of the<br />

different plots are summarised in table 3.<br />

47


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Figure 3: Surfaces of the used rainfall simulation plots<br />

48


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Table 3: Plot parameters and results of the 33 rainfall simulations: RFC e = Rock fragment<br />

cover embedded, VC = Vegetation cover, BS c = Bare soil crusted, RFC o = Rock fragment<br />

cover overlying, BS u = Bare soil uncrusted, RC = runoff coefficient, SQ = sediment quantity,<br />

SSC = sediment concentration<br />

Id RFCe VC BS c RFC o BS u Slope RC SQ SSC<br />

[%] [%] [%] [%] [%] [°] [%] [g] [g L -1 ]<br />

01 85 14 0 1 0 9 72.0 5.8 1.4<br />

02 1 1 98 0 0 11 53.3 4.9 1.7<br />

03 63 36 0 1 0 15 80.6 5.1 1.1<br />

04 6 35 58 1 0 13 65.9 8.7 2.4<br />

05 1 0 0 11 88 5 94.6 70.8 13.4<br />

06 1 0 0 13 86 5 73.5 52.2 12.7<br />

07 1 0 0 40 59 5 112.6 50.6 8.0<br />

08 25 6 68 1 0 9 51.0 13.2 4.6<br />

09 25 18 56 1 0 11 95.2 57.6 10.8<br />

10 91 0 8 1 0 15 43.2 14.9 6.2<br />

11 99 0 0 1 0 15 66.9 11.3 3.0<br />

12 1 3 0 4 92 15 12.2 4.9 7.2<br />

13 1 2 0 11 86 15 5.6 0.4 1.4<br />

14 1 3 0 4 92 15 53.1 25.0 8.4<br />

15 1 2 0 11 86 15 84.4 49.4 10.4<br />

16 18 0 81 1 0 12 39.5 1.9 0.9<br />

17 9 4 86 1 0 10 47.7 6.9 2.6<br />

18 96 0 3 1 0 16 138.0 98.5 12.7<br />

19 27 15 57 1 0 13 53.4 4.1 1.4<br />

20 60 1 38 1 0 11 54.1 4.5 1.5<br />

21 25 10 64 1 0 16 76.3 12.0 2.8<br />

22 1 17 82 0 0 6 58.2 8.1 2.5<br />

23 1 99 0 0 0 11 42.0 3.6 1.5<br />

24 22 17 60 1 0 6 86.8 22.9 4.7<br />

25 77 16 6 1 0 9 39.5 6.7 3.0<br />

26 1 19 80 0 0 4 79.8 7.2 1.6<br />

27 1 10 89 0 0 1.8 46.6 5.9 2.3<br />

28 96 2 1 1 0 19 58.2 17.9 5.5<br />

29 20 8 71 1 0 12.3 57.0 11.2 3.5<br />

30 0 6 94 0 0 2 61.3 20.0 5.8<br />

31 85 2 12 1 0 8 29.6 13.9 8.3<br />

32 52 0 47 1 0 7.5 68.4 30.9 8.1<br />

33 97 0 2 1 0 17 36.6 4.4 2.2<br />

49


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Figure 4: Results of the rainfall simulations: Runoff Coefficients (A), sediment quantities (B)<br />

and sediment concentrations (C). The limits between extremely high, very high, high, middle,<br />

low and very low are marked. D: Relative Measuring errors of the runoff coefficients (RC),<br />

the sediment quantities (SQ) and the sediment concentrations (SSC) of the rainfall<br />

simulations.<br />

The runoff coefficients, the amounts of eroded material and the sediment concentration of the<br />

experiments were classified in 6 classes from very low (class 0) to extreme high (class 5) as<br />

shown in table 4. Each class was symbolised by a box in a certain colour in figure 7. The first<br />

box represented the runoff coefficient, the second the sediment quantity and the third the<br />

sediment concentration. In a fourth box, the average class value from these three parameters<br />

was calculated. These values describe the hazard class of the tested plot against the called<br />

parameters. The class limits for runoff coefficient and sediment quantity were defined on the<br />

basis of the experience from about 400 rainfall simulations of the working group in semi-arid<br />

landscapes in North Africa and Spain during the past 16 years [e.g. 23, 27, 34]. For the<br />

definition of the sediment concentration class limits, we did not divide the sediment quantity<br />

and the runoff quantity, we defined other limits. The reason for this decision was, that in the<br />

other case, nearly all rainfall simulations would show “very high” or even “extreme high”<br />

sediment concentrations even in cases of low sediment quantities (and low runoff values). So<br />

we defined the limits “independent” of runoff coefficient and sediment quantity.<br />

50


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Table 4: Classification of the rainfall simulations regarding runoff coefficient (RC), sediment<br />

quantity (SQ) and sediment concentration (SSC).<br />

Value RC [%] SQ [g] SSC [g L -1 ] Colour<br />

Extreme high 5 > 75 > 40 > 10<br />

Very high 4 50.1 – 75 8.1 – 40 8.1 – 10<br />

High 3 30.1 – 50 4.1 – 8 6.1 – 8<br />

Middle 2 10.1 – 30 1.1 – 4 4.1 – 6<br />

Low 1 1 – 10 0.1 – 1 2 – 4<br />

Very low 0 < 1 < 0.1 < 2<br />

Regarding Figures 7 and 8: First (left) box: RC class value; second box: SQ class value; third box:<br />

SSC class value; fourth (right) box: average class value<br />

Regarding runoff coefficients (Fig 4 A), a wide range from 5.6 % to 100 % could be<br />

observed. In two cases the maximum RC of 100 % was exceeded with values of 112.6 % and<br />

138 % due to seepage of water from outside the plot ring. We still did not omit them because<br />

91 % of the rainfall simulation experiments showed high (21 %), very high (42 %) or extreme<br />

high (27 %) runoff coefficients. Only 9 % showed middle (6 %) or low (3 %) values and no<br />

RC was very low.<br />

Also 91 % of the sediment quantity values were high (33 %) very high (39 %) or extreme<br />

high (18 %) and also only 9 % showed middle (6 %) or low (3 %) values (Fig 4 B). The range<br />

of sediment quantity was between 0.4 g and 98.5 g.<br />

Sediment concentrations of the rainfall simulation experiments ranged between 0.9 g L -1 and<br />

13.4 g L -1 (Fig 4 C). 33 % of these values were high (9 %), very high (9 %) and extreme (15<br />

%). 67 % were middle (12 %), low (27 %) and very low (27 %).<br />

Most of the RMEs were between 13.5 (25 % quartile) and 40.7 % (75 % quartile) for the<br />

runoff coefficients, between 43.4 and 81.8 % for the sediment quantities and between 43.4<br />

and 71.8 for the sediment concentrations. The sediment quantity showed the highest<br />

maximum RME (394.7 %) and the highest median RME (70.4 %). The runoff values<br />

generally showed lower RME values than the erosion parameters, suggesting that the single<br />

values showed lower differences from an average value.<br />

Regarding runoff coefficient classes, sediment quantity classes and sediment concentration<br />

classes separated following the 3 groups (1) with mechanical treatment, (2) without<br />

mechanical treatment and (3) high vegetation cover, a clear statement could be made: plots<br />

with a vegetation cover of maximal about 1/3 (highest value in these groups: 36 %) and<br />

without mechanical treatment showed average class values of 3.9 (very high) for runoff<br />

coefficient, 3.6 (very high) for sediment quantity and 1.5 (low-middle) for sediment<br />

51


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

concentration. Although absolute values showed differences, we found this an appropriate<br />

method for classification of sub-areas. The mechanical treated plots representing goat trails<br />

also showed “very high” runoff coefficient and sediment quantity values (3.7 and 4.0). The<br />

sediment concentration showed in contrast to the other group also the class “very high” (3.7).<br />

That means, we defined for the test site area without vegetation without goat trampling a very<br />

high RC (50.1 – 75 % runoff under a rainfall event of 40 mm h -1 for 30 minutes), a very high<br />

detachment rate (8.1 – 40 g per 0.28 m², that means 29 – 143 g m -2 under the simulated<br />

rainfall event) and a middle sediment concentration (4.1-6 g L -1 ). The only rainfall simulation<br />

(Id 23) with high vegetation cover (99 %) showed lower class values, the RC was still “high”<br />

(30 – 50 %), the sediment loss could be classified as “middle” (4 – 14 g m -2 ) and the sediment<br />

concentration as “very low” (< 2 g L -1 ).<br />

52


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Table 5: Class values separated following “crusted soil”, “rock fragment cover embedded”<br />

and “rock fragment cover overlying”.<br />

Rainfall Sim. Id RC class SQ class SSC class Average class<br />

Group 1: with mechanical treatment<br />

5 5 5 5 5<br />

6 4 5 5 5<br />

7 5 5 3 4<br />

12 2 3 4 3<br />

13 1 1 0 1<br />

14 4 4 4 4<br />

15 5 5 5 5<br />

Average Group 1 3.7 4.0 3.7 3.9<br />

Group 2: without mechanical treatment<br />

1 4 3 0 2<br />

2 4 3 0 2<br />

3 5 3 0 3<br />

4 4 4 1 3<br />

8 4 4 2 3<br />

9 5 5 5 5<br />

10 3 4 3 3<br />

11 4 4 1 3<br />

18 5 5 5 5<br />

16 3 2 0 2<br />

17 3 3 1 2<br />

19 4 3 0 2<br />

20 4 3 0 2<br />

21 5 4 1 3<br />

22 4 4 1 3<br />

24 5 4 2 4<br />

25 3 3 1 2<br />

26 5 3 0 3<br />

27 3 3 1 2<br />

28 4 4 2 3<br />

29 4 4 1 3<br />

30 4 4 2 3<br />

31 2 4 4 3<br />

32 4 4 4 4<br />

33 3 3 1 2<br />

Average Group 2 3.9 3.6 1.5 2.9<br />

Special Group: Vegetation cover<br />

53


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

23 3 2 0 2<br />

3.3 Mapping, GIS analyses and aerial photography<br />

Figure 5 shows the positions of the 5 tested rills and the positions of 33 rainfall simulations.<br />

The different colours represent the year of accomplishment. Additionally, the outlines of the<br />

test site are marked. Figure 6 includes the mapped rills and their catchment areas; in figure 7<br />

the classification of the rainfall simulations is added. The classification system is presented in<br />

table 4. In figure 8, the soil coverage, divided in “vegetation” and “no vegetation”, is<br />

presented additionally.<br />

Figure 5: Positions of the rainfall simulations and the rill experiments<br />

54


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Figure 6: Additionally: Mapped rills and their catchments<br />

55


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Figure 7: Additionally: Classification of runoff coefficient, sediment quantity, sediment<br />

concentration and average assessment of the rainfall simulations<br />

56


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Figure 8: Additionally: soil surface coverage<br />

Table 6 lists the calculated test site parameters. The total rill length in the study area (10.01<br />

ha) was 1694 m. These rills drained an area of 1.98 ha, which was about 20 % of the whole<br />

test area. The measured rills had an average width of 0.15 m (based on field mapping and<br />

SFAP), so there was an estimated rill area of about 0.025 ha (0.25 % of the study area).<br />

Accordingly, we could calculate the total interrill area at about 9.99 ha (99.75 % of the study<br />

area). The rill density was 169 m ha -1 . In the study area, a total trail length of about 365 m was<br />

mapped, the average width of these trails was about 1 m. That means, an area of 365 m² was<br />

covered with uncrusted soil and/or overlying rock fragments what makes up for only 0.5 % of<br />

the non-vegetated area.<br />

57


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Table 6: Different test site parameters<br />

Parameter<br />

Value<br />

Total rill length L R [m] 1694<br />

Area of the test site A T [ha] 10.01<br />

Catchment area of all rills C R [ha] 1.98<br />

Estimated average rill width W R [m] 0.15<br />

Total rill area A R [ha] 0.025<br />

Interrill area A I [ha] 9.99<br />

Rill amount on total test site [%] 0.25<br />

Interrill amount on total test site [%] 99.75<br />

Rill density D R [m ha -1 ] 169<br />

Rill drainage index I R [%] 20<br />

Trail length [m] 365<br />

Trail area [m²] 365<br />

4 Combination of the results and discussion<br />

To compare both, the results of the rill experiments and the rainfall simulations, we wanted to<br />

calculate the inflow intensity which would reach the rill under a real rainfall event of 40 mm<br />

h -1 , an event simulated in our rainfall experiments. From the inflow intensity we could<br />

estimate the average sediment concentration following figure 9. In this figure, we ignored two<br />

peaks: In RE5_2009a was a technical problem in sampling and in RE7_2009a 600 goats were<br />

crossing the rill before the measurement, so the quantity of lost material was too high. We<br />

therefore assumed in both cases the same sediment concentration as measured in run b of the<br />

experiment.<br />

58


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Figure 9: Relationship between the inflow intensity and the sediment concentration in the rill<br />

experiments<br />

The relationship between runoff and the sediment concentration gave a high significance R² in<br />

our experiments but the data base is still too small for general statements of these correlations.<br />

So in this study we used the relationship only for this test site and only for an approximate<br />

estimation of the sediment concentration, caused by runoff of a 40 mm h -1 rainfall event. The<br />

relationship between runoff and sediment yield is to be found in several studies. Parsons et al.<br />

[35] e.g. measured runoff and eroded material from 8 runoff plots during 10 natural storm<br />

events and described a clear correlation between runoff coefficient and the sediment yield.<br />

A critical point in upscaling the results from rainfall simulator area to catchment area scale<br />

was the well-known scale problem [36]. The plot size used in the rainfall simulations (0.28<br />

m²) is too small for rill initiation. On the other hand, only much longer plot or slope lengths<br />

can cause rill development and – in consequence – initiate soil loss processes and runoff<br />

change. An important condition for the initiation of soil erosion in general is exceeding a<br />

certain threshold determined by soil parameters, such as soil shear strength or critical shear<br />

stress, via hydraulic parameters (shear stress, unit length shear force, stream power). Those<br />

values are used to calculate transport and detachment capacities. Soil erosion can occur as<br />

long as the transport or detachment rate does not exceed the transport or detachment capacity.<br />

Exceeding this threshold will cause sedimentation of the excess material [e.g. 37]. Different<br />

research groups defined different thresholds that separate the interrill erosion from the<br />

59


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

initiation of rill erosion. Emmett [38] found, that flowing water on entirely flat surfaces runs<br />

partially in preferred runoff paths with higher flow velocity, runoff depth and consequently<br />

with different erosional behaviour. Other research groups stated 2-3° [39, 40] or 3-7° [41] as<br />

critical slope for rill initiation. When we assume a critical slope of 3°, only two (6 %) of our<br />

rainfall plots showed lower values, 8 plots (24 %) showed slope values lower than 7°. That<br />

denotes that upscaling the results from the rainfall simulation plots to a greater area will<br />

underestimate the real erosion and runoff values. Another critical value for rill initiation is a<br />

Froude number between 2.4 and 3 [39] or 1+0.0035*D, were D is the median of grain size in<br />

µm [42]. Giménez & Govers [43] and Giménez [44] assume that there is a threshold Froude<br />

number. In laboratory experiments, they found that the flow velocity in natural (rough) rills<br />

was independent of the slope angle because of a feedback between rill bed morphology and<br />

flow parameters. The frequency of macro roughness (steps, pools) increased with slope and,<br />

as a consequence, an increase of the flow velocity was prevented. With increasing slope, the<br />

water accelerated until a Froude number between 1.3 and 1.7 was reached. At this point, an<br />

“hydraulic jump” occurred and a plungepool could develop [44, 45]. The Froude number of<br />

the runoff on our rainfall simulation plots was clearly below 1, so the mentioned critical<br />

values were not reached. In contrast to the cited research groups, Torri et al. [46] and Merz &<br />

Bryan [47] stated, that the Froude number was not applicable to discriminate between rill and<br />

interrill flow because runoff in rills was not implicit supercritical.<br />

A third possible factor influencing the rill initiation is the runoff intensity. Loch & Donnollan<br />

[48] observed rill development on clay soils with a slope angle of 2.3°, where the runoff<br />

reached an intensity of 0.3 to 0.6 L sec -1 . Moss et al. [49] determined a value of 0.6 L sec -1 on<br />

non-cohesive quartz sand with a slope of 0.2 – 0.3°. Loch & Thomas [50] found much higher<br />

values: they determined minimum runoff intensities of more than 3 L sec -1 . In our rainfall<br />

experiments, 100 % runoff was 5.6 L 30 min -1 , that means an average value of 0.003 L sec -1 .<br />

In consequence to the reported critical runoff intensities of the research groups, none of our<br />

rainfall simulation plots was able to develop any rills. Merz [16] doubts the importance about<br />

the runoff as a significant value for rill initiation. In his studies, the runoff was measured at<br />

the lowermost part of a test plot. These values could not describe a reliable average of the<br />

hydraulic situation since the runoff was not constant over the whole plot. Merz [16] assumed<br />

that different critical runoff values in both studies are based on variations in cohesive forces<br />

of the substratum and other soil physical parameters. Therefore, the critical runoff should not<br />

be used as a confirmed parameter for rill development.<br />

60


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

More detailed results of soil and water movement under defined conditions are listed below in<br />

Table 7. The conditions for the precipitation were: 40 mm h -1 rainfall event within 30 min.<br />

The field mapping provided size and position of catchments areas. The classification of the<br />

surface in “vegetation” and “no vegetation” was done by aerial images. We assumed that the<br />

complete runoff from “no vegetation” areas can reach the rills. Rainfall experiments of Cerdà<br />

[51] in south-east Spain on Stipa tenacissima areas showed, that surface runoff and erosion<br />

was negligible in the tussock and quite high in bare areas. So we could calculate the rainfall<br />

quantity that reaches to uncovered surface areas of the rill catchments by multiplying the “novegetation”<br />

area in each catchment with 20 (40 mm h -1 but duration just 30 min). The rainfall<br />

simulations were accomplished on areas with very low vegetation cover and the classification<br />

of the rainfall experiments showed a “very high” (class 4) runoff coefficient between 50 %<br />

and 75 %, independent of the soil surface cover (crusted, overlying or embedded rock<br />

fragments). Because of the range in the runoff classification, we also got a range in the runoffand<br />

erosion parameters in table 7. We could calculate the runoff quantity and the runoff<br />

intensity. Based on the runoff values of our rill experiments, we assumed that there was an<br />

average infiltration rate of 2.5 L m -1 in the rill so we could calculate the water quantity and the<br />

runoff intensity which was able to cause erosion. Figure 7 shows the correlation between the<br />

runoff intensity and the sediment concentration calculated from the results of the rill<br />

experiments. Using the now known runoff intensity in the rill and the equation in figure 7, we<br />

could calculate the sediment concentrations in the rills. Runoff quantity, rill lengths and the<br />

average rill widths were known, so we could calculate the absolute erosion quantity and the<br />

erosion rate per unit area.<br />

Table 7: Runoff and erosion parameters of the rill catchments. The presented values base on<br />

the class limits.<br />

Rill Catchment<br />

Rill<br />

Id.<br />

No- Runoff Rill Runoff<br />

Erosion<br />

Area Vegetation<br />

SSC<br />

vegetation quantity length intensity<br />

rate<br />

[m²] area [m²]<br />

[g L -1 ]<br />

area [m²] [L] [m] [L min -1 ]<br />

[g m -2 ]<br />

1 415 89 326 3256 - 4884 26 106 - 161 3 - 4 34 - 61<br />

2 1346 477 869 8695 - 13042 106 281 - 426 6 - 7 54 - 107<br />

3 87 3 84 843 - 1264 26 26 - 40 3 - 3 23 - 38<br />

4 1852 382 1470 14696 - 22044 121 480 - 725 8 - 11 78 - 160<br />

5 59 23 36 365- 547 16 11 - 17 2 - 2 21 - 34<br />

6 280 17 263 2633 - 3950 51 84 - 127 3 - 4 31 - 54<br />

7 466 63 403 4033 - 6049 32 132 - 199 4 - 5 37 - 68<br />

61


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

8 448 64 384 3839 - 5759 35 125 - 189 4 - 4 36 - 66<br />

9 41 11 30 299 - 449 10 9 - 14 2 - 2 21 - 34<br />

10 61 19 42 424 - 636 16 13 - 20 2 - 2 21 - 35<br />

11 66 25 41 407 - 610 19 12 - 19 2 - 2 21 - 34<br />

12 66 17 49 489 - 733 4 16 - 24 2 - 3 24 - 37<br />

13 43 32 11 110 - 166 5 3 - 5 2 - 2 20 - 32<br />

14 133 21 112 1121 - 1682 24 35 - 54 3 - 3 25 - 41<br />

15 1616 393 1223 12235 - 18352 60 403 - 607 7 - 9 69 - 140<br />

16 622 231 391 3913 - 5870 49 126 - 192 4 - 5 36 - 66<br />

17 692 41 651 6513 - 9770 67 212 - 320 5 - 6 46 - 89<br />

18 81 30 51 513 - 770 18 16 - 24 2 - 3 22 - 35<br />

19 155 17 138 1379 - 2068 17 45 - 68 3 - 3 27 - 44<br />

20 443 11 432 4323 - 6484 21 142 - 214 4 - 5 39 - 71<br />

21 491 0 491 4909 - 7363 31 161 - 243 4 - 5 41 - 76<br />

22 1658 253 1405 14053 - 21079 48 464 - 699 8 - 11 77 - 157<br />

23 255 33 222 2222 - 3333 12 73 - 110 3 - 4 30 - 52<br />

24 177 31 146 1460 - 2189 22 47 - 71 3 - 3 27 - 45<br />

25 408 63 345 3454 - 5182 58 110 - 168 4 - 4 34 - 62<br />

26 1120 345 775 7752 - 11629 62 253 - 382 5 - 7 51 - 100<br />

27 916 232 684 6844 - 10267 38 225 - 339 5 - 6 48 - 93<br />

28 183 57 126 1265 - 1897 15 41 - 62 3 - 3 26 - 44<br />

29 109 28 81 811 - 1216 17 26 - 39 3 - 3 24 - 39<br />

30 579 92 487 4870 - 7305 26 160 - 241 4 - 5 41 - 76<br />

31 49 16 33 334 - 500 8 10 - 16 2 - 2 22 - 35<br />

32 37 10 27 268 - 402 6 8 - 13 2 - 2 22 - 34<br />

33 250 90 160 1598 - 2397 20 52 - 78 3 - 3 27 - 46<br />

34 46 10 36 364 - 545 10 11 - 17 2 - 2 22 - 35<br />

35 195 67 128 1285 - 1927 19 41 - 63 3 - 3 26 - 43<br />

36 237 93 144 1444 - 2166 30 46 - 70 3 - 3 26 - 44<br />

37 151 15 136 1363 - 2045 11 45 - 67 3 - 3 27 - 45<br />

38 117 18 99 986 - 1479 8 32 - 49 3 - 3 26 - 41<br />

39 365 96 269 2689 - 4034 73 84 - 128 3 - 4 30 - 54<br />

40 128 30 98 978 - 1466 24 31 - 47 3 - 3 24 - 40<br />

41 132 26 106 1058 - 1587 34 32 - 50 3 - 3 24 - 40<br />

42 68 10 58 579 - 869 12 18 - 28 2 - 3 23 - 37<br />

43 15 4 11 109 - 164 6 3 - 5 2 - 2 19 - 31<br />

44 224 42 182 1824 - 2736 34 58 - 88 3 - 3 28 - 48<br />

45 259 80 179 1791 - 2686 38 57 - 86 3 - 3 27 - 47<br />

46 188 42 146 1456 - 2183 33 46 - 70 3 - 3 26 - 44<br />

47 171 26 145 1447 - 2171 21 46 - 71 3 - 3 27 - 45<br />

48 79 14 65 649 - 973 9 21 - 32 2 - 3 24 - 38<br />

62


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

49 297 44 253 2534 - 3801 21 83 - 125 3 - 4 31 - 55<br />

50 203 35 168 1675 - 2513 26 54 - 82 3 - 3 27 - 47<br />

51 1641 306 1335 13352 - 20028 187 429 - 652 7 - 10 71 - 146<br />

410 30 4 26 257 - 386 12 8 - 12 2 - 2 20 - 33<br />

It has to be considered, that the rill catchments were only used to calculate the runoff for the<br />

rills, erosion only taking place directly in the rills. That means the rill catchments were on the<br />

“rill site” regarding the runoff but on the “interrill site” concerning the erosion parameters.<br />

Table 8 summarises the results of table 7: the complete interrill areas without gully covered<br />

100079 m², 19750 m² of which were catchments for rills which covered 254 m² resulting from<br />

the total rill length and an average rill width. 21912 m² were covered with vegetation, 78167<br />

m² showed bare soil. The erosion rates in the rills reached values between 1685 and 3018 g m -<br />

2 (under a 40 mm h -1 rainfall event of 30 min), whereas the soil loss in the interrill areas was<br />

only between 29 and 143 g m -2 (20-60 times less). This relationship is confirmed by Morgan<br />

[52], who stated that rill sediment transport exceeded interrill transport by a factor of 40 on an<br />

11° slope. Due to the much larger area of the interrill areas, the total soil loss quantity was<br />

about 5 to 15 times higher than in the rills. The runoff intensity of the water from the interrill<br />

area was about 5 times higher than the runoff intensity of the rill flow. Regarding the runoff<br />

intensity per square meter, the rills’ value exceeded the interrills’ value by the factor 50.<br />

Table 8: Comparison of rill and interrill areas. The erosion values are only valid for the rill<br />

areas, not for the rill catchment areas (specific explanation in the text)<br />

Rills / rill catchments Interrill Area without gully<br />

Area complete [m²] 254 / 19750 100079<br />

Area, Vegetation [m²] X / 4174 21912<br />

Area, no vegetation [m²] X / 15576 78167<br />

Erosion rate [g m -2 ] 1685-3018 / X 29-143<br />

Erosion quantity [kg] 425-762 / X 2267-11178<br />

Runoff intensity [m³ s -1 ] 0.08-0.13 / X 0.43-0.65<br />

Runoff intensity per m² [cm s -1 ] 0.03-0.05 / X 0.0006-0.0008<br />

One ignored point in our study is the gully in the test site. This is acceptable in so far as the<br />

gully is not active. An active gully would certainly raise the runoff and erosion values on the<br />

gully-bottom.<br />

63


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

We probably underestimated the rill erosion values: during a real rainfall event, the water<br />

would fill the rill from upstream and from both sides whereas in the experiments, the water<br />

entered the rill at a single point. The measured sediment concentrations in the rills (maximum<br />

19.9 g L -1 in RE5_2009a) showed relatively low values. In the surrounding area of the<br />

presented test site, we still measured much higher sediment concentrations of up to 438 g L -1<br />

caused by an inflow intensity of only 9 L min -1 [25]. This means, the rill effectiveness could<br />

be even higher than the presented results implicate. Unfortunately, we do not (yet) have<br />

enough data of the surrounding test sites for a comparison of the different areas.<br />

5 Conclusions<br />

The aim of this study was to compare the effectiveness of rill and interrill areas concerning<br />

soil erosion on one abandoned land site in Andalusia. The results are only valid for the tested<br />

area, they cannot easily be adapted to other areas or used as general statement about the<br />

relation between rill- and interrill erosion. We combined the results of rill experiments,<br />

rainfall simulations, field mapping and small scale aerial photographs to get an idea about the<br />

dimension of the soil loss caused by rill- and interrill flow. In our test site, the rills reached<br />

much higher erosion rates than the interrill erosion (20-60 times higher). Because of the much<br />

larger interrill area, the absolute erosion values of the interrill areas were 5-15 times higher<br />

than these of the rill area. The rills in our study area drained 19800 m² of the 100079 m² study<br />

area; this means 0.25 % of the study area were responsible for 20 % of the area providing<br />

runoff. The results clearly proved the importance of the rills as sediment provider as well as<br />

runoff accumulator. The possible influence of the gully in our study area was ignored in this<br />

study, so we cannot state the possibly erosion-activating effect of our applied rainfall intensity<br />

of 40 mm h -1 . We know about the problems and the limitations of upscaling experimental<br />

field assessments but as we only presented ranges or class limits, we assume that the results<br />

remain in realistic dimensions.<br />

Acknowledgement<br />

The research was supported by the 'Internationales Graduiertenzentrum' of Trier University,<br />

by the Deutsche Forschungsgemeinschaft (DFG) project numbers Ri 835/2 and Ri-835/6-1<br />

and by the 'Freundeskreis Trierer Universität e.V.'. Additionally we thank all participants of<br />

the field trips to Andalusia in September 2006, September 2008 and September 2009 who<br />

supported the accomplishment of the experiments. Special thanks go to Dr. A. Keller for the<br />

support in data analysis. Finally, we would like to thank the reviewers for their remarks.<br />

64


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

References<br />

1. Beuselinck, L., Govers, G., Hairsine, P., Sander, G. and Breynaert, M.: The influence<br />

of rainfall on sediment transport by overland flow over areas of net deposition,<br />

Journal of Hydrology, 257, 145-163, 2002.<br />

2. Kuhn, N. and Bryan, R.: Drying, soil surface condition and interrill erosion on two<br />

Ontario soils, Catena, 57, 113-133, 2004.<br />

3. Kuhn, N., Bryan, R.B. and Navar, J.: Seal formation and interrill erosion on a<br />

smectite-rich Kastanozem from NE-Mexico, Catena, 52, 149-169, 2003.<br />

4. Le Bissonnais, Y., Cerdan, O., Lecomte, V., Benkhadra, H., Souchère, V. and<br />

Martin, P.: Variability of soil surface characteristics influencing runoff and interrill<br />

erosion, Catena, 62, 111-124, 2005.<br />

5. Brodie, I. and Rosewell, C.: Theoretical relationships between rainfall intensity and<br />

kinetic energy variants associated with stormwater particle washoff, Journal of<br />

Hydrology, 340(1-2), 40-47, 2007.<br />

6. Bryan, R. B.: Soil erodibility and processes of water erosion on hillslope,<br />

Geomorphology, 32(3-4), 385-415, 2000.<br />

7. Govers, G., Giménez, R. and Van Oost, K.: Rill erosion: Exploring the relationship<br />

between experiments, modelling and field observations, Earth-Science Reviews,<br />

84(3-4), 87-102, 2007.<br />

8. Knapen, A., Poesen, J., Govers, G., Gyssels, G. and Nachtergaele, J.: Resistance of<br />

soils to concentrated flow erosion: A review, Earth-Science Reviews, 80(1-2), 75-<br />

109, 2007.<br />

9. Cerdan, O., Le Bissonnais, Y., Couturier, A., Bourennane, H. and Souchère, V.: Rill<br />

erosion on cultivated hillslopes during two extreme rainfall events in Normandy,<br />

France, Soil and Tillage Research, 67(1), 99-108, 2002.<br />

10. Poesen, J.: Transport of rock fragments by rill flow – a field study. In: R.B. Bryan<br />

(Editor): Rill erosion - process and significance, Catena, Supplement 8, Cremlingen,<br />

35-54, 1987.<br />

11. Oostwoud Wijdenes, D.J., Poesen, J., Vandekerckhove, L. and Ghesquiere, M.:<br />

Spatial distribution of gully head activity and sediment supply along an ephemeral<br />

channel in a Mediterranean environment, Catena, 39, 147-167, 2000.<br />

65


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

12. Vandekerckhove, L., Poesen, J., Oostwoud Wijdenes, D. and De Figueiredo, T.:<br />

Topographical thresholds for ephemeral gully initiation in intensively cultivated areas<br />

of the Mediterranean, Catena, 33, 271-292, 1998.<br />

13. Woodward, D. E. 1999. Method to predict cropland ephemeral gully erosion. Catena,<br />

37(3-4), 393-399.<br />

14. Hancock, G.R., Crawter, D., Fityus, S.G., Chandler, J. and Wells, T.: The<br />

measurement and modelling of rill erosion at angle of repose slopes in mine spoil,<br />

Earth Surface Processes and Landforms 33, 1006-1020, 2008.<br />

15. Ries, J.B.: Landnutzungswandel und Landdegradation in Spanien – Eine Einführung,<br />

in: Marzolff, I., Ries, J.B., De La Riva, J. and Seeger, M. (Eds):<br />

Landnutzungswandel und Landdegradation in Spanien, 11-29, 2003.<br />

16. Merz, W.: Experimentelle Untersuchungen zur Rillenerosion auf landwirtschaftlich<br />

genutzten Böden in Kanada und der Volksrepublik China. Freiburger Geographische<br />

Hefte, Heft 40, Selbstverlag des Institutes für Physische <strong>Geographie</strong> der Albert-<br />

Ludwig-Universität Freiburg i.Br., Freiburg, 147 pp., 1993.<br />

17. Poesen, J., Valentin, C., Nachtergaele, J. and Verstraeten, G.: Gully erosion and<br />

environmental change: importance and research needs, Catena, 50(2–4), 91–133,<br />

2003.<br />

18. Faust, D. and Schmidt, M.: Soil erosion processes and sediment fluxes in a<br />

Mediterranean marl landscape, Campina de Cádiz, SW Spain, Z.Geomorph. N.F. 52<br />

(2), 247-265, 2009.<br />

19. Meyer, L.D., Forster, G.R. and Römkens, M.J.M.: Source of soil eroded by water<br />

from upland slopes, Proc. 1972 Sediment Yield Workshop, US Dept. of Agric.<br />

Sediment. Lab., Oxford, Mississippi, 177-189, 1975.<br />

20. Morgan, R.P.C., Martin, L. and Noble, C.A.: Soil erosion in the United Kingdom: a<br />

Case Study from Mid-Bedfordshire, England, Occasional Paper No.14., 1987.<br />

21. Giménez, R., and Govers, G. 2002. Flow Detachment by Concentrated Flow on<br />

Smooth and Irregular Beds. Soil Sci Soc Am J, 66(5), 1475-1483.<br />

22. Hessel, R. and Jetten, V. 2007. Suitability of transport equations in modelling soil<br />

erosion for a small Loess Plateau catchment. Engineering Geology, 91(1), 56-71.<br />

23. Seeger, M.: Uncertainty of factors determining runoff and erosion processes as<br />

quantified by rainfall simulations, Catena, 71(1), 56-67, 2007.<br />

66


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

24. <strong>Wirtz</strong>, S., Seeger, M. and Ries, J.B.: The rill experiment as a method to approach a<br />

quantification of rill erosion process activity, Zeitschrift für Geomorphologie, 54,1,<br />

47-64, 2010.<br />

25. <strong>Wirtz</strong>, S., Seeger, M. and Ries, J.B.: Field experiments for understanding and<br />

quantification of rill erosion processes, Catena 91, 21-34, 2012.<br />

26. Ministerio de Fomento, Secretaria de Estado de Infraestructuras y Transportes,<br />

Directión General de Carreteras: Máximas illuvias diarias en la espana peninsular,<br />

1999.<br />

27. Ries, J.B., Langer, M. and Rehberg, C.: Experimental investigations on water and<br />

wind erosion on abandoned fields and arable land in the Central Ebro Basin,<br />

Aragón/Spain, Z. Geomorphol. N.F. (Supplement 121), 91– 108, 2000.<br />

28. Iserloh, T., Fister, W., Seeger, M., Willger, H. and Ries, J.B.: A small portable<br />

rainfall simulator for reproducible experiments on soil erosion. Soil and Tillage<br />

Research 124, 131-137, 2012.<br />

29. Ries, J.B., Seeger, M., Iserloh, T., Wistorf, S. and Fister, W.: Calibration of simulated<br />

rainfall characteristics for the study of soil erosion on agricultural land, Soil & Tillage<br />

Research 106, 109-116, 2009.<br />

30. Fister, W., Iserloh, T., Ries, J.B. and Schmidt, R.-G.: Comparison of rainfall<br />

characteristics of a small portable rainfall simulator and a combined portable wind<br />

and rainfall simulator. Zeitschrift für Geomorphologie, 55(3), 109-126, 2011.<br />

31. DIN 1319-1.: Grundlagen der Messtechnik – Teil 1: Grundbegriffe, Deutsches Institut<br />

für Normung e.V., Ausgabe: 1995-01 , Deutsch, 1995.<br />

32. Aber, J. S., Marzolff, I., and Ries, J. B.: Small-Format Aerial Photography:<br />

Principles, techniques and geoscience applications., Elsevier, Amsterdam, 2010.<br />

33. Ries, J. B. and Marzolff, I.: Identification of sediment sources by large-scale aerial<br />

photography taken from a monitoring blimp, Phys. Chem. Earth, 22 (3-4), 295-302,<br />

1997.<br />

34. Ries, J.B.: Methodologies for soil erosion and land degradation assessment in<br />

mediterranean-type ecosystems, Land Degradation & Development, 21, 171-187,<br />

2010.<br />

35. Parsons, A.J., Brazier, R.E., Wainwright, J. and Powell, D.M.: Scale relationships in<br />

hillslope runoff and erosion, Earth Surf. Process. Landforms, 31, 1384-1393, 2006.<br />

67


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

36. Amore, E., Modica, C., Nearing, M.A. and Santoro, V.C.: Scale effect in USLE and<br />

WEPP application for soil erosion computation from three Sicilian basins, Journal of<br />

Hydrology, 293, 100-114, 2004.<br />

37. Scherer, U.: Prozessbasierte Modellierung der Bodenerosion in einer Lösslandschaft,<br />

Ph.D. thesis, Fakultät für Bauingenieur-, Geo- und Umweltwissenschaften,<br />

Universität Fridericiana zu Karlsruhe (TH), Karlsruhe, Germany, 2008.<br />

38. Emmett, W.W.: The hydraulics of overland flow on hillslopes, US Geological Survey<br />

Professional Paper 662-A, United States Government Printing Office, Washington,<br />

1970.<br />

39. Savat, J. and De Ploey, J.: Sheetwash and rill development by surface flow, in:<br />

Badland geomorphology and Piping, Geo Books, Norwich, 1982.<br />

40. Govers, G.: Spatial and temporal variability in rill development processes at the<br />

Huldenberg experimental site, in: Rill erosion, Catena Supplement 8, Catena Verlag,<br />

Braunschweig, 17-34, 1987.<br />

41. Evans, R.: Water erosion in british farmers’ fields, Progress in Physical Geography,<br />

14, 2, 198-219, 1990.<br />

42. Boon, W. and Savat, J.: A Nomogram for the Prediction of Rill Erosion, in: Soil<br />

Conservation: Problems and Prospects, John Wiley & Sons Ltd., Cranfield, Bedford,<br />

303-319, 1981.<br />

43. Giménez, R. and Govers, G.: Interaction between bed roughness and flow hydraulics<br />

in eroding rills, Water resources Research, 37,3, 791-799, 2001.<br />

44. Giménez, R.: The interaction between rill hydraulics, rill geometry, and sediment<br />

detachment: an experimental approach, Proefschrift ingedient tot het behalen van de<br />

graad van Doctor in de Wetenschappen, Katholieke Universiteit Leuven, Faculteit<br />

Wetenschappen, Department Geografie-Geologie, 2003.<br />

45. Giménez, R., Planchon, O., Silvera, N. and Govers, G.: Longitudinal velocity patterns<br />

and bed morphology interaction in a rill, Earth Surface Processes and Landforms, 29,<br />

105-114, 2004.<br />

46. Torri, D, Sfalanga, M. and Chisci, G.: Threshold conditions for incipient rilling, in:<br />

Rill erosion, Catena, Supplement 8, Catena Verlag, Braunschweig, 97-105, 1987.<br />

47. Merz, W. and Bryan, R.B.: Critical conditions for rill initiation on sandy-loamy<br />

Brunisols – laboratory and field experiments in Southern Ontario, Canada, Geoderma,<br />

57, 4, 357-385, 1993.<br />

68


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

48. Loch, R.J. and Donnollan, T.E.: Field rainfall simulator studies on two clay soils of<br />

the Darlington Downs, Queensland: I. The effect of plot length and tillage orientation<br />

on erosion processes and runoff and erosion rates, Australien Journal of Soil<br />

Research, 21, 33-46, 1983.<br />

49. Moss, A.J., Green, P. and Hutka, J.: Small channels: Their experimental formation,<br />

nature, and significance, Earth surface processes and landforms, 7, 401-415, 1982.<br />

50. Loch, R.J. and Thomas, E.C.: Resistance to rill erosion: observations on the efficiency<br />

of rill erosion on a tilled clay soil under simulated rain and run-on water, in: Rill<br />

erosion – processes and significance, Catena Supplement 8, Catena Verlag,<br />

Braunschweig, 71-83, 1987.<br />

51. Cerdà, A.: The effect of patchy distribution of Stipa tenacissima L. on runoff and<br />

erosion, Journal of Arid Environments, 36, 37-51, 1997.<br />

52. Morgan, R.P.C.: Soil erosion in the United Kingdom: field studies in the Silsoe area,<br />

1973-75, National College of Agricultural Engineering Occasional Paper 4, 1977.<br />

69


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Kapitel 3<br />

<strong>Wirtz</strong> et al. (2010): The rill experiment as a method to approach a quantification<br />

of rill erosion process activity. Zeitschrift für Geomorphologie 54 (1), 47-64.<br />

70


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

The rill experiment as a method to approach a quantification of rill erosion process<br />

activity<br />

<strong>Wirtz</strong> S. (1), Seeger M. (1,2) & Ries J.B. (1)<br />

(1) Dep. of Physical Geography, Trier University (2) Dep. of Land Degradation and<br />

Development, Wageningen University<br />

with 9 figures and 6 tables<br />

Summary<br />

Within this paper a standardized method to quantify sediment transport and runoff in natural<br />

rills is described. In order to achieve this, several rill experiments (RE) were accomplished in<br />

March 2007 in the Arnás catchment in the Spanish Pyrenees. Both, anthropogenically<br />

initiated and naturally developed rills were flushed with a total water quantity of 72 l in 8<br />

minutes (equivalent to 9 l min -1 ). Following the developed experimental routine flow<br />

velocities are recorded along the whole flushed rill, sediment concentrations are measured at<br />

different points and different times during the experiment and runoff values are measured<br />

after 25 m. For the characterisation of the rill, slope is measured and micromorphological<br />

features like scours are registered.<br />

By means of the developed rill experiments it becomes possible to assess the effectivity of<br />

single rills in a catchment. However, the assessments can not be generalized because it is<br />

impossible to project the behaviour of one rill to other rills, even if the spatial distance in<br />

between is small. But it gets possible to detect the qualitative process activity in a rill.<br />

The average slopes oscillated between 7.6° and 11.3°, the highest slope was 16°. The<br />

sediment concentrations reached average values between 0.69 and 2.21 g l -1 , the maximum<br />

values ranged between 1.59 and 6.31 g l -1 . Comparing the sediment concentrations measured<br />

in the rills to the sediment concentrations in the runoff of the river Arnás, it can be stated that<br />

the concentrations in the rills are usually higher.<br />

Keywords: soil erosion, rill erosion, field experiments, quantification of process activity,<br />

Arnás catchment<br />

71


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

1 Introduction<br />

Soil erosion is a complex process, which involves different processes at variable spatial and<br />

temporal scales. Rill erosion is considered to be the most important process affecting soils.<br />

Not only because of the large amounts of eroded soil material, but also due to the generation<br />

of more or less persisting forms, which can develop into gullies and disable further land use.<br />

(OOSTWOUD WIJDENES et al. 2000, VANDEKERCKHOVE et al. 1998, WOODWARD<br />

1999, HANCOCK et al. 2008)<br />

The outstanding role of linear erosion in geomorphological, hydrological and economical<br />

aspects can be explained by the fact that the runoff water reaches its maximal power as both,<br />

eroding and transporting medium, when concentrated in a rill (MERZ 1993). Runoff depth<br />

and flow velocity reach higher values than within diffuse sheet flow, thus the forces affecting<br />

soil particles are multiplied as compared to sheet flow (POESEN 1987). Consequently, more<br />

energy for mobilization and transport of soil material is available (MERZ 1993) leading to a<br />

clear increment in erosion rates. MEYER et al. (1975) noticed a triplication of the erosion<br />

rates due to rill development. Corn fields in Bedfordshire, England for example produce 9-21<br />

times as much sediment with rill erosion under corn as a comparable field only affected by<br />

sheet erosion (MORGAN et al. 1987). According to CERDAN et al. (2002) up to 90 % of soil<br />

loss in several heavy rainfall-events in Normandy was caused by rill erosion.<br />

Despite the efforts made for understanding rill erosion processes and building models to<br />

explain the retreat of headcuts (e. g. FLORES-CERVANTES et al. 2006) or to model gully<br />

erosion by identification of stochastic components (SIDORCHUK 2005) there is still much<br />

work to do for a comprehensive modelling of rill erosion and to the development of field<br />

methods for understanding and quantifying the rill erosion processes. The lack of rill erosion<br />

models is all that weighty, if is compared with the approaches for modelling sheet erosion<br />

(BRUNO et al. 2008).<br />

BRUNO et al. (2008) tested the evolution of a rill network under natural rainfall on testplots<br />

in Sicily. Investigating the effect of initial conditions, rill headcut development and the most<br />

important hydraulic variables, they tried to compile simple mathematical models. Their results<br />

are remarkable, because they demonstrate that the erosion occurs within the upper part of the<br />

rill, but the eroded material is only passed through the lower rill section.<br />

Correct and reproducible measurement of rill erosion processes are needed for the<br />

understanding and the correct model building of a rill erosion model. CASALI et al. (2006)<br />

tested three frequently used methods to quantify the intensity of rill erosion processes and<br />

showed that different methods offer clear differences in resulting data. Depending on the<br />

72


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

method used, the form of the tested rill, as well as position and distance between the measured<br />

cross sections, the measuring error was up to 40 %. So it is clear, that results taken with<br />

different methods still cannot be compared and a standardized method is needed. These<br />

methods are useful to quantify erosion, but it remains impossible to give a statement about the<br />

dynamics in the rill.<br />

Still not all factors influencing the evolution and the behaviour of rills are known, and the<br />

knowledge of the relationship between factors is weak as well. KNAPEN et al. (2007)<br />

attribute this to the lack of comparable data that can be used for a comparative investigation<br />

of different studies on rill erosion. Furthermore, there is still little known about the function of<br />

rills in specific environments, this is: what is the rill's contribution to the sediment dynamics?<br />

How effective are rills for the routing of surface runoff?<br />

In the present paper, a field method is presented for the accomplishment of comparable<br />

experiments on ephemeral rill types in different landscapes. Hereby, a contribution to<br />

understand the functioning of natural rills and their role concerning runoff concentration and<br />

erosion is made. Additionally, the interconnection of the rills with their environment,<br />

especially the contributing areas, is seized by means of rainfall simulations. Some of the<br />

possibilities will be shown exemplarily on experiments performed in the Central Spanish<br />

Pyrenees.<br />

2 Material and Methods<br />

Study area:<br />

The Arnás catchment is located in the Upper Aragón River Basin, a northern tributary of the<br />

Ebro River. The bedrock is Eocene Flysch with alternating sandstones with carbonate<br />

cementation and marl layers sloping northward, which is characteristic of a wide sector of the<br />

Central Spanish Pyrenees. The climate is mountainous Mediterranean sub-humid. The<br />

average annual precipitation is about 1100 mm, mostly concentrated from October to May but<br />

divided by a secondary minimum in March. The average annual temperature is 10ºC. The<br />

Arnás ravine drains a 284 ha headwater catchment into the Lubierre River, a small tributary of<br />

the Aragón River. The highest peak is at 1330 m a.s.l. and the outflow at about 900 m a.s.l.<br />

The ravine runs from west to east, building up a valley with a strong contrast between the<br />

steep south facing slope and the gentle north facing slope. The morphology of the slopes is<br />

characterised by big rotational landslides and earthflows. Some poorly drained areas can be<br />

found related to the rotational landslides, especially in the shady aspect. They are mostly<br />

disconnected from the drainage network. Most of the catchment is covered by shrubs<br />

73


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

(SEEGER et al. 2005). Due to sheep grazing, the vegetation succession is strongly retarded<br />

and absent on many trails (RIES et al. 2000b; RIES et al. 2003).<br />

Figure 1 shows the Arnás catchment with the dominating runoff processes in winter, when the<br />

experiments were carried out and the positions of the rill experiments. The tested rills have<br />

developed in regions with dominating surface runoff generation processes: the first rill in a<br />

Hortonian overland flow (HOF) area, rill 2 and rill 3 in an saturation overland flow (SOF)<br />

area.<br />

Fig. 1: Overview over the Arnás catchment with the dominating runoff processes in winter:<br />

HOF= Horton overland flow, SOF=saturation overland flow, SSF= subsurface flow, DP=<br />

deep percolation (modified according to BUTZEN et al. 2009 submitted).<br />

Tested rills:<br />

In the Arnás catchment the experiment was accomplished in three rills that are very similar to<br />

each other for the first 12 m. The slope stays constant at about 8° in rill 1 whereas rill 2 and<br />

rill 3 become steeper. (Fig. 3). All three rills drain over a track and finish with a step at the<br />

track scarp.<br />

The first rill has developed along a motorcycle trail, thus it is an anthropogenically initiated<br />

rill with an average slope of 7.6°, the maximum slope is 12°. The rill is weakly developed, the<br />

width is between 1 and 10 cm, the depth between 1 and 5 cm. The rill is nearly completely<br />

74


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

covered with grass. The rock fragment cover is about 10%. The dominating runoff process<br />

mapped for this rill is HOF.<br />

The second rill is a naturally developed rill on an abandoned field with an average slope of<br />

11° and a maximum slope of 16°. This medium developed rill (width between 5 and 20cm,<br />

depth between 5 and 10 cm) shows a dense rock fragment cover (about 30%) and is only<br />

partially (about 10%) grass-covered. The dominating runoff process is the SOF.<br />

Rill 3 is also a naturally developed rill, that has cut into an old slump. The average slope is<br />

11.3° and the maximum slope 14.5°. The rill is rather deeply incised with width between 10<br />

and 50 cm and depth between 15 and 20 cm. It shows a dense rock fragment cover of about<br />

30% and almost no vegetation cover (about 5%). Rill 3 was just tested over 15 m, because it<br />

was expected that the water would not reach the 25 m standard test length on the slump<br />

material. So, also the slope was just measured over the 15 m. (Fig. 3) The dominating runoff<br />

process is also SOF.<br />

Fig. 2: Profiles of the 3 rills. The measuring points are also marked.<br />

75


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Fig. 3: The three tested rills.<br />

Table 1: Rill parameters of the three tested rills<br />

Factor RE 1 RE 2 RE 3<br />

Tested length [m] 25 25 15<br />

Average slope [°] 7.6 11 11.3<br />

Maximum slope [°] 12 16 14.5<br />

Catchment area [m²] 450 400 ND<br />

Rock fragment cover [%] 10 30 30<br />

Vegetation Cover [%] 90 10 5<br />

depth [cm] 1-5 5-10 15-20<br />

width [cm] 1-10 5-20 10-50<br />

runoff process HOF SOF SOF<br />

Rainfall simulations:<br />

For the characterisation of the rill catchments there were performed rainfall simulations, with<br />

a small mobile rainfall simulator based on the one used by CALVO CASES et al. (1988),<br />

LASANTA et al. (1994) and CERDÁ & GARCÍA-FAYOS (1994-95). The plot is bounded<br />

by a steel ring of 60 cm in diameter (0,28 m²) and 6 cm height, it has an outlet of 10 cm width<br />

with a narrowing runoff collector (RIES et al. 2000a, PFAHLS et al. 1999). The rainfall<br />

intensity is 40 mm/h, the fall height is 2 m. The suitability of this method for the explanation<br />

of the influence of external factors on runoff generation and erosion with this methodology<br />

was demonstrated in the same catchment by SEEGER (2007).<br />

76


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Rill experiment:<br />

The objective of the rill experiment is to simulate concentrated runoff without the effect of<br />

splash or additional inflow into a rill. Each experiment consists of two runs: first the rill is<br />

tested under dry conditions; in a second run, about 15 min later, the same rill is tested under<br />

wet conditions.<br />

With a motor driven pump, a constant discharge of 9 l min -1 is maintained during 8 minutes,<br />

resulting in a total water inflow of 72 l.<br />

In order to reduce the flow velocity of the inflowing water and consequently avoiding the<br />

mobilization of soil material at the starting point, the water inflow is initiated through the<br />

“Field-Adapter-Tool A” (FAT-A) (Fig.4). This is a longish 55x30 cm plastic plate on which<br />

plastic pipes and an inverted meadow of synthetic turf are fixed. The hose can be adapted to<br />

the pipe system with a compressed air adapter. The flow velocity is firstly reduced by the<br />

course of the pipes and secondly by the flow path under the artificial turf. In this way,<br />

accelerated erosion can be avoided, as the water leaves the plastic plate through a cut and<br />

flows into the rill in a rather 'natural' speed.<br />

Fig.4: Device for water inflow into the rill (Fat-A).<br />

The travel time of the waterfront and of two colour tracers are measured for each meter using<br />

a chronograph. The first colour tracer is induced after three minutes at the starting point, the<br />

second follows after six minutes. By means of this procedure, three velocity curves are<br />

recorded and changes in flow dynamics can be detected. As colour tracers, food colourings (E<br />

124 (red) and E 13 (blue)) are used for reasons of safety.<br />

The experiment ends after 25 m flow length. For the continuous measurement of the water<br />

level by means of a pressure transducer (Ecotech DL/n, V2.35), the FAT-Z (Fig. 5) has been<br />

designed. It consists of a metal sheet with a plastic drainpipe with a junction. The sheet is<br />

driven into the ground as far until the opening of the drainpipe is on rill’s ground level. The<br />

77


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

pressure transducer is installed at the junction piece, and this is filled with water so that the<br />

water level sensor is permanently awashed. For calibrating the gauge, the runoff at the<br />

outflow is measured volumetrically in regular intervals.<br />

Fig. 5: Mobile flume (FAT-Z).<br />

The rills slope is measured with a spring bow of 1m range and a digital air lever (Fig. 6), thus<br />

dividing it into 25 segments. It has to be considered, that the slope measuring provides only<br />

average slopes for 1 meter. A step or a knickpoint in the rill is not accounted. The positions<br />

and the heights of the steps are noted.<br />

78


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Fig. 6: The 1m-spring bow.<br />

For gathering intermediate data on suspended sediment transport, three adequate sampling<br />

points are selected. Small knickpoints in the rill have proven to be adequate to collect water<br />

samples without having to press the bottle into the rill bottom. At these points in the rills<br />

longitudinal profile, four water samples are taken: the first directly after the waterfront has<br />

reached the sampling point, the second after 30 seconds, the third after 1:30 min and the<br />

79


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

fourth 2:30 min after the arrival of the water. The sampling disturbs flow behaviour, so it<br />

should be taken quickly. There is no default quantity, but the sample should be enough for<br />

filtration and determination of sediment concentration.<br />

Laboratory analysis and data processing:<br />

The samples are filtered with a 25 µm and a 2 µm pore size filter from Schleicher & Schuell<br />

Company (Whatman 589/1 black ribbon and Whatman 589/3 black ribbon) and the sediment<br />

quantity is determined gravimetrically. The calculated sediment amounts are related to the<br />

water quantity in the sample to get a sediment concentration in g l -1 .<br />

Data evaluation:<br />

The experiments are characterised by means of the following indicators, where the water<br />

flowing out of the rill system is considered as runoff:<br />

intensity factor: the relationship between the maximum runoff value and the inflow<br />

intensity, I = RI / II<br />

duration factor: the relationship between the runoff duration and the inflow duration,<br />

DF = RD / ID<br />

detention: the difference between the runoff start and the inflow finish, D = RS - IF<br />

time factor: the relationship between these two values, T = RS / IF<br />

runoff rate: relationship between runoff quantity and inflow quantity. This factor<br />

neglects the influence of the flow length, but it is an indicator to the infiltration<br />

characteristics within the rill, R = RQ / IQ<br />

r-l-factor (runoff-length-factor): the relationship between the runoff- and the inflow<br />

quantity, depending on the flow length: r-l-factor = (runoff quantity / inflow quantity )<br />

* length [m], RL = (RQ / IQ ) * L<br />

For characterising the rainfall simulations, we used the overall runoff coefficient. With this,<br />

the size of the contributing area is determined, which is needed to produce a runoff of 9 l min -<br />

1 at the starting point of each rill with a rainfall intensity of 40 mm h -1 . With a given<br />

catchment area size and a known runoff coefficient of the rainfall simulations, the RR (Rill<br />

runoff caused by a rainfall intensity of 40 mm h -1 in the catchment [l s -1 ] ) can be calculated,<br />

at least the dimension.<br />

The runoff of a river catchment can not easily be compared to a runoff in a rill. In the Arnás<br />

catchment, runoff and sediment concentration values were measured (SEEGER et al. 2005).<br />

We can calculate the dimension of the theoretical runoff, a runoff quantity that the Arnás river<br />

80


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

would offer if he had not his real catchment area size but the catchment area size of the rills.<br />

This value is calculated as follow:<br />

runoff Arnás * area rill catchment<br />

Theoretical runoff =<br />

area Arnás catchment<br />

It is not possible to compare the absolute runoff values of a river and a rill catchment, so we<br />

calculated the specific runoff, the theoretical runoff relating to 1 ha for the Arnás catchment<br />

and the rill catchments.<br />

For the Arnás catchment, the specific average runoff is calculated as follow:<br />

average runoff min average runoff max<br />

Specific average runoff = / 2<br />

catchment area catchment area<br />

The specific runoff peak is calculated respectively.<br />

The values for the rill catchments are calculated as follow:<br />

RQ run a RQ run b<br />

Specific average runoff =<br />

/ (2 * catchment area)<br />

RD run a RD run b<br />

Specific runoff peak =<br />

RI run a RI run b<br />

2 * catchment area<br />

with RQ = runoff quantity [l], RD = runoff duration [s] and RI = maximum runoff intensity [l<br />

s -1 ].<br />

3 Results<br />

Table 2: Results of the rainfall simulations: ND = No Data<br />

Factor RE 1 RE 2 RE 3<br />

Sed.Conc. rainfall simulation 1 [g l -1 ] 0.7 0.9 ND<br />

Sed.Conc. rainfall simulation 2 [g l -1 ] ND 1.86 ND<br />

Runoff of rainfall simulation 1 [%] 25 80 ND<br />

Runoff of rainfall simulation 2 [%] ND 100 ND<br />

Start runoff rainfall simulation 1 [s] 99 61 ND<br />

Start runoff rainfall simulation 2 [s] ND 14 ND<br />

RR = Rill runoff caused by a rainfall intensity of 40 mm h -1 in the<br />

catchment [l s -1 ]<br />

1.3 4.15 ND<br />

In the 450m² catchment area of rill 1, one rainfall simulation was accomplished. This<br />

simulation showed a runoff coefficient of 25% and a average sediment concentration of 0.7 g<br />

l -1 . The runoff started after 99 s. The RR reaches 1.3 l s -1 .<br />

81


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

The catchment of rill 2 has an area of 400m², in this area two rainfall simulations were<br />

accomplished. The runoff coefficients were 80% and 100%, the average sediment<br />

concentrations 0.9 and 1.86 g l -1 . The runoff started after 61 s in the first simulation and still<br />

after 14 s in the second simulation. The RR reaches 4.15 l s -1 .<br />

In the catchment area of rill 3, no rainfall simulation was accomplished .<br />

Table 3: Overview over the runoff values of the rill experiments: n.D. = no Data<br />

rill 1 rill 2 rill 3<br />

Factor 1a 1b 2a 2b 3a 3b<br />

L = length [m] 25 25 25 25 15 15<br />

IQ = inflow quantity [l] 72 72 72 72 72 72<br />

II = inflow intensity [l s -1 ] 0.15 0.15 0.15 0.15 0.15 0.15<br />

IF = inflow finish [s] 480 480 480 480 480 480<br />

ID = inflow duration [s] 480 480 480 480 480 480<br />

RQ = runoff quantity [l] 8.41 17.01 20 49.23 n.D. 10.09<br />

RI = maximum runoff intensity [l s -1 ] 0.042 0.059 0.066 0.102 n.D. 0.052<br />

RS = runoff start [s] 554 398 608 414 n.D. 442<br />

RF = runoff finish [s] 904 832 1420 1612 n.D. 796<br />

RD = runoff duration [s] 350 434 812 1198 n.D. 354<br />

I = intensity factor 0.28 0.39 0.44 0.68 n.D. 0.34<br />

DF = duration factor 0.73 0.90 1.69 2.5 n.D. 0.74<br />

D = detention [s] 74 -82 128 -66 n.D. -38<br />

T = time factor 1.15 0.83 1.27 0.86 n.D. 0.92<br />

R = runoff rate [%] 11.68 23.63 27.78 68.38 n.D. 14.01<br />

RL = r-l-factor 2.92 5.91 6.94 17.09 n.D. 2.1<br />

82


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Fig. 7: Runoff values after 25m, respectively. 15m for rill experiment 3.<br />

In run A of RE 1 8.41 l of the 72 l inflow reached the measuring point at 15 m this is<br />

equivalent to a runoff rate of 11.68 %. The maximum runoff at the outlet is 0.042 l s -1 that<br />

means 28% of the inflow intensity. The runoff period counts 350 s and it begins 74 s after<br />

stopping inflow. In the second run, 17.01 l reached the 25 m with a maximum intensity of<br />

0,059 l s -1 , (39% of the inflow intensity) corresponding to a runoff rate of 23.63 %. The<br />

duration is longer in the second run (434 sec) and starts earlier, 82 s before stopping<br />

discharge. The r-l-factor in the first run is 2.92, in the second run 5.91. (Fig. 7a)<br />

83


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

The first run of RE 2 leads to a total runoff of 20 l, thus representing a proportion of 27.78 %,<br />

at the end of the rill. The runoff peak reaches 0.066 l s -1 , 44% of the inflow intensity. During<br />

the second run, 49.23 l reach the gauge at 25 m, indicating a runoff rate of 68.38%. Also the<br />

runoff intensity is higher than in the other experiments, the maximum value reaches 0.102 l s -1<br />

(68% of the inflow intensity). In the first run the runoff lasts for 812 s and starts 128 s after<br />

the water input is stopped, in the second run, 66 s before stopping the discharge, the runoff<br />

duration of 1198 s starts. The r-l-factor is 6.94 for the first run and 17.09 for the second run.<br />

(Fig. 7b)<br />

In RE 3A the water did not reach the measuring point after 15 m. In the second run (RE 3B)<br />

10.09 l (14.01%) reached the measuring point with a maximum intensity of 0.052 l s -1 (34%<br />

of the inflow intensity). The runoff starts 38 s before the pump is stopped and goes on for 354<br />

s The r-l-factor is calculated here with 2.1.<br />

Table 4: Overview over the flow velocities of the rill experiments: n.D.= no data<br />

rill 1 rill 2 rill 3<br />

Factor 1a 1b 2a 2b 3a 3b<br />

MEAN water front [m s -1 ] 0.05 0.04 0.05 0.07 0.02 0.03<br />

MEAN Tracer 1 [m s -1 ] 0.1 0.11 0.09 0.1 0.06 0.07<br />

MEAN Tracer 2 [m s -1 ] 0.1 0.1 0.17 0.13 0.05 0.08<br />

MAX waterfront [m s -1 ] 0.12 n.D. 0.13 0.11 0.04 0.06<br />

MAX Tracer 1 [m s -1 ] 0.23 0.24 0.52 0.38 0.25 0.27<br />

MAX Tracer 2 [m s -1 ] 0.2 0.24 1.35 1.09 0.11 0.28<br />

84


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Fig. 8: Flow velocities of the 3 rill experiments.<br />

In the first experiment it is plain to see that the waterfront flows much slower than the two<br />

tracers. The front is even outrun by tracer 1 after 22 m. Run a and run b also show similar<br />

average and maximum velocity values. In the first run, the waterfront shows flow velocities<br />

between 0.02 and 0.12 m s -1 , the average is 0.05 m s -1 , in the second run the flow velocity<br />

could not be measured for the first 20 m. Tracer 1 is similar in the two runs, the velocities<br />

range between 0.23 and 0.02 m s -1 in the first run and between 0.24 and 0.04 m s -1 in the<br />

second run. The averages are 0.10 and 0.11 m s -1 respectively. Tracer 2 is very similar to<br />

Tracer 1, the values are between 0.2 and 0.04 m s -1 in the first, and between 0.24 and 0.03 m<br />

s -1 in the second run. The averages are both 0.10 m s -1 .<br />

In experiment 2 are no big differences between the waterfront and the two tracers. In both<br />

runs the highest flow velocities are reached over the first 3-4 m, at this place also the highest<br />

differences in velocity between the waterfront and the tracers can be observed. The course is<br />

very constant. The flow velocities of the waterfront range between 0.13 and 0.02 m s -1 in the<br />

first run and between 0.11 and 0.03 m s -1 in the second run. The averages are 0.05 and 0.07 m<br />

85


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

s -1 . Tracer 1 shows a higher maximum value in the first run (0.52 m s -1 ) than in the second run<br />

(0.38 m s -1 ), whereas the mean values are similar reaching 0.09 m s -1 in the first and 0.1 m s -1<br />

in the second run. The maximum velocity value of tracer 2 is measured in the first run with<br />

1.35 m s -1 compared to 1.09 m s -1 in the second run. The average is also higher in the first run<br />

with 0.17 m s -1 against 0.13 m s -1 in the second run. The minimum velocities are 0.03 m s -1 in<br />

the second run for the waterfront, tracer 1 and also tracer 2; in the first run, the minimum of<br />

front and first tracer is 0.02 m s -1 , the lowest flow velocity of tracer 2 is 0.04 m s -1 .<br />

In experiment 3 variations can be detected as well but they are not as high as in the first<br />

experiment. The water does not reach the rill's end in the first run but the movement stops<br />

after 11m. The waterfront shows again the lowest flow velocities and is cached by tracer 1 in<br />

the first run after 6.5 m, in the second run after 13 m. The two tracers show again very similar<br />

flow velocities, until tracer 1 outruns the waterfront and turns into the waterfront itself. The<br />

maximum velocity of the waterfront is 0.04 m s -1 in the first run and 0.06 m s -1 in the second<br />

run. The averages are 0.02 m s -1 and 0.03 m s -1 respectively. Tracer 1 reaches 0.25 m s -1 in the<br />

first run and 0.27 m s -1 in the second run, the average is 0.06 m s -1 and 0.07 m s -1 . Also tracer 2<br />

shows a higher maximum in the second run with 0.28 m s -1 against 0.11 m s -1 in the first run<br />

and mean values of 0.05 m s -1 in the first and 0.08 m s -1 in the second run. The minimum<br />

values are 0.01 m s -1 for the front, tracer 1 and tracer 2 in both runs.<br />

Table 5: Overview over the sediment concentrations of the rill experiments<br />

rill 1 rill 2 rill 3<br />

Factor 1a 1b 2a 2b 3a 3b<br />

MAX [g l -1 ] 2.06 2.31 2.37 6.31 2.31 1.59<br />

MIN [g l -1 ] 0 0.05 0 0.3 0 0<br />

MEAN [g l -1 ] 0.78 0.69 0.82 2.21 0.83 0.88<br />

86


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Fig.9: Sediment concentrations of the 3 rill experiments: MP=Measuring point, n.D.= no data.<br />

In the first experiment, the maximum sediment concentrations are 2,06 g l -1 in the first run<br />

and 2.31 g l -1 in the second run. Minimum values are 0 g l -1 (no sediment detected) in the first<br />

run and 0.05 g l -1 in the second run. The averages are 0.78 g l -1 and 0.69 g l -1 .<br />

The second experiment shows a maximum sediment concentration of 2.37 g l -1 in the first and<br />

6.31 g l -1 in the second run, the minimum values are 0 g l -1 for the first run and 0.3 g l -1 for the<br />

second run. The averages are 0.82 and 2.21 g l -1 .<br />

In the third experiment, the last measuring point is not reached by the water in the first run, so<br />

no samples could be collected there. regarding the first two measuring points, the first run<br />

shows a maximum of 2.31 g l -1 , a minimum of 0 g l -1 and an average of 0.83 g l -1 . In the<br />

second run, all three measuring points offered samples. The maximum value is 1.59 g l -1 , the<br />

minimum 0 g l -1 and the average 0.88 g l -1 .<br />

Concerning the sediment concentrations, high differences are to notice. The highest value is<br />

reached with 6.31 g l -1 in experiment 2b. But there are also several samples where no<br />

sediment could be detected. The averages of the experiments mostly range below 1 g l -1 that is<br />

87


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

even slightly below the sediment concentrations of the rainfall simulations in the rill<br />

catchments. Experiment 2b is an exception with several samples reaching values between 2<br />

and 4 g l -1 , and a maximum value of over 6 g l -1 .<br />

4 Discussion<br />

POLYAKOV & NEARING (2003) tested in laboratory experiments the behaviour of rills<br />

under different inflow quantities and rill lengths on a loamy substrate. The sediment<br />

concentrations reached values between 10 and 130 g l -1 . LOCH (2000) tested the influence of<br />

vegetation on soil erosion. In one of his plots, both interrill and rill erosion was observed. The<br />

sediment concentrations of this plot reached a maximum value of 210 g l -1 . In the experiments<br />

accomplished in the Arnás-catchment, the maximum sediment concentration reached only<br />

6.31 g l -1 . The low sediment concentrations may be explained by the fact that the induced<br />

water inflow rate is not sufficient to activate a further incision of the rill. Also the take-up of<br />

loose material from the rill's ground does not raise the sediment concentrations significantly<br />

because directly before the experiments, snow melt removed the available material. The<br />

necessary presuppositions for further rill incision are a higher discharge rate and/or the<br />

preparation of transportable material e.g. by frost weathering.<br />

For all rill experiments the total runoff is higher in the second run, in experiment 2b for<br />

example 49 l of the 72 l inflow reach the end of the rill at 25 m, this is a rate of about 68 %.<br />

The lowest runoff rate (11,68%) reaches the 25 m in experiment 1a. The third experiment is<br />

not directly comparable to the other two experiments because of the changed flow lengths.<br />

After 15 m, there is certainly still a higher water quantity in the rill as after 25 m. The rate<br />

reaches 14 % here. Regarding the r-l-factor, experiment 3 shows the lowest value, it is even<br />

still lower than the r-l-factor of experiment 1a. In experiment 3, the runoff was measured only<br />

in run b, because in run a, over a flow length of 15m, 100% of the 72 l of inflow could<br />

infiltrate into the substrate.<br />

For each rill experiment, the second run shows a negative detention, indicating that the water<br />

reaches the measuring point before the inflow is finished. The lower the detention is, the<br />

faster the water reaches the measuring point. The second runs show higher maximum runoff<br />

quantities and so also higher intensity factors, the earlier runoff starts and the longer the<br />

period. Here we can deduce, that the moisture status of the rill not only has an influence on<br />

the infiltration within the rill, but also has a positive relation with the flow velocity.<br />

The three tested rills show different starting and developing conditions. The first rill is<br />

anthropogenically initiated and has developed on a motorcycle trail, the two other rills are<br />

88


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

naturally developed. Rill 3 has developed on an old rotational slump and is the deepest incised<br />

of the three rills. The deep incision suggests a more intense activity of the rill, but in the<br />

experiments, the used 72l water did not reach the rill’s end. That means, that for development<br />

and evolution of this rill, a much higher water quantity is needed as used in the experiments.<br />

The estimated area of the catchments of rill 1 and rill 2 was 450 and 400 m² respectively. In<br />

the rill catchments rainfall simulations were accomplished and the results are used to get a<br />

ballpark estimation of the water amounts that can reach the rill at least for the given soil<br />

moisture situation and the experimental rainfall intensity of 40 mm h -1 . Under a rainfall<br />

intensity of 40 mm h -1 , rill 1 can expect a runoff of about 1.3 l s -1 , rill 2 even 4.15 l s -1 , these<br />

values are about 10 to 25 times higher than the used input discharge rate of 0.15 l s -1 for each<br />

rill. The rainfall simulation in the catchment of rill 1 reached a sediment concentration of 0.7<br />

g l -1 , the average of the two runs of RE1 was 0.73 g l -1 , so there is no clear difference between<br />

rill erosion and sheet erosion in this case, because the used quantity of water was not<br />

sufficient to lead to further rill incision. The two rainfall simulations in the catchment of rill 2<br />

reached an average sediment concentration of 1.38 g l -1 , the two runs of RE2 reached an<br />

average of 1.52 g l -1 .<br />

At the outlet of the Arnás catchment, SEEGER et al. (2005) measured runoff values as well as<br />

sediment concentrations. The average sediment concentrations ranged between 0.03 and 2.29<br />

g l -1 , with maximum values between 0.09 and 4.8 g l -1 . In the rill experiments, the maximum<br />

sediment concentrations ranged between 1.59 and 6.31 g l -1 thus being in the same range than<br />

the values of the whole Arnás catchment. It is not possible to compare the runoff values of the<br />

rill experiments with the runoff values of the catchment because rill and river are different<br />

forms and there are different processes proceeding. What we can do is to compare the runoff<br />

values of the rill's catchment with the Arnás catchment: The area of the Arnás catchment is<br />

284 ha, the average runoff values are between 48 and 708 l s -1 . For a catchment of 450 m² (rill<br />

1) and 400 m² (rill 2) the theoretical runoff values would reach values between 0.008 and 0.11<br />

l s -1 and between 0.007 and 0.1 l s -1 . The maximum values between 87 and 2347 l s -1 of the<br />

Arnás catchment calculated for the rill catchments are between 0.014 and 0.37 l s -1 and<br />

between 0.012 and 0.34 l s -1 . The used inflow was 0.15 l s -1 and this value caused the same<br />

sediment concentration range as measured at the Arnás catchment outlet. The rainfall<br />

simulations in the rill catchments show, that a rainfall intensity 40 mm h -1 could cause an<br />

inflow in the rills of 1.3 for rill 1 or even 4.15 l s -1 for rill 2 (table 2).<br />

89


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Table 6: Comparison of the runoff values of the Arnás catchment and the rill catchments; c.a.<br />

= catchment area<br />

Arnás c.a. 284 ha RE1 c.a. 450 m² RE2 c.a. 400 m²<br />

specific average runoff [l s -1 /m²] 0.0001 0.00007 0.00008<br />

specific runoff peak [l s -1 /m²] 0.0004 0.0001 0.0002<br />

It is to notice, that the connectivity between rill and river is not given in each case. The<br />

sediment and the water from the rill may not reach the river. The river transports the material<br />

from those areas that have a connectivity to the river. The trails improve the connectivity, but<br />

in some cases, it is not sufficient.<br />

Accordingly the rills can be regarded as very effective sediment providers for the receiving<br />

waters in a catchment because they can cause heavy sediment losses with relatively low<br />

runoff amounts: the calculated specific runoff values of the rills are clearly lower than the<br />

specific runoff values of the Arnás catchment. But this lower specific runoffs reach the same<br />

sediment concentrations as the Arnás catchment (table 6).<br />

The results clearly indicate that the rill erosion in the Arnás catchment, despite the relatively<br />

low measured values, is characterized by a high efficiency.<br />

The given standardized experimental setup makes it possible to compare different rills and to<br />

take statements about their efficiency for runoff and sediment budget of a catchment. Only<br />

with comparable conditions concerning water quantity and inflow intensity, it becomes<br />

possible to achieve appropriate statements on the basis of the measured results and to compare<br />

single experiments to each other at least qualitatively or semi-quantitatively.<br />

Measuring the slope is very important, but it has to be considered, that only average values<br />

over 1 meter slope length are measured. The position and the elevation of the knickpoints in<br />

the rill must be recorded exactly and accounted in the analysis.<br />

Using colour tracers, it became possible to record changes in flow velocity. This can be<br />

caused for example by the overflow of a plunge pool, a process, that delivers a lot of water<br />

and sediment in a short time. Without the knowledge of a change in flow velocity, the notice<br />

of an increase in sediment concentration could cause a wrong conclusion.<br />

In these experiments, 72 l of water are discharged with an intensity of 9 l min -1 into a rill, an<br />

intensity also used in laboratory experiments by POLYAKOV & NEARING (2003). They<br />

measured clearly higher sediment concentrations. That means, under other conditions, this<br />

water quantity is sufficient to cause higher sediment loss. The Arnás catchment is still as<br />

90


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

much degraded, that the most erodible material is removed and the rill's bottoms still reached<br />

the Flysch-plates . But it is a fact, that even in the slump material rills are developed. But for<br />

developing or increasing much higher water quantities and runoff intensities are needed.<br />

Considering the influence of initial soil moisture, one experiment consists of two runs. The<br />

influence of initial soil moisture was accounted by GOVERS et al. (1990) and GOVERS<br />

(1991). In their experiments under dry conditions the sediment concentrations almost reached<br />

the runoff's transport capacity, under higher initial soil moistures, the sediment concentrations<br />

were lower. In the rill experiments in the Arnás-catchment, the transport capacity was not<br />

reached because of the lack of transportable material and the low water amount. The influence<br />

of the initial soil moisture was clearly to notice, in the second run under saturated conditions<br />

the flow velocities and the runoff values were higher than in the first run with a soil moisture<br />

of about field capacity caused by snow melt . The starting moistures of the rainfall<br />

simulations were between 22 and 24 grav. %. But some experiments also show higher<br />

sediment concentrations under the higher initial soil moisture conditions in the second run<br />

(almost saturated). Under wet conditions, the flow velocities are higher and the cohesion<br />

forces are weaker, but there is less loose material in the rill. Under dry conditions, (in this<br />

case: less wet) the sediment concentrations are higher if there is a high quantity of lose<br />

material in the rill, if all this material is removed in the first run and if the higher flow<br />

velocities in the second run cannot cause an incision into the rill’s bed.<br />

5 Conclusion<br />

Three rill experiments according the described method were executed in the Arnás-catchment<br />

in the Pyrenees.<br />

In these experiments, as well naturally developed rill, anthropogenically induced rills like rills<br />

developed from motorcycle trails as real lanes were tested. The used 72 liters of water, that<br />

were disposed with an intensity of 9 l min -1 into the rill, caused material loss in the rill, but the<br />

dimensions were really different. There were samples without detectable sediment, the highest<br />

sediment concentrations reached values of 6.31 g l -1 . These results indicate that in the Arnás<br />

catchment, the rills have developed and are still active today under higher runoff intensities<br />

than the ones used in the experiments.<br />

From the induced 72 l of water, between about 8 and 49 l reached the end of the test length,<br />

what means a rate between 11 and 68 %. Is the passed flow length accounted, r-l-factors of<br />

between 2.1 and 17.1 were calculated. This factor is necessary, because in the third<br />

91


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

experiment, the flow length was reduced from 25 m to 15 m for accounting the given<br />

conditions. It is clearly to detect, that the different rills have a different efficiency for runoff<br />

and sediment budget.<br />

By recording the flow velocities, slope, sediment concentrations and runoff values, the<br />

behaviour of the three tested rills in the Arnás catchment could be comprehended. But the<br />

results can only be applied to this three tested rills, it is not possible to give a statement about<br />

the behaviour of rills in a catchment. Despite similar experiment conditions, the factors slope,<br />

different development state of the rill (depth and width) and position of the rill in the<br />

landscape determine the behaviour of the rills. The furthermost developed rill shows the<br />

lowest r-l-factor, the water can’t reach the measuring point after 15 m in the first run. So, it<br />

must be assumed, that this rill has formed under much higher rainfall-runoff situations as<br />

simulated in the experiments. It is necessary to go on testing rills with a standardised method<br />

like the rill experiments to collect more information about a rill's behaviour concerning runoff<br />

and sediment transport, and maybe to give universal statements. This rill experiments offer<br />

qualitative or semi-quantitative analysis of single rills.<br />

References<br />

- BRUNO, C., DI STEFANO, C. & FERRO, V. (2008): Field investigation on rilling in<br />

the experimental Sparacia area, South Italy. - Earth Surface Processes and Landforms<br />

33: 263-279.<br />

- BUTZEN, V., SEEGER, M. & CASPER, M. (2009 submitted): Spatial pattern and<br />

temporal variability of runoff processes in Mediterranean Mountain environments -<br />

coupling experimental measurement and GIS-analysis. - Zeitschrift für<br />

Geomorphologie; submitted.<br />

- CALVO -CASES, A., GISBERT, J.M., PALAU, E. & ROMERO, M. (1988): Un<br />

simulador de lluvia portátil de fácil construcción. - In: Sala, M. & Gallart, F. (eds.):<br />

Métodos y técnicas para la medición en el campo de processo geomorfológicos: 6-15.<br />

- CASALI, J., LOIZU, J., CAMPO, M.A., DE SANTISTEBAN, L.M. & ÁLVAREZ-<br />

MOZOZ, J. (2006): Accuracy of methods for field assessment of rill and ephemeral<br />

gully erosion. - Catena 67: 128-138.<br />

- CERDÁ, A. & GARCÍA-FAYOS, P. (1994-95): Relaciones entre las pérdidas de agua<br />

y semillas en zonas acarcavadas. Influencia de la pendiente. - Cuadernos I Geográfica<br />

20-21: 47-63.<br />

92


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

- CERDAN, O., LE BISSONNAIS, Y., COUTURIER, A., BOURENNANE, H. &<br />

SOUCHÈRE, V. (2002): Rill erosion on cultivated hillsplope during two extreme<br />

rainfall events in Normandy, France. - Soil & Tillage Research 67: 99-108.<br />

- FLORES-CERVANTES, J.H., ISTANBULLUOGLU, E. & BRAS, R.L. (2006):<br />

Development of gullies on the landscape: A model of headcut retreat resulting from<br />

plunge pool erosion. - J. Geophys. Res. 111: F01010, doi:10.1029/2004JF000226.<br />

- GOVERS, G., EVERAERT, W., POESEN, J., DE PLOEY, J. & LAUTRIDOU, J.P.<br />

(1990): A Long flume study of dynamic factors affecting the resistance of loamy soil<br />

to concentrated flow erosion. - Earth surface processes and landforms 15: 313-328.<br />

- GOVERS, G. (1991): Time-dependency of runoff velocity and erosion: the effect of<br />

the initial soil moisture profile. - Earth Surface Processes and Landforms 16: 713-729.<br />

- HANCOCK, G.R., CRAWTER, D., FITYUS, S.G., CHANDLER, J. & WELLS, T.<br />

(2008): The measurement and modelling of rill erosion at angle of repose slopes in<br />

mine spoil. - Earth Surface Processes and Landforms 33: 1006-1020.<br />

- KNAPEN, A., POESEN, J., GOVERS, G., GYSSELS, G. & NACHTERGAELE, J.<br />

(2007): Resistance of soils to concentrated flow erosion: A review. - Earth-Science<br />

Reviews 80: 75-109.<br />

- LASANTA, T., PÉREZ RONTOMÉ, M.C. & GARCÍA RUIZ, J.M. (1994): Efectos<br />

hidromorfológicos de differentes alternativas de retirada de tierras en ambientes<br />

semiáridos de la depressión del Ebro. - In: García Ruiz, J.M. & Lasanta, T. (eds.):<br />

Efectos geomorfológicos del abandono de tierras: 69-82; Sociedad Española de<br />

Geomorfología.<br />

- LOCH, R.J. & THOMAS, E.C. (1987): Resistance to rill erosion: observations on the<br />

efficiency of rill erosion on a tilled clay soil under simulated rain and run-on water. -<br />

In: Bryan, R.B. (ed.): Rill erosion – processes and significance: 71-83; Catena Verlag.<br />

- LOCH, R.J. (2000): Using rainfall simulation to guide planning and management of<br />

rehabilitated areas: part I Experimental methods and results from a study at the<br />

Northparkes mine, Australia. - Land degradation and development 11: 221-240.<br />

- MERZ, W. (1993): Experimentelle Untersuchungen zur Rillenerosion auf<br />

landwirtschaftlich genutzten Böden in Kanada und der Volksrepublik China. -<br />

Freiburger Geographische Hefte, Heft 40, Selbstverlag des Institutes für Physische<br />

<strong>Geographie</strong> der Albert-Ludwig-Universität Freiburg i.Br., Freiburg, 147 pp.<br />

93


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

- MEYER, L.D., FORSTER, G.R. & RÖMKENS, M.J.M. (1975): Source of soil eroded<br />

by water from upland slopes. - Proc. 1972 Sediment Yield Workshop, US Dept. of<br />

Agric. Sediment. Lab., Oxford, Mississippi: 177-189.<br />

- MORGAN, R.P.C., MARTIN, L. & NOBLE, C.A. (1987): Soil erosion in the United<br />

Kingdom: a Case Study from Mid-Bedfordshire, England. - Occasional Paper No.14.<br />

- OOSTWOUD WIJDENES, D.J., POESEN, J., VANDEKERCKHOVE, L. &<br />

GHESQUIERE, M. (2000): Spatial distribution of gully head activity and sediment<br />

supply along an ephemeral channel in a Mediterranean environment. - Catena 39: S.<br />

147-167.<br />

- PFAHLS, C., SEEGER, M. & SAUER, T. (1999): Infiltrationsmessungen,<br />

Niederschlagssimulationen, TDR- und Einstichtensiometermessungen sowie<br />

konventionelle gravimetrische Bodenwassererhebungen – ein Methodenvergleich. -<br />

APT-Berichte 4, Nr. 10: 51-77.<br />

- POESEN, J. (1987): Transport of rock fragments by rill flow – a field study. - In:<br />

Bryan R.B. (ed.): Rill erosion – processes and significance: 35-54; Catena Verlag.<br />

- POLYAKOV , V.O. & NEARING, M.A. (2003): Sediment transport in rill flow under<br />

deposition and detachment conditions. - Catena 51: 33-43.<br />

- RIES, J.B., LANGER, M. & REHBERG, C. (2000a): Experimental investigations on<br />

water and wind erosion on abandoned fields and arable land in the central Ebro Basin,<br />

Aragón/Spain. - Zeitschrift für Geomorphologie N.F. Suppl.-Bd. 121: 91-108.<br />

- RIES, J.B., MARZOLFF, I. & SEEGER, M. (2000b): Der Einfluss von Beweidung<br />

auf die Vegetationsbedeckung und Bodenerosion in der Flyschzone der spanischen<br />

Pyrenäen. - In: G. Zollinger (ed): Aktuelle Beiträge zur angewandten physischen<br />

<strong>Geographie</strong> der Tropen, Subtropen und der Regio Trirhena: 167-194; Freiburger<br />

Geographische Hefte.<br />

- RIES, J.B., MARZOLFF, I. & SEEGER, M. (2003): Einfluss der Beweidung auf<br />

Vegetationsbedeckung und Geomorphodynamik zwischen Ebrobecken und Pyrenäen.<br />

- Geographische Rundschau 55(5): 52-59.<br />

- SEEGER, M., ERREA-ABAD, M.-P. & LANA-RENAULT, N. (2005): Spatial<br />

Distribution of Soils and their Properties as Indicators of Degradation/Regradation<br />

Processes in a Highly Disturbed Mediterranean Mountain Catchment. - Journal of<br />

Mediterranean Ecology Vol. 6 No.1: 53-59.<br />

- SEEGER, M., ERREA, M.-P., LANA-RENAULTE, N., BEGUERÍA, S., ARNÁEZ,<br />

J., MARTÍ, C., REGÜÉS, D. & GARCÍA-RUIZ, J.-M. (2005): Charakteristika des<br />

94


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Sedimenttransportes in einem Kleineinzugsgebiet der Aragonesischen Pyrenäen. -<br />

Zeitschrift für Geomorphologie N.F. Suppl.-Bd. 138: 211-228.<br />

- SIDORCHUK, A. (2005): Stochastic components in the gully erosion modeling. -<br />

Catena 63: 299–317.<br />

- VANDEKERCKHOVE, L., POESEN, J., OOSTWOUD-WIJDENES, D. & DE<br />

FIGUEIREDO, T. (1998): Topographical thresholds for ephemeral gully initiation in<br />

intensively cultivated areas of the Mediterranean. - Catena 33: 271-292.<br />

- WOODWARD, D.E. (1999): Method to predict cropland ephemeral gully erosion. -<br />

Catena 37: 393-399.<br />

95


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Kapitel 4<br />

<strong>Wirtz</strong> et al. (2012a): Field experiments for understanding and quantification of rill<br />

erosion processes. Catena 91, 21-34.<br />

96


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Field experiments for understanding and quantification of rill erosion processes<br />

<strong>Wirtz</strong> S. a , Seeger M. b , Ries J.B. a<br />

a: Dep. of Physical Geography, Trier University<br />

b: Dep. of Land Degradation and Development, Wageningen University<br />

Abstract<br />

Despite many efforts over the last decades to understand rill erosion processes, they remain<br />

unclear. This paper presents the results of rill experiments accomplished in Andalusia in<br />

September 2008 using a novel experimental set up. 72 L of water are introduced with an<br />

intensity of 9 L min -1 into a rill. Rill cross sections, slope values, flow velocities and sediment<br />

concentrations were measured and these values were used to calculate sediment detachment<br />

and transport. Each experiment was repeated once within 15 min. With this new experimental<br />

setup it is possible to calculate several hydraulic parameters like hydraulic radius, wetted<br />

perimeter, flow cross section, transport rate and transport capacity which are usually<br />

estimated from coarse flow and rill parameters. In rill experiments, four different natural rills<br />

were flooded with the same experimental setup. Several processes like transport of loose<br />

material, erosion, bank failure and knickpoint retreat and the runoff effectiveness showed<br />

different and variable intensities. The sediment concentrations ranged between 5.2 – 438 g L -<br />

1 . In most cases, detachment rates are close to the transport capacity and, in some cases, the<br />

transport capacity is even exceeded. This can be explained by the occurrence of different<br />

erosion processes within a rill (e.g. detachment, bank failure, headcut retreat) which are not<br />

all explained by the given equations. The results suggest that the existing soil erosion<br />

equations based on shear forces exerted by the flowing water are not able to describe rill<br />

erosion processes satisfactory. Too many different processes with a high spatial and temporal<br />

variability are responsible for rill development.<br />

Key words: soil erosion, rill erosion, experimental method, quantification of process activity<br />

Introduction<br />

Soil erosion in general, and the development of rills in special, is the result of a very complex<br />

interaction of soil properties with a high spatial and temporal variability (Nachtergaele et al.,<br />

97


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

2001, 2002; Poesen et al., 1999) in which the morphology of a rill and the rill's headcut<br />

morphology (Flores-Cervantes et al., 2006) may be determinant as well as stochastically<br />

driven processes (Sidorchuk, 2005). This leads to great difficulties in quantifying soil erosion<br />

processes and makes soil erosion measurements hardly comparable (Knapen et al., 2007;<br />

Auerswald et al., 2009; Stroosnijder, 2005).<br />

The two main processes in soil erosion are inter-rill and rill erosion by flowing water,<br />

however the mechanisms of these two processes are completely different. The detachment in<br />

inter-rill erosion is caused and enhanced by drop-impact (Beuselinck et al., 2002) and, in<br />

addition to the soil’s intrinsic characteristics (Kuhn and Bryan, 2004; Kuhn et al., 2003; Le<br />

Bissonnais et al., 2005), is thought to depend mainly on rainfall intensity (Brodie and<br />

Rosewell, 2007; Bryan, 2000). Rill erosion is, in contrast, caused by the concentrated flow of<br />

water (Bryan, 2000; Govers et al., 2007; Knapen et al., 2007) and is considered to be the most<br />

important process of sediment production (and thus, soil loss) (Cerdan et al., 2002; Poesen,<br />

1987). The resulting rills may be persistent and develop into gullies, hindering further land<br />

use (Woodward, 1999; Vandekerckhove et al., 1998). Especially on fallowland and shrubland,<br />

rills can develop without disturbance by land management measures like ploughing. In the<br />

Mediterranean, huge areas of fallowland and shrubland exist (Ries, 2003) thus rills can<br />

develop very fast and cause high soil losses.<br />

Generally, rill erosion is understood as the effect of flowing water exceeding a certain<br />

threshold of soil resistance (Knapen et al., 2007). During the last decades, several approaches<br />

to describe and predict soil detachment and sediment transport in rills have been developed,<br />

and great effort has been made to evaluate their suitability for that purpose (Giménez and<br />

Govers, 2002; Govers et al., 2007; Hessel and Jetten, 2007). Unfortunately, the different<br />

approaches to describe this phenomenon have turned out to be at least weak, if not<br />

contradictory (Giménez and Govers, 2002; Govers et al., 2007; Merz and Bryan, 1993). This<br />

is attributed mainly to methodological differences in all monitoring and experimental set-ups<br />

to achieve the rills (Knapen et al., 2007; Merz and Bryan, 1993). It also appears that particle<br />

detachment and sediment transport may be controlled by different characteristics of the<br />

flowing water and, therefore, a comprehensive description may not be possible (Govers et al.,<br />

2007). However, soil erosion measurements are still lacking (Stroosnijder, 2005) and there is<br />

a recognized need to perform field experiments to ascertain the role of rills in soil erosion<br />

(Govers et al., 2007). As the observation of erosion in the field is subordinated to the<br />

stochastic character of the erosion events (Auerswald et al., 2009) and to a high dependency<br />

of the measurement technique (Casali et al., 2006), standardized and reproducible field<br />

98


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

experiments are needed, in which the relevant parameters of runoff and sediment transport<br />

can be measured and which, at the same time, can produce data to characterise the rill's<br />

behaviour in its environment.<br />

Most experimental work about rill erosion that has been carried out, both in the laboratory<br />

(Brunton and Bryan, 2000; Bryan and Poesen, 1989; Gilley et al., 1990; Huang et al., 1996;<br />

Mancilla et al., 2005) as well as under field conditions (De Santisteban et al., 2005; Helming<br />

et al., 1999; Rejman and Brodowski, 2005) used soils with different textures and natural or<br />

simulated rainfall. The aim of the research groups was to observe rill network formation<br />

(Bruno et al., 2008; Mancilla et al., 2005), define the initial conditions for rilling (Bruno et al.,<br />

2008, Bryan et al., 1998; Govers and Poesen, 1988; Slattery and Bryan, 1992; Torri et al.,<br />

1987), study the development of rill head morphology (Bruno et al., 2008; Brunton and<br />

Bryan, 2000), estimate the main hydraulic variables like cross-section area, wetted perimeter,<br />

hydraulic radius, mean velocity and shear stress (Bruno et al., 2008; Foster et al., 1984; Gilley<br />

et al., 1990; Giménez et al., 2004; Govers, 1992b) or propose mathematical models for<br />

estimating soil loss due to rill erosion (Favis Mortlock, 1998; Favis Mortlock et al., 2000;<br />

Bruno et al., 2008; Foster, 1982; Nearing et al., 1989).<br />

In most laboratory experiments the effort is made to find relationships between different<br />

factors. The influence of runoff on soil detachment is an often investigated question. Other<br />

parameters often tested are slope length, percolation, rill development, critical Froude<br />

number, critical shear stress different soil characteristics, slope, rainfall intensity, flow<br />

velocity or flow velocity distribution, bed morphology and flow area.<br />

In such a way, Bryan and Poesen (1989) tested, in laboratory experiments, the relationship<br />

between slope length, percolation, runoff and rill development. The flume used had a<br />

maximum length of 24.5 m, consisting of ten segments of 2.45 m. At the end of the flume,<br />

runoff was measured. They showed that runoff is not a simple function of rainfall excess and<br />

slope length but a more complex process dominated by surface sealing, rill development and<br />

headcut incision. Rill initiation is controlled by established threshold hydraulic conditions, the<br />

further development of the rills and headcuts is complex and depending on different<br />

thresholds. Torri et al. (1987) related in laboratory experiments, with variable slope, runoff<br />

and rainfall intensity, the critical Froude number and the critical shear stress to some soil<br />

characteristics. Critical shear stress was found to be correlated to soil shear strength. But this<br />

result could not be confirmed in all further studies. In an other laboratory study, Nearing et al.<br />

(1991) measured flow shear stresses ranged from 0.5 to 2 Pa, while tensile strengths ranged<br />

from 1 to 2 kPa, a difference in magnitude of 1000. Despite this conflict, detachment rates of<br />

99


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

nearly 300 g m -2 s -1 were measured. He explained this result with turbulent burst events which<br />

are much greater than the average flow shear stresses.<br />

Giménez et al. (2004) tested, in a laboratory flume experiment, the velocity distribution in<br />

rills and the relationship between flow velocity and rill bed morphology. They showed that<br />

bed roughness increases with slope while flow velocity decreases. Flow velocity increases<br />

until a threshold Froude number between 1.3 and 1.7 is reached, and a hydraulic jump occurs<br />

leading to the formation of a pool. Govers (1992b) tested, in another laboratory flume, the<br />

relationship between discharge, flow velocity and flow area. He showed that mean flow<br />

velocity is not at the front of the water, but where the water reached 80-90% of its maximum<br />

width. The mean flow velocity and cross-sectional flow area can be related to discharge for<br />

rills eroding loose, non layered materials like agricultural soils. Soil characteristics and slope<br />

appear to be of minor importance in this study.<br />

The main problem with these laboratory experiments is that the results cannot be easily<br />

transferred to natural rills. In the laboratory, flumes with compacted soil material are used, but<br />

Giménez & Govers (2002) showed that most of the data attained on rill models with smooth<br />

beds cannot be applied to naturally developed rills with rough beds. In many cases, hydraulic<br />

parameters are extracted from equations created to describe flow behaviour in rivers. Govers<br />

(1992a) and Govers et al. (2007) showed that these parameters cannot simply be transferred to<br />

flow behaviour in rills. This process oriented research needs also to be conducted in natural<br />

rills, using a mixture of process oriented (laboratory) experiments and field research.<br />

However field research is often in pursuit of other targets. Experimental work is very rare.<br />

Interests are adjusted to catchment areas, long-term-observations on plots or, in best cases, the<br />

measurement of different parameters under natural rainfall. Some examples are given here.<br />

De Santisteban et al. (2005) tested two different indices to characterise the influence of<br />

watershed topography on channel erosion. The first is defined as the product of watershed<br />

area and the partial area-weighted average slope. The other one is similar but uses the slope as<br />

the weighting factor, i.e. it is the product of watershed area and the length-weighted average<br />

slope. It was shown, that for a wide range of soil, climate, soil use and management<br />

conditions, the close relationship between soil erosion and topography can be quantified using<br />

the two indices. Govers & Poesen (1988) observed on a 7500 m² field plot the evolution of a<br />

rill and gully system. The periodic survey started on 15.11.1983 and finished at 3.10.1984.<br />

They measured detachment rates and used splash cups to get data about splash erosion. The<br />

sediment being detached by splash on inter-rill areas is transported to the channel system<br />

mainly by inter-rill wash. Rill and gully erosion is more important than inter-rill erosion, but<br />

100


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

the relative importance of inter-rill erosion varies in time and space, due to changes of the<br />

inter-rill surface characteristics and the activation of sidewall and gullying processes in the<br />

channel network. Bruno et al. (2008) accomplished field investigations under natural rainfall.<br />

They measured cross sections, runoff and soil loss and proposed a simple mathematical model<br />

for estimating soil loss. The analysis of the measured erosive events allowed establishment of<br />

the proportion of both rill erosion and inter-rill erosion on total soil loss. The measurements<br />

showed that rill erosion increases the total sediment transport efficiency because rill flow is<br />

able to transport both the inter-rill eroded sediments and sediment particles eventually<br />

detached from the rill wetted perimeter. The measurements also allowed verification of a<br />

relationship between rill length and rill volume, which was theoretically deduced by the<br />

dimensional analysis and self-similarity theory. This equation shows that rill length can be<br />

usefully employed as a severity index of the rilling process. The morphological evolution of<br />

the rill cross-section showed that in the first part of the rill length the channelized flow is able<br />

to erode the wetted perimeter and to transport the eroded particles, while in the terminal part<br />

of the rill the actual sediment load is high and the flow is only able to transport the particles<br />

coming from upstream without scouring the rill perimeter. The shear stress profile along the<br />

rill length confirmed that in the terminal part of the rill the flow is able to transport the<br />

upstream eroded sediment particles, but not able to detach additional material.<br />

The review of the experimental research on rill erosion processes shows the weak and<br />

partially contradictory results of attempts to understand rill erosion processes. This can be<br />

seen in the large number of studies dealing with the relationships between flow parameters<br />

and particle detachment and transport. This is even clearer regarding the rills' development<br />

and behaviour in the field. There is still a lack of direct observation of the throughflow and<br />

sediment transport/detachment characteristics in natural rills. On the other hand, rills and their<br />

behaviour in the landscape can give insight to the main dominating soil erosion processes and<br />

their magnitude and frequency. But for this we have to understand their functioning with<br />

determined amounts of throughflowing water. Therefore, a novel and simple method for<br />

characterizing the rills hydraulic effectiveness and the erosion dynamics within was<br />

developed.<br />

In the landscape context the following questions have to be addressed:<br />

1. What is the rill's contribution to soil erosion?<br />

2. How effective are rills for routing water and sediments through their environment?<br />

The rill flow and erosion experiments are thought to provide insights to contribute finding<br />

answers. This experiment should be able to reflect, at different levels of detail, the rill's<br />

101


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

hydraulic and erosion dynamics within the landscape in which they developed. Experiments<br />

have been designed so that they can be easily adapted to different rill sizes and morphologies<br />

as well as to variable runoff, thus providing information about the different processes<br />

involved in rill erosion. Within, we attempt to approach standardised description methods for<br />

the rill erosion processes.<br />

The aim of the novel experimental field setup presented here is to provide a reproducible tool,<br />

in which the magnitude of rill erosion processes can be understood, and at the same time<br />

contribute to understanding the effect of rills in their environment. The new method can be<br />

used to test existing rills in field experiments and generate information about their hydraulic<br />

effectiveness and erosion dynamics. This experimental setup only allows to identify and to<br />

compare different processes and their activity within a specific rill. Similar to the uncertain<br />

erosion behaviour of replicated test plots (Nearing, 1998; Nearing et al., 1999) a direct<br />

comparison of results from different experiments is not valid.<br />

Material and Methods<br />

Study areas<br />

The four study areas in Andalusia are located at Negratin, Freila, Salada and Belerda as<br />

presented in Figure 1. UTM coordinates of the tested rills are given in Table 1.<br />

Fig. 1: Location of the test areas in South East Spain<br />

102


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Negratin and Freila: The areas are located within the Hoya de Baza sedimentary basin and<br />

composed of marls, in which calcareous Regosols have developed. The climate is semi-arid<br />

and vegetation is dominated by low shrubs and Stipa tenacissima grass tussocks. The land<br />

cover at the south side of the Negratin-dam is dominated by abandoned cereal fields, which<br />

are extensively grazed by sheep and agricultural land use is comprised mainly of cereal dryfarming<br />

and almond grooves (Seeger, 2007).<br />

Salada: Located at the SE-margin of the Betic range (SE-Spain), inside the penibetic complex.<br />

The area is composed of conglomerates with a clayey to loamy matrix in which Regosols as<br />

well as to fairly developed (Calcic) Cambisols have developed. Vegetation is similar to that<br />

found in the Freila and Negratin-area. The climate is semi-arid too, but less accentuated than<br />

in the previously mentioned area (Seeger, 2007). Here we can find a mixed pattern of rainfed<br />

agricultural areas, mainly cereals, olives and almonds, and abandoned or uncultivated areas.<br />

Belerda: This test area is located in the Guadix basin. The parent material consists of tertiary<br />

and quaternary conglomerates, sands, silts and clays. The soil texture class following the FAO<br />

is a silty clay loam. The land use is separated into cultivated areas, with almond and olive<br />

groves, and abandoned agricultural fields (Vandekerckhove et al., 2003). The climate is,<br />

though still semi-arid, characterised by higher average annual temperatures and precipitations<br />

in comparison with the other test areas.<br />

The climatic parameters of the test fields are summarized in Table 1.<br />

Table 1: Temperature, precipitation, northing and easting of the four test areas.<br />

Tested Rill meteorological station average annual temperature annual precipitation easting rill northing rill<br />

Negratin Baza 14.2°C 368 mm 505835 4156275<br />

Freila Baza 14.2°C 368 mm 504874 4152414<br />

Salada Embalse Valdeinfierno 13.4°C 311 mm 595759 4187298<br />

Belerda Granada 15.6°C 473 mm 477694 4132866<br />

Tested rills<br />

An overview of the main descriptors of the rills is given in Table 2. Particle size analysis was<br />

conducted following AD-HOC-Arbeitsgruppe Boden (2005). Soil texture is classified<br />

following FAO guidelines (FAO, 2006).<br />

103


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Table 2: Rill parameters; NCA = needed catchment are to reach a runoff of 9 L min -1 ; gravel<br />

content refers to total soil, the rest only to fine earth fraction. Particle-size class limits are<br />

determined following AD-HOC-Arbeitsgruppe Boden 2005.<br />

Negratin Freila Salada Belerda<br />

Ø Slope [°] 20.6 9.7 7.7 17.1<br />

max. Slope [°] 31 19 11 28<br />

tested flow length [m] 20 21 15 16<br />

catchment area [m²] 1200 1100 170 5400<br />

texture class L SiL SiCL L<br />

Gravel [%] > 2000 μm 1 13 1 13<br />

coarse sand [%] 2000-630 μm 10 1 2 10<br />

middle sand [%] 630-200 μm 10 6 2 10<br />

fine sand [%] 200-125 μm 8 2 1 8<br />

very fine sand [%] 125-63 μm 17 9 7 17<br />

coarse silt [%] 63-20 μm 13 28 17 13<br />

middle silt [%] 20-6.3 μm 13 22 17 13<br />

fine silt [%] 6.3-2 μm 14 17 24 14<br />

Clay [%] < 2 μm 15 15 29 15<br />

organic matter [%] 11.5 11.5 8.7 12.15<br />

starting soil moisture [% w/w] 5.08 2.4 4.48 5.87<br />

Kt [s² m 0,5 kg -0,5 ] 0.0095 0.0095 0.0095 0.0091<br />

location WEPP dataset Frederick Frederick Frederick Opequion<br />

maximum width [m] 0.3 0.3 0.3 0.3<br />

maximum depth [m] 0.1 0.2 0.2 0.2<br />

vegetation cover [%] ~ 5 ~ 0 ~ 0 ~ 10<br />

rock fragment cover [%] ~ 5 ~ 5 ~ 5 ~ 10<br />

volume [m³] 0.2 0.26 0.23 0.19<br />

MP 1<br />

cross section [m²] 0.0107 0.0217 0.0221 0.0140<br />

slope [°] 30 3 6 14<br />

MP 2<br />

cross section [m²] 0.0102 0.0160 0.0244 0.0200<br />

slope [°] 19 19 11 18<br />

MP 3<br />

cross section [m²] 0.0160 0.0100 0.0144 0.0213<br />

slope [°] 15 10 7 18<br />

grain density [kg m -3 ] 2650 2710 2660 2610<br />

dry bulk density [kg m -3 ] 1410 1510 1320 1450<br />

runoff rainfall simulations [%] 71.5 90 85 51<br />

NCA [m²] 19 15 16 27<br />

The rill in Negratin (Fig. 2, upper left) is situated on the steepest slope (average 20.6°) of the<br />

tested rills with a catchment area of the rill’s headcut of about 1200 m². The total rill length<br />

tested was 20 m in length with a maximum width of about 0.3 m and a maximum depth of 0.1<br />

m. Vegetation and rock fragments cover only approximately 5% of the rill. The areas of the<br />

three measured cross sections are 0.0107 m² at 6 m, 0.0102 m² at 10 m and 0.016 m² at 16 m<br />

of the tested rill's length. With these values, the total volume was estimated to be about 0.2<br />

m³. The rill was carved into a loamy soil with a moderate bulk density of 1410 kg m -3 (BD3,<br />

104


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

(FAO, 2006)), in which the particle density was measured to be 2.65 g cm -3 . Starting soil<br />

moisture was 5.08 % w/w.<br />

The rill at Freila (Fig 2, upper right) developed on a leveled field. The maximum depth of the<br />

rill is about 0.2 m, the maximum width 0.25 m. Vegetation cover is not detectable, rock<br />

fragment cover is about 5%. Using the cross sections of 0.0217 m² (5 m), 0.016 m² (12.5 m),<br />

0.01 m² (15.5 m) and the flow length of 21 m, a volume of about 0.26 m³ was estimated for<br />

the whole rill. The soil texture is classified as medium clay silt, with a particle density of 2.71<br />

g cm -3 and a moderate bulk density of 1510 kg m -3 (BD3). Starting soil moisture was 2.4 %<br />

w/w.<br />

The rill at the Salada site (Fig 2 bottom left) formed on an unpaved road with an average<br />

slope of 7.7°. The rill's headcut catchment covers an area of about 170 m². The maximum<br />

width within the 15 m length tested is about 0.3 m, the maximum depth round 0.2 m. Only 5<br />

% of the rill is covered by rock fragments, and there is no vegetation cover in the rill. The<br />

cross section at the first measuring point at 4 m has an area of 0.0221 m², after 7 m, a cross<br />

section of 0.0244 m² was measured and after 12 m, the cross section was 0.0144 m². For the<br />

tested length of 15 m, there is a volume of 0.23 m³ calculated. The soil texture is classified as<br />

silty clay loam, with a low bulk density (BD2) of 1320 kg m -3 . The particle density value<br />

reaches 2.66 g cm -3 . Starting soil moisture was 4.48 % w/w.<br />

The Belerda rill (Fig 2 bottom right) developed in a loam soil with a high content of sand (45<br />

% of the fine soil) and coarse fragments (13 % of the total soil). It is located on a steep<br />

agricultural road with an average slope of 17.1° and a catchment area of about 5400 m². We<br />

tested 16 m of rill length, with a maximum width of 0.3 m and a maximum depth of 0.2 m.<br />

Vegetation and rock fragments cover about 10%. The cross sections were measured at 4.5 m,<br />

at 8 m and at 11 m, they reach areas of 0.014 m², 0.020 m² and 0.021 m² respectively leading<br />

to an estimated rill volume of about 0,19 m³ . The bulk density is with 1450 kg m -3 moderate<br />

(BD3), the grain density is only 2.61 g cm -3 . Starting soil moisture was 5.87 % w/w.<br />

105


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Fig. 2: The 4 tested rills: top left Negratin; top right Freila; bottom left Salada; bottom right<br />

Belerda<br />

Rill experiment<br />

The rill experiments consist of two runs: first the rill is tested under field conditions (“dry”);<br />

in a second run, about 15 min later, the same rill is tested under “wet” conditions.<br />

With a motor driven pump, a constant discharge of 9 L min -1 is maintained for 8 minutes,<br />

resulting in a total water inflow of 72 L. To avoid mobilization of soil particles by the<br />

inflowing water at the starting point of the experiment, the inflow velocity and pressure are<br />

reduced with an inflow device, which allows the water to enter the rill as a light stream (<strong>Wirtz</strong><br />

et al., 2010).<br />

The flow velocity within the rill is characterized in three steps: the travel time of the<br />

waterfront and of two different colour tracers is measured for every meter using a<br />

106


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

chronograph. The first colour tracer is introduced at the starting point after three of the eight<br />

minutes of experiment time, the second follows at six minutes. By means of this procedure,<br />

three velocity curves are recorded and changes in flow dynamics can be detected. For safety<br />

reasons, food colourings (E 124 (red) and E 13 (blue)) are used as colour tracers.<br />

The runoff height is continuously measured in a small flume at the end of the rill by a<br />

pressure transducer (Ecotech DL/n, V2.35). For runoff curve calibrating, runoff at outflow is<br />

measured volumetrically at regular intervals. With this, a constant measurement of the<br />

discharge is possible in high temporal resolution and throughout the whole experiment.<br />

The rill's morphology is characterized by measuring its slope with a spring bow of 1m range<br />

and a digital air lever. It should be noted that slope measuring provides only average slopes<br />

per 1 meter. A step or a knick-point in the rill is not accounted for, however its position and<br />

height are recorded.<br />

For gathering intermediate data on suspended sediment transport, three suitable measuring<br />

points (MP) are selected. Small knickpoints in the rill have proven to be useful sampling<br />

points because there is no need to press the bottle into the rill bottom. Four water samples are<br />

taken: the first as soon as the waterfront has reached the sampling point, the second after 30<br />

seconds, the third after 1:30 and the fourth 2:30 after the arrival of the water. The sediment<br />

concentration is determined by filtration of the samples in laboratory (<strong>Wirtz</strong> et al., 2010).<br />

The rill cross section was measured at each measuring point. With thin metal sticks, the<br />

distance between ground level and rill bottom was measured in 0.02 m steps. This allows an<br />

accurate calculation of the rills transversal area and an estimation of the rills volume.<br />

Descriptors of rill experiment<br />

Several simple descriptors are used for characterizing the hydraulic function of the rills:<br />

The total runoff volume at the outlet V R [L] is calculated. When divided by the inflow<br />

volume, V I [L] , the runoff coefficient (RC) [without unit] is obtained by<br />

V<br />

V<br />

R<br />

RC =<br />

(eq. 1)<br />

I<br />

To compare runoff values measured in rill experiments with different experimental lengths,<br />

we calculate the runoff length factor (RC L ) [m] by multiplying the runoff coefficient by the<br />

tested length [m]. It is an expression of runoff effectiveness and an inverse measure for<br />

infiltration capacity within the rill:<br />

VR<br />

RC<br />

L<br />

= L<br />

(eq. 2)<br />

V<br />

I<br />

107


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

The maximum runoff intensity Q m [L s -1 ] is set into relation with the inflow intensity Q I [L s -<br />

1 ], leading to Q f [without unit], which is the intensity factor. Some additional factors of the<br />

flow duration are calculated as they are also indicators of the transformation of the input pulse<br />

such as the begin and end times of the outflow at the end of the rill and the flow duration (t S<br />

[s], t E [s] and T R [s] respectively). The changes in flow duration are indicated by the quotient<br />

(d T ) [without unit] between runoff duration and inflow duration T I [s].<br />

Velocity is recorded as an average value (v a ) [m s -1 ] and the maximum (v max ) [m s -1 ] value for<br />

each of the 3 measurements in each experiment. The flow velocity at which each of the<br />

samples was collected is estimated as follows: We assume linear changes between the<br />

measured flow velocities (front, tracer 1 and tracer 2). Knowing the exact time of arrival of<br />

the water front to the sampling point, we estimate the flow velocity for each of the samples<br />

collected interpolating between the measured velocities.<br />

Taking into account the rill’s morphology, the estimated velocity at the sampling location, the<br />

points in time and the sediment concentration, we calculated the maximum hydraulic shear<br />

stress τ [Pa] (eq. 3) according to Nearing et al. (1997).<br />

τ = ρgRS<br />

(eq. 3)<br />

where ρ represents the fluid density [kg m -3 ] which was calculated taking into account the<br />

sediment concentration in the samples, g the gravitational acceleration (9.81 m s -2 ), R the<br />

hydraulic radius [m], and S the effective slope (sin(slope angle)). The hydraulic radius is<br />

calculated as follow:<br />

A<br />

R (eq. 4)<br />

WP<br />

with A = flow cross section [m²] and WP = wetted perimeter [m]. Flow cross section is<br />

calculated using the runoff and the flow velocity:<br />

Q<br />

A (eq. 5)<br />

v<br />

with Q = runoff [m³ s -1 ] and v = flow velocity [m s -1 ]. The runoff, needed to calculate the<br />

hydraulic radius at the measuring points, was set equal to inflow (0.00015 m 3 s -1 ), so the<br />

calculated values can be understood as the maximum possible. Using the known flow cross<br />

section and the rill cross section, water level and wetted perimeter can be determined.<br />

The sediment transport within the rill was characterized by the sediment concentration in the<br />

samples. Additionally, we calculated the sediment transport per unit time and unit rill length<br />

(D L [kg m -1 s -1 ]) and per unit rill bed area (D A [kg m -2 s -1 ]).<br />

108


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

The transport capacity T C [kg s -1 ] at each of the sampling points was calculated according to<br />

WEPP (Foster et al., 1995):<br />

3<br />

2<br />

T<br />

C<br />

= k t f<br />

(eq. 6)<br />

where k t is the transport coefficient [m 0.5 s 2 kg -0.5 ] and τ f is the flow shear stress [Pa]. Here we<br />

used the maximum hydraulic shear stress calculated as mentioned above.<br />

Rainfall simulations<br />

For the characterisation of the rill catchments rainfall simulations were performed with a<br />

small mobile rainfall simulator based on the type of Calvo Cases et al. (1988), Lasanta et al.<br />

(1994) and Cerdá and García-Fayos (1994-95). The plot was bounded by a steel ring of 0.6 m<br />

in diameter (0.28 m²) and 0.006 m height, having an outlet of 0.1 m width with a narrowing<br />

runoff collector (Ries et al., 2000; Ries et al., 2009). The rainfall intensity was 40 mm h -1 , the<br />

fall height 2 m. The suitability of this method for the explanation of the influence of external<br />

factors on runoff generation and erosion with this methodology was demonstrated by Seeger<br />

(2007).<br />

The results gathered from the rainfall simulations were used to estimate the minimum area<br />

needed to generate the runoff which was used to perform the rill experiments (9 L min -1 ). This<br />

information provides insight regarding the possibility of runoff in the rill. This factor called<br />

“needed catchment area” (NCA) is calculated as follow:<br />

NCA<br />

I<br />

RO<br />

RS<br />

100<br />

RE<br />

* I<br />

RS<br />

(eq. 7)<br />

with I RE = inflow intensity of the rill experiment, in this study, the value is 9 L min -1 . RO RS is<br />

the runoff of the rainfall simulation [%] and I RS is the rainfall intensity of the rainfall<br />

simulation per area in L min -1 m -2 .<br />

Results<br />

Runoff dynamics<br />

In the Negratin-rill catchment the runoff coefficient of the rainfall simulations reached 71.5%.<br />

Only 19 m² of the catchment is needed to generate the experiments' runoff during a moderate<br />

heavy rain (Table 2).<br />

In the Freila test site, the runoff coefficient increased to 90 %, reducing the needed<br />

contributing area to only 15 m 2 . Salada showed similar values, (RC=85 %), for which only 16<br />

109


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

m 2 would be necessary to generate the experimental runoff in the rill. Due to the considerably<br />

lower RC in Belerda (51 %), the contributing area was estimated to be 27 m 2 (Table 2).<br />

Rill hydraulics<br />

The descriptors of the rills’ hydraulics are summarized in Table 3.<br />

After the given flow length, only 19 L water reach in the first run in Salada the end of the<br />

tested rill part. The highest value (59 L) is measured in the second run in the Belerda<br />

experiment.<br />

The highest maximum runoff intensity was measured in the first run in the Belerda<br />

experiment (0.15 L s -1 ), the lowest maximum runoff intensity in the first run of the Salada<br />

experiment (0.07 L s -1 ).<br />

The relationship between inflow intensity and runoff intensity shows the highest value in the<br />

first Belerda run, here the inflow and the runoff intensity are equal. The lowest value is<br />

measured in Salada, run a (0.47).<br />

The highest runoff coefficient is measured in Belerda run b, here 82 % of the inflow water<br />

reach the runoff measuring point. In the first Salada run, only 27 % reach the end of the tested<br />

rill part.<br />

The runoff-length-factor shows its lowest value in Salada run a (3.96 m) the highest value is<br />

measured in Freila run a (13.71 m). In all experiments, the second run shows higher values<br />

than the first run but in the second run in Freila, there are no data available because of<br />

technical problems.<br />

The runoff at the rill’s end starts in the second run in the Belerda experiment still after 70 sec.,<br />

in the first run of the Freila experiment, the water reaches the rill’s end after 175 sec. After<br />

593 sec., the runoff ends in the first run of the Salada experiment, in the first run of the Salada<br />

experiment, the runoff has finished after 866 sec. The runoff shows the longest duration in the<br />

first run in Negratin (700 sec.), the fastest runoff wave has been noticed in the first run in<br />

Freila (88 sec.).<br />

The relationship between runoff duration and inflow duration shows the highest value in<br />

Negratin run a (1.46), the lowest value in Salada run a (0.92).<br />

Average flow velocities show the highest values in Freila run b (Front), Negratin run a<br />

(Tracer 1) and Salada run b (Tracer 2), the lowest values in Freila and Salada run a (Front),<br />

and Belerda run b (Tracer 1 and Tracer 2). The maximum flow velocities show the highest<br />

values in Freila run b (Front) and Negratin run a (Tracer 1 and Tracer 2), the lowest values are<br />

measured in Freila run a (Front), Freila run b (Tracer 1) and Belerda run b (Tracer 2).<br />

110


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Table 3: Hydraulic parameters, ND = no data, F = waterfront, T1 = tracer 1, T2 = tracer 2<br />

Negratin Freila Salada Belerda<br />

run a b a b a b a b<br />

VR [L] 26 44 47 ND 19 55 43 59<br />

Qm [L s -1 ] 0.08 0.11 0.13 ND 0.07 0.13 0.15 0.14<br />

Qf 0.53 0.73 0.87 ND 0.47 0.87 1 0.93<br />

RC [-] 0.36 0.61 0.65 ND 0.27 0.76 0.6 0.82<br />

RCL [m] 7.22 12.22 13.71 ND 3.96 11.46 9.56 13.11<br />

tS [s] 166 108 175 ND 155 70 118 110<br />

tE [s] 866 666 663 ND 593 726 646 794<br />

TR [s] 700 558 88 ND 438 656 528 684<br />

va [m s -1 ]<br />

vmax [m s -1 ]<br />

dT 1.46 1.16 1.02 ND 0.92 1.37 1.15 1.43<br />

F 0.16 0.21 0.12 0.27 0.12 0.25 ND 0.15<br />

T1 0.36 0.3 0.29 0.28 0.29 0.31 0.26 0.25<br />

T2 0.32 0.31 0.35 0.34 0.38 0.39 0.29 0.24<br />

F 0.36 0.44 0.18 0.70 0.22 0.49 ND 0.29<br />

T1 1.05 0.78 0.42 0.37 0.44 0.46 0.49 0.5<br />

T2 0.93 0.85 0.65 0.56 0.6 0.61 0.44 0.38<br />

Erosion dynamics<br />

The erosion dynamic parameters are all summarized in table 4.<br />

Negratin: The average sediment concentration in the first run (run a) was 270 g L -1 with an<br />

average maximum flow velocity of 0.28 m s -1 . The maximum sediment concentration of 428<br />

g L -1 was measured after 30 sec at MP3. In the second run (run b), the average sediment<br />

concentration reached only 58 g L -1 , with an average maximum flow velocity of 0.27 m s -1 .<br />

The maximum sediment concentration increased to only 111 g L -1 and was measured at the<br />

waterfront at MP3 (figure 3).<br />

Freila: In the 1 st run, the average sediment concentration reached 224 g L -1 with an average<br />

maximum flow velocity of 0.25 m s -1 . In the second run 189 g L -1 and 0.3 m s -1 were<br />

reached. The maximum sediment concentration in the first run increased to 438 g L -1 ,<br />

measured at MP3 after 2:30 min. In run b, a maximum value of 267 g L -1 was measured at<br />

MP2 after 2:30 min (figure 4).<br />

Salada: The average sediment concentration in the 1 st run reached 183 g L -1 and increased to<br />

295 g L -1 in the 2 nd run. The average maximum flow velocities were 0.27 and 0.32 m s -1 . The<br />

111


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

maximum sediment concentration in both runs was measured after 2:30 min at MP2, reaching<br />

in run a 302 g L -1 , and 402 g L -1 in run b (figure 5).<br />

Belerda: In this experiment, the average sediment concentration in the first run reached 28 g L<br />

-1 at an average maximum flow velocity of 0.25 m s -1 . It should be noted that in this<br />

experiment the flow velocity of the waterfront could not be measured. The waterfront of the<br />

first run always shows the lowest flow velocity, so the 0.25 m s -1 mentioned here is too high.<br />

In the second run, the average sediment concentration decreased to 16 g L -1 . The average<br />

maximum flow velocity (waterfront included) reached 0.21 m s -1 . In the first run, the<br />

maximum sediment concentration reached was 54 g L -1 at the waterfront at MP3. In the<br />

second run, the waterfront at MP1 reached 29 g L -1 (figure 6).<br />

The average detachment rate shows the highest values in most cases at the second measuring<br />

point. In experiments Negratin b and Freila a, the highest values were reached at MP3, in<br />

experiment Belerda b, the highest value was reached at MP1. The maximum detachment rates<br />

also reach the highest values mostly at MP2, although Negratin b and Freila a reached them at<br />

MP3, and Negratin a at MP1.<br />

The highest average sediment flux values are reached at MP 3 for experiments Negratin a and<br />

b and Freila a; at MP 2 for experiment Freila b, Salada a and b and Belerda a. Belerda b<br />

reached the highest value at MP1. The maximum sediment flux values show the same trend,<br />

except that in experiment Belerda a the highest value is not reached at MP2 but at MP3.<br />

The transport capacity also shows a variable trend. In Negratin, the highest values were<br />

reached at the first measuring point, in Freila at the second measuring point. In Salada, the<br />

highest values were reached at the third measuring point. Only the maximum T C for the<br />

second run was reached at the second measuring point. In the second Belerda run, the highest<br />

values were reached at the third measuring point.<br />

112


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Table 4: Erosion dynamic parameters, ND = no data<br />

Negratin Freila Salada Belerda<br />

run a b a b a b a b<br />

1 0.0683 0.0113 0.0489 0.0350 0.0614 0.1060 ND 0.0086<br />

Ø Da [kg s -1 m -2 ] 2 0.0785 0.0149 0.0688 0.0784 0.0994 0.1545 ND 0.0054<br />

3 0.0652 0.0191 0.0844 0.0604 0.0559 0.0993 ND 0.0054<br />

1 0.1081 0.0197 0.0734 0.0444 0.0928 0.1279 ND 0.0117<br />

max. Da [kg s -1 m -2 ] 2 0.1063 0.0179 0.0963 0.0988 0.1543 0.1976 ND 0.0117<br />

3 0.0898 0.0239 0.1198 0.0686 0.0801 0.1256 ND 0.0063<br />

1 0.0366 0.0058 0.0259 0.0178 0.0238 0.0394 0.0040 0.0030<br />

Ø Qs [kg s -1 ] 2 0.0393 0.0072 0.0274 0.0337 0.0311 0.0503 0.0044 0.0019<br />

3 0.0454 0.0131 0.0474 0.0334 0.0273 0.0431 0.0040 0.0020<br />

1 0.0633 0.0106 0.0368 0.0226 0.0343 0.0483 0.0071 0.0044<br />

max. Qs [kg s -1 ] 2 0.0550 0.0087 0.0339 0.0400 0.0453 0.0602 0.0076 0.0041<br />

3 0.0641 0.0166 0.0656 0.0376 0.0349 0.0542 0.0081 0.0025<br />

1 0.3204 0.2265 0.0085 0.0045 0.0208 0.0150 ND 0.3197<br />

Ø Tc [kg s -1 ] 2 0.0719 0.0421 0.1365 0.1574 0.0444 0.0581 ND 0.3358<br />

3 0.1107 0.0689 0.0426 0.0299 0.0685 0.2412 ND 0.4060<br />

1 0.5549 0.3070 0.0127 0.0049 0.0274 0.0179 ND 0.3788<br />

max. Tc [kg s -1 ] 2 0.1070 0.0508 0.2795 0.1819 0.0526 0.0722 ND 0.2864<br />

3 0.1746 0.0882 0.0538 0.0333 0.1382 0.0143 ND 0.4970<br />

Fig. 3: Results RE Negratin<br />

113


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Fig. 4: Results RE Freila<br />

Fig. 5: Results RE Salada<br />

114


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Fig. 6: Results RE Belerda<br />

Discussion<br />

The estimation of the contributing areas for the experimental runoff shows that these are<br />

considerably smaller than the catchments of the rills. This leads to the conclusion that the<br />

runoff used within these experiments represents a low magnitude-high frequency event.<br />

Therefore, the measured amounts of erosion and hydraulic dynamic have to be valued as low.<br />

This should be remembered when considering or comparing sediment concentrations.<br />

The sediment concentrations measured in our experiments are interesting and important when<br />

compared with previous experiments. Polyakov and Nearing (2003) tested, in laboratory<br />

experiments, the behaviour of rills in al loamy substrate under different inflow quantities (6<br />

and 9 L min -1 ) and rill lengths. The sediment concentrations in their experiment reached<br />

values between 10 and 130 g L -1 . Loch (2000) tested the influence of vegetation on soil<br />

erosion. In one of his plots, both inter-rill and rill erosion were observed. The sediment<br />

concentrations of this plot reached a maximum value of 210 g L -1 . The maximum value<br />

measured in our rill experiments in Andalusia was 437.6 g L -1 , which is about double the<br />

maximum value in the experiments of Loch (2000) and about 3.5 times higher than the values<br />

in the laboratory experiments of Polyakov and Nearing (2003). This is remarkable because, in<br />

115


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

laboratory experiments, the rills are formed in a relatively loose substrate while the rills tested<br />

in our experiments developed in natural soils or on heavily compacted areas. In our<br />

experiments, the sediment concentrations showed a large range, from a low value of 5.2 g L -1 ,<br />

measured in Belerda, to the highest value (437.6 g L -1 ) in Freila. Sediment concentrations<br />

used to parametrize WEPP ranged from 0.1 to 415 g L -1 (Elliot et al., 1989). Our results show<br />

that in Mediterranean sedimentary basins, on abandoned fields, the erosion rates may exceed<br />

the upper limits of the parametrization data set of WEPP. Taking into account the low runoff<br />

used here, it can be assumed that the parameterization dataset does not cover the range of soils<br />

that can be found in practice. This confirms the assumption of Sidorchuk (2009), who<br />

postulates that most empirical data used for establishing erosion models is too narrow to cover<br />

the variability of erosion controls and processes.<br />

The reason for such a large range of sediment concentrations could be the fact that several<br />

processes take place in the rill: the transport of lose material, erosion or both processes<br />

together. Erosion processes can be separated into deepening of the rill's bottom caused by the<br />

shear forces of water (and transported sediments), but the main sediment load is supplied by<br />

other processes. Rill erosion is the result of the combination of different processes including<br />

headcut erosion, sidewall sloughing, tunnelling, micro-piping, slaking piping and sapping<br />

(Bryan et al. 1989, Bryan 1990, Knapen et al. 2007, Owoputi & Stolte 1995, Rapp 1998, Zhu<br />

et al. 1995). Zhu et al. (1995) concluded in laboratory experiments that the contribution of<br />

headcutting in detachment processes was four times larger than the contribution of bed scours.<br />

Kohl (1988) found that headcutting accounted for up to 60% of total rill erosion for the soil<br />

erodibility data of WEPP. <strong>Stefan</strong>ovic and Bryan (2009) tested in laboratory experiments the<br />

development of rills on a loamy sand and on a sandy loam. They also showed that<br />

concentrated flow causes sediment production primarily from knickpoints, chutes, meanders<br />

and bank failure. Govers (1987) distinguished between hydraulic erosion, mass wasting<br />

processes on rill sidewalls, gullying and piping. During his study in Huldenberg, the loamy<br />

hilly region of Flandres, the field was conventionally tilled and a seedbed was prepared.<br />

Hydraulic rill erosion mostly occurs during three observed runoff events with peak discharges<br />

between 70 and 90 L s -1 (4200 – 5400 L min -1 ). Runoff rates during other events were always<br />

below 25 L s -1 (1500 L min -1 ). We used runoff rates of 9 L min -1 however this rate turned out<br />

to be too low to produce hydraulic rill erosion. With our runoff rates, we can cause mass<br />

wasting processes on rill sidewalls. In the observed runoff events (Govers 1987), mass<br />

wasting processes caused 37% of total erosion in rills. But in erosion modelling, rill erosion is<br />

considered to be only dependant of the erosive power of the flowing water, represented by<br />

116


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

shear stress, unit length shear force or stream power. The process of gullying, the retreat<br />

erosion at knickpoints and headcuts is not considered in rill erosion formulas. This process<br />

caused about 12% of rill erosion rates in the study of Govers (1987). In our experiments, we<br />

only cause mass wasting and gullying processes, so the relations between hydraulic<br />

parameters and sediment concentration are mostly low. But the hydraulic rill erosion only<br />

occurs in extreme runoff events, in most cases, the runoff values are too low to cause this<br />

process (Govers 1987). All these observations agree with our own observations and<br />

measurements. In these cases, we can also observe that the areas with high erosion rates<br />

exceed the calculated transport capacities and show a trend towards an increase with flow<br />

duration. This suggests that sediment production by knickpoints, chutes, etc. is important for<br />

understanding rill behaviour. Nevertheless, they are highly variable throughout a single rill,<br />

and the combination of the processes can change between run a and run b of one experiment.<br />

The different processes can be identified by comparing the sediment concentrations, the flow<br />

velocities and the absolute level of the sediment concentrations.<br />

The first process is the “bulldozer-effect” (Seeger et al., 2005, Regües et al., 2000), a<br />

mobilisation of the loose material available in the rill before an experiment starts. The highest<br />

sediment concentration at the MP is collected with the water front despite the lowest flow<br />

velocity. This effect can be detected in the first run of the Negratin experiment. The flow<br />

velocity in run a and run b are close but in the first run the average sediment concentration is<br />

clearly higher. The material transport within the second run has to be attributed only to the<br />

erosion effect of the flowing water (bulldozer-effect (marginal) + erosion (low)), since the<br />

loose material has removed in the first run, mostly by the water front (bulldozer effect (high)<br />

+ erosion (low)). In run a, the bulldozer-effect is more powerful than the erosion process. If<br />

the erosion process is more powerful, the sediment concentrations increase at the MP from the<br />

first to the last sample. The waterfront only mobilizes the available loose material, but<br />

because of the low flow velocity, the erosion process starts later and the sediment<br />

concentration increases (bulldozer-effect (lower) + erosion (higher)). This can be observed in<br />

the first run of the Freila experiment. In the second run, the sediment concentrations are<br />

lower, because the bulldozer-effect is able to mobilize only the material left behind after run a<br />

(bulldozer-effect (marginal) + erosion (high)). If there is no available loose material and the<br />

erosion process is very active from the beginning, the sediment concentrations increase at<br />

each measuring point and the run with the higher flow velocity shows the higher average<br />

sediment concentration. This was given in the Salada experiment: run a (bulldozer-effect<br />

(marginal) + erosion (high)) had a lower average flow velocity and a lower average sediment<br />

117


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

concentration than run b (bulldozer-effect (marginal) + erosion (even higher)) with a higher<br />

average flow velocity and also a higher average sediment concentration. In some rills the<br />

erosion process could not be activated with the inflow intensity and water quantity used. Even<br />

the higher flow velocity in the second run was not sufficient to start erosion processes. In this<br />

case the available loose material is removed by the slowly flowing waterfront of the first run.<br />

This situation occurred in the Belerda experiment, where the highest sediment concentrations<br />

were measured at the waterfront run a (bulldozer-effect (high) + erosion (marginal)), while in<br />

run b both the bulldozer-effect and erosion processes were very low (bulldozer-effect<br />

(marginal) + erosion (marginal)).<br />

The differences in runoff-length-coefficient between the two runs can be seen clearly. In<br />

Salada, the RC L factor in run b is about twice as high as in run a, in Salada even the factor 3 is<br />

observed. In Belerda, the differences between the two runs are not significant, but the absolute<br />

values show that this rill is an effective runoff collector for the gully. In Freila only the first<br />

run could be measured, but this first run had the highest RC L factor of all runs. This is likely<br />

due to the fact that the rills in this area show a low infiltration rate, which corresponds to the<br />

high bulk densities measured. It also shows that, in these areas, flow concentration and the<br />

runoff accelerating effect of the rills within the landscapes are considerable regardless of the<br />

erosion processes inside the rills.<br />

Differences in the flow velocities of the water fronts between the two runs are particularly<br />

noteworthy. The velocities of the waterfronts in the first runs are clearly lower than all other<br />

velocity curves. This may be attributed, in part, to the higher transporting activity the water<br />

has to perform during the first runs, combined with a higher infiltration rate in most cases,<br />

resulting in considerably higher loss of water mass during the experiment. But the other, and<br />

surely most important, factor is that of initial soil moisture as was reported also by Govers et<br />

al. (1990) and Govers (1991): higher soil moisture reduces the infiltration rate which results in<br />

more water flowing along the experimental length. Higher soil moisture at the surface layer of<br />

the soil reduces the water repellency of the soils. This can be important for the dynamics of<br />

the water front under dry conditions. This was observed, even under very moist conditions, in<br />

the Pyrenees (<strong>Wirtz</strong> et al., 2010). Here we have that the observed velocities during the first<br />

run may underestimate the runoff during a rainfall event, as the rain producing the runoff<br />

would moisten the rill at the same time, thus generating the conditions for fast flow.<br />

The influence of initial soil moisture on sediment transport was reported by Govers et al.<br />

(1990) and Govers (1991). In their experiments under dry conditions the sediment<br />

concentrations almost reached the runoff's transport capacity while, under higher initial soil<br />

118


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

moistures, the sediment concentrations were lower. In three of our rill experiments, the first<br />

run (dry conditions) showed higher sediment concentrations than the second run (wet<br />

conditions). Only in the experiment at Salada, were higher sediment concentrations reached in<br />

the second run. The differences between run a and run b are caused by the differences in the<br />

supply of loose material, which is normally higher in the first run, and the detachment, which<br />

is normally higher in the second run (higher flow velocity).<br />

The ranges of transport capacity calculated are within the same orders of magnitude within<br />

the rills as reported by Giménez and Govers (2002). Plotting the sediment flow rate Q s against<br />

the transport capacity T c shows some differences between the tested rills (Fig. 7 - 10). In<br />

Negratin and Belerda the measured values are below the calculated transport capacity, due to<br />

a limitation in sediment supply or sediment detachment. This is especially accentuated in the<br />

Belerda experiment. Both Negratin and Belerda have loam soils, however Belerda has in<br />

addition a high coarse fragment content and the higher organic matter content which may be a<br />

reason for the lower detachment rates. Within the tested rills of Freila and Salada, the<br />

observed sediment load is at least partially above the calculated sediment transport capacity.<br />

Here, transport of sediments within the rill should be limited by the transport capacity<br />

however gravitational processes like bank failure and headcut retreat at knickpoints are the<br />

main sediment producing processes rather than erosive deepening of the rill's bottom. As a<br />

result it is possible for the sediment load to surpass the transport capacity. As Q s and T c have<br />

been calculated based on the inflow into the rill, both values are doubtless lower than the<br />

values given here, but as T c decreases more quickly (flow shear stress with an exponent 1.5,<br />

Q S with an exponent 1) the relationship is expected to shift towards exceeding the transport<br />

capacity as runoff decreases. However a common observation in all experiments was the fact,<br />

that at each measuring point and at each run we found a wide range of values for the transport<br />

capacity and the sediment load. We even found shifts between exceeding the transport<br />

capacity and dropping below at the same measuring point. This is a clear indicator of the<br />

spatially (along the rill) and temporally (during different stages of the water flow) variable<br />

processes and process intensities. The variability of erosion processes was tested from<br />

different research groups. Nearing (1998) tested the variability between replicated soil erosion<br />

field plots under natural rainfall and determined the principal factor or factors which correlate<br />

to the magnitude of variability. The coefficient of variation ranged on the order of 14 % for a<br />

measured soil loss of 20 kg m -2 to grater than 150 % for a measured soil loss of less than 0.01<br />

kg m -2 . Ruttiman et al. (1995) analysed statistically the data of four sites, each with five to six<br />

reported treatments; each of them with three replications. The coefficients of variation in soil<br />

119


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

loss ranged from 3.4 % to 173.2 %, with an average value of 71 %. Wendt et al. (1986)<br />

measured soil erosion rates on 40 experimental plots. All plots were cultivated and treated<br />

identically. The coefficients of variation for 25 storms ranged from 18 % to 91 %, with a<br />

clearly decreasing variability of soil loss with increasing erosivity of the storms: 15 storms<br />

with erosion rates higher than 0.1 kg m -2 were noted and 13 showed coefficients of variation<br />

of less than 30 %. Risse et al. (1993) applied the USLE to a large data-set of plot-years and<br />

natural runoff plots. Annual values of measured soil loss averaged 3.51 kg m -2 , the average<br />

magnitude of prediction error was 2.13 kg m -2 , that means 60 %. Zhang et al. (1996) tested the<br />

WEPP-model in a similar way. The average soil loss was 2.18 kg m -2 with an average<br />

prediction error of 1.34 kg m -2 this means 61% of the mean. Liu et al. (1996) compared<br />

measured values with WEPP-calculated values. For one of the tested catchments, the<br />

sediment yield was underpredicted by approximately 50 %. Govers (1991) tested the rill<br />

erosion rates on arable land in Central Belgium. The relevant characteristics of the selected<br />

fields were similar; the highest standard deviation was 65 m in length and 25.5 % in sand<br />

content. Mean rill erosion rate from 156 measurements reached 0.36 kg m -2 , with a maximum<br />

of 3.5 kg m -2 , but also no erosion was observed. The results reflected here show the high<br />

variability of soil erosion measurements, even under controlled (this is experimental)<br />

conditions. This is partially the result of non-homogeneous parameters concerning soil<br />

characteristics and rainfall. On experimental plots, infiltration rates and soil aggregate<br />

stability can be highly variable (Ajayi & Horta 2007) as well as the rainfall shows a high<br />

spatial and temporal variability (Dunkerley 2008). Therefore, the input parameters to the<br />

different measurements reflected in the mentioned studies were not really comparable.<br />

Nevertheless, the results also make clear, that modelling soil erosion has to tackle with<br />

uncertainty in model input as well as in the data that can be used for model calibration and<br />

validation.<br />

In the field, the spatial and temporal variability of soil conditions cannot be avoided, and is,<br />

furthermore, part of the investigations. Therefore, additional input parameters as rainfall or<br />

flow should be maintained constant to generate reproducible data.<br />

120


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Fig. 7: Sediment flux vs. transport capacity RE Negratin<br />

Fig. 8: Sediment flux vs. transport capacity RE Freila<br />

121


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Fig. 9: Sediment flux vs. transport capacity RE Salada<br />

Fig. 10: Sediment flux vs. transport capacity RE Belerda<br />

Conclusions<br />

In Andalusia, four rill experiments were conducted. The aim of these experiments was to<br />

estimate the effectiveness of different rills as sediment sources and runoff pathways and to<br />

detect the processes having the most influence on the rills’ behaviour. To accomplish this, 72<br />

122


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

L of water were induced with an intensity of 9 L min -1 into the rill. Flow velocities, sediment<br />

concentrations, rill cross sections, water levels, runoff values and slope were measured and<br />

different hydraulic parameters were calculated. It was noticeable that, despite the low inflow<br />

intensity, extremely high sediment concentrations were reached in some cases, with the<br />

transport of loose material having in some cases a higher, and in other cases a lower,<br />

influence on the sediment budget. In one experiment, where only low erosion activity took<br />

place, the measured sediment concentrations were mainly caused by the removal of the loose<br />

material. The runoff effectiveness showed very different values, and the higher effectiveness<br />

was reached in each case under conditions of higher initial soil moisture.<br />

During the experiment, the main sediment delivering processes were observed to be retreat<br />

erosion at knickpoints or steps and bank failure processes. Depending on the tested rill, a<br />

simple cutting in the rill's bottom also played an important role. In some cases, the high<br />

sediment concentrations during the first run indicate that the mobilisation of loose material<br />

may be important too. This may lead to loose material dominating the amount of transported<br />

material unless the other processes described above occur.<br />

In most cases, the erosion rate was near the transport capacity. In each experiment, the<br />

transport capacities and the sediment loads showed high variability even at a single measuring<br />

point. In some cases, samples exceeded the traditionally calculated transport capacity. This<br />

behaviour is a clear indicator of variable processes and process intensities in the natural<br />

environment. The limitation of detachment or transport equations is that important sediment<br />

preparing processes, like headcut retreat, step-pool effects or bank failure, are not considered.<br />

Measured total rill erosion rates are the sum of erosion rates caused by a combination of<br />

different soil erosion processes with different spatial and temporal distribution. There are two<br />

different ways. The first way is to modify existing physical based model concepts or, the<br />

second way, to start a new direction, the concept of stochastic soil erosion modelling. In both<br />

cases, much more field experiments are needed to provide the required data and the question<br />

is, if the actual used experimental setups can deliver the requested data or if a totally new<br />

experimental setup must be created.<br />

The novel method for monitoring rill processes presented here allows a lot of data which are<br />

useful for understanding the behaviour of rills in a catchment area to be directly measured or<br />

calculated from the measured data.<br />

123


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Acknowledgements<br />

The research was supported by the Internationale Graduiertenzentrum of Trier University. We<br />

want to acknowledge the reviewers for their valuable comments. They were very helpful in<br />

clarifying the objectives and the discussion.<br />

References<br />

AD-HOC-Arbeitsgruppe Boden 2005. Bodenkundliche Kartieranleitung. 5. verbesserte und<br />

erweiterte Auflage, Hannover. Schweizerbart’sche Verlagsbuchhandlung, Stuttgart.<br />

Ajayi, A.E., Horta, I.D.M.F. 2007. The effect of spatial variability of soil hydraulic properties<br />

on surface runoff processes. Anais XIII Simpósio Brasileiro de Sensoriamento<br />

Remoto, Florianópolis, Brasil, 21-26 abril 2007, p. 3243-3248.<br />

Auerswald, K., Fiener, P. and Dikau, R. 2009. Rates of sheet and rill erosion in Germany - A<br />

meta-analysis. Geomorphology, 111(3-4), 182-193.<br />

Beuselinck, L., Govers, G., Hairsine, P., Sander, G. and Breynaert M. 2002. The influence of<br />

rainfall on sediment transport by overland flow over areas of net deposition. Journal of<br />

Hydrology, 257, 145-163.<br />

Brodie, I and Rosewell C. 2007. Theoretical relationships between rainfall intensity and<br />

kinetic energy variants associated with stormwater particle washoff. Journal of<br />

Hydrology, 340(1-2), 40-47.<br />

Bruno, C., Di <strong>Stefan</strong>o, C. and Ferro, V., 2008. Field investigation on rilling in the<br />

experimental Sparacia area, South Italy. Earth Surface Processes and Landforms, 33,<br />

263-279.<br />

Brunton, D.A. and Bryan, R.B., 2000. Rill network development and sediment budgets. Earth<br />

Surface Processes and Landforms, 25, 783-800.<br />

Bryan, R.B. 1990. Knickpoint evolution in rillwash. Catena 17, 111-132.<br />

Bryan, R. B. 2000. Soil erodibility and processes of water erosion on hillslope.<br />

Geomorphology, 32(3-4), 385-415.<br />

Bryan, R.B. and Poesen, J., 1989. Laboratory experiment on the influence of slope length on<br />

runoff, percolation and rill development. Earth surface processes and Landforms, 14,<br />

211-231.<br />

Bryan, R.B., Govers, G., Poesen, J. 1989. The concept of soil erodibility and some problems<br />

of assessment and application. Catena, 16 (4-5), 393-412.<br />

124


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Bryan, R.B., Hawke, R.M. and Rockwell, D.L., 1998. The influence of subsurface moisture<br />

on rill system evolution. Earth Surface processes and Landforms, 23, 773-789.<br />

Calvo -Cases, A., Gisbert, J.M., Palau, E. and Romero, M. 1988. Un simulador de lluvia<br />

portátil de fácil construcción. In: M. Sala, F. Gallart (Editors), Métodos y técnicas<br />

para la medición en el campo de processo geomorfológicos. pp. 6-15.<br />

Casali, J., Loizu, J., Campo, M., De Santisteban, L. and Alvarez-Mozos, J. 2006. Accuracy<br />

of methods for field assessment of rill and ephemeral gully erosion. Catena, 67(2),<br />

128-138.<br />

Cerdá, A. and García-Fayos, P. 1994-95. Relaciones entre las pérdidas de agua y semillas en<br />

zonas acarcavadas. Influencia de la pendiente. Cuadernos I Geográfica, 20-21, 47-63.<br />

Cerdan, O., Le Bissonnais, Y., Couturier, A., Bourennane, H. and Souchère, V. 2002. Rill<br />

erosion on cultivated hillslopes during two extreme rainfall events in Normandy,<br />

France. Soil and Tillage Research, 67(1), 99-108.<br />

De Santisteban, L.M., Casalì, J., Lòpez, J.J., Giràldez, J.V., Poesen, J. and Nachtergaele, J.,<br />

2005. Exploring the role of topography in small channelerosion. Earth Surface<br />

Processes and Landforms, 30, 591-599.<br />

Dunkerley, D. 2008. Rain event properties in nature and in rainfall simulation experiments: a<br />

comparative review with recommendations for increasingly systematic study and<br />

reporting. Hydrological Processes, 22, 4415-4435.<br />

Elliot, W.J., Liebenow, A.M., Laflen, J.M. and Kohl, K.D., 1989. A compendium of soil<br />

erodibility data from WEPP cropland soil field erodibility experiments 1987 and 1988.<br />

NSERL Report No. 3, The Ohio State University and USDA Agricultural Research<br />

Service.<br />

FAO 2006. Guidelines for soil description. 4. ed. Rome.<br />

Favis-Mortlock, D. 1998. A self-organizing dynamic systems approach to the simulation of<br />

rill initiation and development on hillslopes. Computers and Geosciences, Vol. 24, No.<br />

4, 353-372<br />

Favis-Mortlock, D., Boardman, J., Parsons, A.J. and Lascelles, B. 2000. Emergence and<br />

erosion: a model for rill initiation and development. Hydrological Processes, 14, 2173-<br />

2205.<br />

Flores-Cervantes, J. H., Istanbulluoglu, E. and Bras, R. L. 2006. Development of gullies on<br />

the landscape: A model of headcut retreat resulting from plunge pool erosion. J.<br />

Geophys. Res., 111, F01010, doi:10.1029/2004JF000226.<br />

125


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Foster, G.R., 1982. Modelling the erosion process. In: H.P. Johnson, D.L. Brakensiek<br />

(Editors), Hydraulic Modelling of Small Watersheds. ASAE monograph 5, Hann CT,<br />

St. Joseph, MI, pp. 295-380.<br />

Foster, G.R., Huggins, L.F. and Meyer, L.D., 1984. A laboratory study of rill hydraulics: I.<br />

Velocity relationsships. Transactions of ASAE, 27, 790-796.<br />

Foster, G., Flanagan, D., Nearing, M., Lane, L., Risse, L. and Finkner, S. 1995. Hillslope<br />

Erosion Component. USDA-Water Erosion Prediction Project (WEPP). West<br />

Lafayette, pp. 11.1-11.12.<br />

Gilley, J.E., Kottwitz, E.R. and Simanton, J.R., 1990. Hydraulic characteristics of rill.<br />

Transactions of the ASAE, 33, 1900-1906.<br />

Giménez, R., and Govers, G. 2002. Flow Detachment by Concentrated Flow on Smooth and<br />

Irregular Beds. Soil Sci Soc Am J, 66(5), 1475-1483.<br />

Giménez, R., Planchon, O., Silvera, N. and Govers, G., 2004. Longitudinal velocity patterns<br />

and bed morphology interaction in a rill. Earth Surface Processes and Landforms, 29,<br />

105-114.<br />

Govers, G. 1987. Spatial and temporal variability in rill development processes at the<br />

Huldenberg experimental site. Catena Supplement, 8, 17-34.<br />

Govers, G. 1991. Rill erosion on arable land in central Belgium: Rates, controls and<br />

predictability. Catena, 18, 133-155.<br />

Govers, G. 1992 a. Evaluation of transporting capacity formulae. In: A.J. Parsons, A.D.<br />

Abrahams (Editors), Overland flow – Hydraulics and erosion mechanics. UCL-Press,<br />

Guildford, pp. 243-273.<br />

Govers. G., 1992b. Relationship between discharge, velocity and flow area for rills eroding<br />

loose, non layered materials. Earth Surface Processes and Landforms, 17, 515-528.<br />

Govers, G. and Poesen, J., 1988. Assesment of the interrill and rill contributions to total soil<br />

loss from an upland field plot. Geomorphology, 1, 343-354.<br />

Govers, G., Giménez, R. and Van Oost, K. 2007. Rill erosion: Exploring the relationship<br />

between experiments, modelling and field observations. Earth-Science Reviews, 84(3-<br />

4), 87-102.<br />

Helming, K., Römkens, M.J.M., Prasad, S.N. and Somme, H., 1999. Erosional development<br />

of small scale drainage networks. In: S. Hergatren, H.J. Neugebauer (Editors), Process<br />

Modelling and Landform Evolution. Lecture Notes in Earth Science 78. Springer,<br />

Berlin, pp. 123-146.<br />

126


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Hessel, R. and Jetten, V. 2007. Suitability of transport equations in modelling soil erosion for<br />

a small Loess Plateau catchment. Engineering Geology, 91(1), 56-71.<br />

Huang, C., Bradford, J.M. and Laflen, J.M., 1996. Evaluation of the detachment-transport<br />

coupling concept in the WEPP rill erosion equation. Soil Science Society American<br />

Journal, 60, 734-739.<br />

Knapen, A., Poesen, J., Govers, G., Gyssels, G. and Nachtergaele, J. 2007. Resistance of<br />

soils to concentrated flow erosion: A review. Earth-Science Reviews, 80(1-2), 75-109.<br />

Kohl, K.D., 1988. Mechanics of rill headcutting. Ph.D. Thesis, Iowa State University, Ames.<br />

Kuhn, N. and Bryan, R. 2004. Drying, soil surface condition and interrill erosion on two<br />

Ontario soils. Catena, 57, 113-133.<br />

Kuhn, N., Bryan, R.B. and Navar, J. 2003. Seal formation and interrill erosion on a smectiterich<br />

Kastanozem from NE-Mexico. Catena, 52, 149-169.<br />

Lasanta, T., Pérez Rontomé, M.C. and García Ruiz, J.M. 1994. Efectos hidromorfológicos de<br />

differentes alternativas de retirada de tierras en ambientes semiáridos de la depressión<br />

del Ebro. In: J.M. García Ruiz, T. Lasanta (Editors), Efectos geomorfológicos del<br />

abandono de tierras. Sociedad Española de Geomorfología, pp. 69-82.<br />

Le Bissonnais, Y., Cerdan, O., Lecomte, V., Benkhadra, H., Souchère, V. and Martin, P.<br />

2005. Variability of soil surface characteristics influencing runoff and interrill erosion.<br />

Catena, 62, 111-124.<br />

Liu, B.Y., Nearing, M.A., Baffaut, C., Ascough II, J.C., 1996. The WEPP watershed model:<br />

III. Comparisons to measured data from small watersheds. Transactions of the ASAE,<br />

40, 945-951.<br />

Loch, R.J. 2000. Using Rainfall simulation to guide planning and management of<br />

rehabilitated areas: Part I. Experimental methods and results from a study at the<br />

northparkes mine, Australia. Land Degradation and Development, 11, 221-240.<br />

Mancilla, G.A., Chen, S. and McCool, D.K., 2005. Rill density prediction and flow velocity<br />

distribution on agricultural areas in the Pacific Northwest. Soil and Tillage Research,<br />

84, 54-66.<br />

Merz, W. and Bryan, R. B. 1993. Critical conditions for rill initiation on sandy loam<br />

Brunisols: laboratory and field experiments in southern Ontario, Canada. Geoderma,<br />

57(4), 357-385.<br />

Nachtergaele, J., Poesen, J., Sidorchuk, A. and Torri, D. 2002. Prediction of concentrated<br />

flow width in ephemeral gully channels. Hydrological Processes, 16, 1935-1953.<br />

127


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Nachtergaele, J., Poesen, J., Steegen, I., Takken, L., Beuselinck, L., Vandekerckhove, L.<br />

and Govers, G. 2001. The value of a physically based model versus an empirical<br />

approach in the prediction of ephemeral gully erosion for loess-derived soils.<br />

Geomorphology, 40, 237-252.<br />

Nearing, M.A., Bradford, J.M., Parker, S.C. 1991. Soil detachment by shallow flow at low<br />

slopes. Soil Science Society of America Journal, 55 (2), 339-344.<br />

Nearing, M.A., 1998. Why soil erosion models over-predict small soil losses and underpredict<br />

large soil losses. Catena, 32, 15-22.<br />

Nearing, M.A., Govers, G. and Norton, L.D. 1999. Variability in Soil Erosion Data from<br />

Replicated Plots. Soil Science Society of America Journal, 63, 1829-1835.<br />

Nearing, M. A., Norton, L. D., Bulgakov, D. A., Larionov, G. A., West, L. T. and<br />

Dontsova K. M. 1997. Hydraulics and Erosion in Eroding Rills. Water Resources<br />

Research, 33 (4), 865-876.<br />

Nearing, M.A., Foster, G.R., Lane, L.J. and Finkner, S.C., 1989. A process-based soil erosion<br />

model for USDA - Water Erosion Predict Project technology. Transactions of the<br />

ASAE, 32, 1587-1593.<br />

Owoputi, L.O., Stolte, W.J., 1995. Soil detachment in the physically based soil erosion<br />

process: a review. Transactions of the ASAE, 38 (4), 1099-1110.<br />

Poesen, J., De Luna, E., Franca, A., Nachtergaele, J. and Govers, G. 1999. Concentrated<br />

flow erosion rates as affected by rock fragment cover and initial soil moisture content.<br />

Catena, 36, 315-329.<br />

Poesen, J. 1987. Transport of rock fragments by rill flow – a field study. In: R.B. Bryan<br />

(Editor), Rill erosion - process and significance. Catena, Supplement 8, Cremlingen,<br />

pp. 35-54.<br />

Polyakov, V.O. and Nearing, M.A., 2003. Sediment transport in rill flow under deposition and<br />

detachment conditions. Catena, 51, 33-43.<br />

Rapp, I., 1998. Effects of soil properties and experimental conditions on the rill erodibilities<br />

of selected soils. Ph. D. Thesis, Faculty of Biological and Agricultural Sciences,<br />

University of Pretoria, South Africa.<br />

Regüés, D., Balasch, J.C., Castelltort, X., Soler, M., Gallart, F., 2000. Relación entre las<br />

tendencias temporales de producción y transporte de sedimentos y las condiciones<br />

climáticas en una pequeña cuenca de montaña mediterránea (Vallcebre, Pirineos<br />

Orientales). Cuadernos de Investigación Geográfica, 26, 41-65.<br />

128


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Rejman, J. and Brodowski, R., 2005. Rill characteristics and sediment transport as a function<br />

of slope length during a storm event on loess soil. Earth Surface Processes and<br />

Landforms, 30, 231-239.<br />

Ries, J.B., Langer, M. and Rehberg, C. 2000. Experimental investigations on water and wind<br />

erosion on abandoned fields and arable land in the central Ebro Basin, Aragón/Spain.<br />

Zeitschrift für Geomorphologie N.F. Suppl.-Bd. 121, 91-108.<br />

Ries, J.B., Seeger, M., Iserloh, T., Wistorf, S. and Fister, W. 2009. Calibration of simulated<br />

rainfall characteristics for the study of soil erosion on agricultural land. Special issue<br />

on "soil erosion and degradation on agricultural land "Soil and Tillage Research 106,1,<br />

109-116.<br />

Ries, J.B. 2003 Landnutzungswandel und Landdegradation in Spanien – Eine Einführung. In:<br />

I. Marzolff, J.B. Ries, J. De La Riva., M. Seeger (Editors), Landnutzungswandel und<br />

Landdegradation in Spanien, pp.11-29.<br />

Risse, L.M., Nearing , M.A.., Nicks, A.D., Laflen, J.M., 1993. Assessment of error in the<br />

universal soil loss equation. Soil Sci. Soc. Am. J. 57, 825-833.<br />

Ruttimann, M., Schaub, D., Prasuhn, V., Ruegg, W., 1995. Measurement of runoff and soil<br />

erosion on regulary cultivated fields in Switzerland – some critical considerations.<br />

Catena, 25, 127-139.<br />

Seeger, M. 2007. Uncertainty of factors determining runoff and erosion processes as<br />

quantified by rainfall simulations. Catena, 71(1), 56-67.<br />

Seeger, M., Errea, M.P., Beguería, S., Arnáez, J., Martí, C., García-Ruiz, J.M., 2004.<br />

Catchment soil moisture and rainfall characteristics as determinant factors for<br />

discharge/suspended sediment hysteretic loops in a small headwater catchment in the<br />

Spanish Pyrenees. Journal of Hydrology, 288, 299–311.<br />

Sidorchuk, A. 2005. Stochastic modelling of erosion and deposition in cohesive soils.<br />

Hydrological Processes, 19(7), 1399-1417.<br />

Sidorchuk, A. 2009: A third generation erosion model: The combination of probabilistic and<br />

deterministic components. Geomorphology, 110 (1-2), 2-10.<br />

Slattery, M.C. and Bryan, R.B., 1992. Hydraulic conditions for rill incision under simulated<br />

rainfall: a laboratory experiment. Earth Surface Processes and Landforms, 8, 97-105.<br />

Stroosnijder, L. 2005. Measurement of erosion: Is it possible?. Catena, 64, 162-173.<br />

Torri, D., Dfalanga, M. and Chisci, G., 1987. Threshold conditions for incipient rilling. In:<br />

R.B. Bryan (Editor): Rill Erosion. Catena Supplement, 8, 97-105.<br />

129


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Vandekerckhove, L., Poesen, J. and Govers, G. 2003. Medium-term gully headcut retreat<br />

rates in Southeast Spain determined from aerial photographs and ground<br />

measurements. Catena, 50, 329-252.<br />

Vandekerckhove, L., Poesen, J., Oostwoud Wijdenes, D. and de Figueiredo,T. 1998.<br />

Topographical thresholds for ephemeral gully initiation in intensively cultivated areas<br />

of the Mediterranean. Catena, 33(3-4), 271-292.<br />

Wendt, R.C., Alberts, E.E., Hjelmfelt, Jr., 1986. Variability of runoff and soil loss from<br />

fallow experimental plots. Soil Sci. Soc. Am. J. 50, 730-736.<br />

<strong>Wirtz</strong>, S., Seeger, M. and Ries, J.B. 2010. The rill experiment as a method to approach a<br />

quantification of rill erosion process activity. Zeitschrift für Geomorphologie, 54,1,<br />

47-64.<br />

Woodward, D. E. 1999. Method to predict cropland ephemeral gully erosion. Catena, 37(3-4),<br />

393-399.<br />

Zhang, G., Liu, B., Liu, G., He, X., Nearing, M.A., 2003. Detachment of undisturbed soil by<br />

shallow flow. Soil Science Society of America Journal 67, 713-719.<br />

Zhu, J.C., Gantzer, C.J., Peyton, R.L., Alberst, E.E., Anderson, S.H., 1995. Simulated smallchannel<br />

bed scour and head cut erosion rates compared. Soil Science Society of<br />

America Journal, 59 (1), 211-218.<br />

130


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Kapitel 5<br />

<strong>Wirtz</strong> et al. (2013): Do deterministic sediment detachment and transport equations<br />

adequately represent process-interactions in eroding rills? An experimental field study.<br />

Catena 101, 61-78.<br />

131


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Do deterministic sediment detachment and transport equations adequately represent the<br />

processes interactions in eroding rills? An experimental field study.<br />

<strong>Stefan</strong> <strong>Wirtz</strong> • Manuel Seeger • Alexander Remke • René Wengel • Jean-Frank Wagner •<br />

Johannes B. Ries<br />

Dep. of Physical Geography, Trier University<br />

Abstract<br />

This paper tackles two main questions by linking observations, determination of hydraulic<br />

parameters and measurement of sediment transport with the formulae used in soil erosion<br />

models. First, do constant shear stress values in different rills with constant soil parameters<br />

result in the same soil detachment values? Secondly, deterministic soil erosion models make<br />

the assumption that there is a relationship (often further assumed to be linear) between shear<br />

stress and soil detachment; is this suitable for representing real erosion process combinations<br />

in natural rills? Following most process based deterministic soil erosion models, derived<br />

hydraulic and erosion parameters should be similar. However, the results from the different<br />

experiments showed clear differences in sediment concentration, transport rates and other<br />

measured as well as calculated values. In contrast to our experimental results, a model<br />

simulation would produce erosion parameters with low variations, represented by the relative<br />

measurement error and the empirical variation coefficient. This reveals the general problems<br />

of using process based deterministic models for erosion in shallow rills. While soil erosion<br />

models simulate the processes resulting from the shear forces of flowing water on the soil<br />

surface, other processes like side wall failure, headcut retreat and plunge pool dynamics are<br />

not taken into account. Our results suggested that these other processes may contribute<br />

substantially to rill erosion processes. The results of this study strongly suggested that the<br />

model concept of most physical based soil erosion models is inadequate for modeling rill<br />

erosion processes. Measured total rill erosion rates were the sum of erosion rates caused by a<br />

combination of different soil erosion processes with different spatial and temporal<br />

distribution. This combination cannot be described by a single equation.<br />

Keywords Experimental field work • Process based soil erosion model • Rill erosion • Soil<br />

erosion<br />

132


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

1 Introduction<br />

Modelling the results of the processes of soil erosion by water is an important task for<br />

geomorphologic research. Soil erosion models are also used to evaluate land use systems, and<br />

derive guidelines for land use management. Model concepts can be empirical (e.g. USLE;<br />

Wishmeier and Smith, 1965, 1978) or based on physical process description (e.g. WEPP;<br />

Laflen et al., 1997; EUROSEM; Morgan et al., 1998; LISEM; De Roo et al., 1996). There are<br />

also stochastic approaches found in the literature for describing and modelling sediment<br />

detachment and transport (Einstein, 1937, Mirtskhoulava, 1988, Wilson, 1993a, Govindaraju<br />

and Kavvas, 1994, Lisle et al., 1998, Hairsine and Rose, 1991, Govers, 1991, Nearing, 1991,<br />

Shaw et al., 2008, Sidorchuk, 2002, Sidorchuk et al., 2004, Sidorchuk, 2005 a, Sidorchuk,<br />

2005 b, Sidorchuk et al., 2008, Sidorchuk, 2009). However to date no no practical, field-scale<br />

or catchment-scale, model exists which uses stochastic approaches. Different approaches are<br />

currently under construction; RillGrow 1 and RillGrow 2 are examples for model<br />

developments which represent detachment from a distributional perspective (Favis-Mortlock<br />

et al., 1998, 2000).<br />

What is the background of process based soil erosion models?<br />

Many current soil erosion models are described as physically based, or process-based. Knapen<br />

et al. (2007) distinguish between excess shear stress models and excess stream power models.<br />

In the first case, transport and detachment capacities are calculated using shear stress alone<br />

and in the other, shear stress is replaced by stream power which is the product of shear stress<br />

and flow velocity. In both cases, a critical soil parameter (critical shear stress or critical<br />

stream power) is exceeded by a hydraulic factor (shear stress or stream power) which enables<br />

an entrainment of soil particles. Shear stress is therefore a fundamental factor for the process<br />

based soil erosion models, describing the drag force exerted by the flow on the bed (Giménez<br />

and Govers, 2002). For inferring shear stress and the derived detachment and transport<br />

capacity in soil erosion modelling, a set of input parameters are typically used in slightly<br />

varying combinations: slope, liquid density, flow velocity, hydraulic radius, wetted perimeter,<br />

flow cross section and water depth (Knapen et al., 2007). Due to the deterministic nature of<br />

process based models, the same input parameters will always deliver the same results.<br />

Moreover, these models often assume a linear relation between shear stress and soil<br />

detachment (Lyle and Smerdon, 1965, Torri et al., 1987, Ghebreiyessus et al., 1994, Nearing<br />

et al., 1997).<br />

133


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Where have experiments been conducted and what questions have been addressed?<br />

Following Kleinhans et al. (2010) there are several ways for geoscientists to create results or<br />

data sets. These are (i) field observations, (ii) field experiments, (iii) laboratory experiments<br />

or (iv) a model simulation.<br />

Measurements of soil erosion have been conducted in both the field and experimentally in the<br />

lab. In laboratory experiments, the initial and boundary conditions can be well controlled. Soil<br />

parameters, rill forms and slope can be adapted to specific questions. In this way, physical<br />

laws can be tested in a relatively precise environment. However, Giménez and Govers (2002)<br />

showed that parameters determined under laboratory conditions cannot be easily transferred to<br />

natural environments. The main reason is that in laboratory experiments smooth beds are<br />

often used while in natural rills the beds are always irregular. Nonetheless, laboratory<br />

experiments have given detailed insight into different processes of soil erosion and have been<br />

used to derive empirical relationships between descriptors of the forces exerted by flowing<br />

water and sediment detachment and transport. For example, Brunton and Bryan (2000) and<br />

Bryan and Poesen (1989) showed in laboratory experiments that headcut incision and bank<br />

collapse are important processes in the development of rills and rill networks. But the<br />

consideration of these processes has not found its way into applied soil erosion modelling.<br />

Different research groups have often presented different shear stress – based factors like: unit<br />

length shear force (Giménez and Govers, 2002), stream power (Bagnold, 1977, Hairsine and<br />

Rose, 1992, Elliot and Laflen, 1993, Nearing et al., 1997, Zhang et al., 2003), unit stream<br />

power (Yang,1972, Moore and Burch, 1986) and effective stream power (Bagnold, 1980,<br />

Govers, 1992a), which show a more or less good fit with soil detachment rates measured in<br />

their experiments. These experimentally deduced factors are highly dependent on the<br />

experimental setup, due to a none standardised experimental method available. Casali et al.<br />

(2006) clearly shows the consequences of comparing the results from different methods.<br />

These results are often used in process based soil erosion models. As a result, soil erosion<br />

models which make use of these experimentally deduced factors have a clear empirical<br />

foundation (Stroosnijder, 2005). However, this is not always adequately acknowledged.<br />

The United States Department of Agriculture (USDA) made great efforts in field<br />

experimental research for developing the Universal soil loss equation USLE (Flanagan et al.,<br />

2003).<br />

Nevertheless, most field data about runoff and erosion in rills have been gained from<br />

observations during natural rainfall and runoff events or long-term-plot measurements; in only<br />

few cases controlled experiments have been conducted (Abrahams and Li, 1996, De<br />

134


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Santisteban et al., 2005, Helming et al., 1999, Parsons et al., 1999, Parsons and Wainwright,<br />

2006, Rejman and Brodowski, 2005). The aim of these studies was:<br />

- to observe rill network formation (Bruno et al., 2008, Mancilla et al., 2005),<br />

- to define the initial conditions for rilling (Bruno et al., 2008, Bryan et al., 1998,<br />

Govers and Poesen, 1988, Slattery and Bryan, 1992, Torri et al., 1987),<br />

- to study the development of rill head morphology (Bruno et al., 2008, Brunton and<br />

Bryan, 2000),<br />

- to estimate the main hydraulic variables like cross-section area, wetted perimeter,<br />

hydraulic radius, mean velocity and shear stress for calculating other hydraulic<br />

parameters which could not be measured or estimated (Abrahams and Li, 1996, Bruno<br />

et al., 2008, Foster et al., 1984, Gilley et al., 1990, Giménez et al., 2004, Govers, 1992<br />

b), to validate existing models (Huang et al., 1996),<br />

- to measure the transmission losses in rills (Parsons et al., 1999) or<br />

- to propose mathematical models for estimating soil loss due to rill erosion (Bruno et<br />

al., 2008, Foster, 1982, Nearing et al., 1989) or for rill initiation (Parsons and<br />

Wainwright, 2006).<br />

Field data are as close to reality as possible. But observations can be collected from a long<br />

term plot measurement or only by scientists who are in the right place at the right time. This is<br />

not always possible, so field experiments are used to trigger the processes to be observed<br />

when the measurement team is on site. However, both observations and experiments have<br />

certain disadvantages: (i) Measurement techniques may disturb the processes being observed,<br />

(ii) the time scale of human observations is shorter than that of the process under study, (iii)<br />

some processes cannot easily be measured and (iv) some processes are apparently chaotic and<br />

their spatial and temporal variations are difficult to capture (Favis-Mortlock and de Boer,<br />

2003, Kleinhans et al., 2010).<br />

Nevertheless, simplified experimental setups are ruled by the same natural laws as the<br />

processes found in nature. Thus, experimental observations can be considered as a simplified<br />

but valid representation of the reality (Paola et al., 2009). A model is always simpler than the<br />

thing which it aims to represent (Favis-Mortlock et al., 2001).<br />

Mathematical models describe reality in terms of mathematical equations. Physical laws are<br />

often simplified to allow numerical solutions, but in many cases it is not clear which laws<br />

apply and to what extent simplification is possible. Model parameters are always the result of<br />

simplifications. Values of some parameters are not well known, so the model has to be<br />

calibrated. Here, the phenomenon of equifinality is a problem: A wide range of parameters<br />

135


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

can produce the same result. Or, as Favis-Mortlock et al. (2001, p. 487) state, models can give<br />

“the right answer for the wrong reason”. Another problem is that parameters for rill hydraulics<br />

are often taken from equations for describing flow behaviour in rivers. Govers (1992 a) and<br />

Govers et al. (2007) showed that these equations are not suitable for rill erosion processes.<br />

Therefore, there is often a mismatch between model results and observed or measured<br />

“reality” (Kleinhans et al., 2010).<br />

What is our scientific question?<br />

As discussed above, models can give rise to large uncertainties. This is, in some cases, the<br />

result of poor quality or inappropriate input data. At the same time, field experiments deliver<br />

reliable data, as processes run under natural conditions. Field experiments enable a direct<br />

observation of the processes involved in rill dynamics. Hence, to bridge the gap between<br />

models, parameterisation and observations, the field experiments in this study were performed<br />

with specific attention given to the basic hydraulic parameters needed to calculate shear<br />

stress. By linking the observations, the determination of hydraulic parameters and the<br />

measurement of sediment transport with fundamental formulae used in soil erosion models,<br />

we aimed to tackle the following questions:<br />

1. Do constant shear stress values in different rills on the same field with constant<br />

soil parameters (critical shear stress for example) always result in the same soil<br />

detachment values?<br />

And as a consequence of this question:<br />

2. Is the model concept of a linear relation between shear stress and soil detachment<br />

suitable for real erosion process combinations in natural rills?<br />

The characteristics of the field site and the experimental set up allowed a set of experiments<br />

with very similar boundary conditions to be conducted under natural conditions. All necessary<br />

parameters could be measured directly, thus avoiding the need for estimates. As a<br />

consequence, it was expected that the shear stress exerted by the flowing water would remain<br />

within the same order of magnitude for all experiments.<br />

2 Materials and methods<br />

2.1 Study area<br />

The Natural Park Bardenas Reales, a 425 km² semiarid landscape, is located in northeast<br />

Spain (Navarra), in the Ebro-Basin (Fig. 1). It is bounded by the Aragón River on the north<br />

and the Ebro on the south (Desir and Marín, 2007, Murelaga et al., 2002, Sancho et al., 2008).<br />

136


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

The tertiary and quaternary sediments in the Bardenas Reales consist of open playa-lake<br />

deposits, red, grey and green clayey marl and pockets of lacustrine, limestone-like, sandy and<br />

gypsum-containing sediments as well as massive marly and lacustrine limestones that form<br />

cuestas (Desir and Marín, 2007, Murelaga et al., 2002, Sancho et al., 2008). The climate in the<br />

test area is semi-arid and characterised by irregular, heavy rainfall events with an average of<br />

380 mm a -1 , the average annual temperature is 19.2°C, and the potential evapotranspiration<br />

rate reaches 1084 mm. In summer and winter, only sporadic rainfall events reach the test<br />

areas; the convective precipitation occurs in spring and autumn (Causapé et al., 2006,<br />

Causapé et al., 2004, Desir and Marín, 2007, Sancho et al., 2008).<br />

Fig. 1 Location of the Bardenas Reales<br />

The position of our experimental site was 624817 E 4681907 N (UTM Zone 30). With highdefinition<br />

aerial photography it was possible to estimate area and length values and to<br />

describe soil surface characteristics such as vegetation cover or rills.<br />

On a part of the 0.75 ha field with nearly bare soil, i.e. vegetation cover was only 1% and rock<br />

fragments cover 2%, approximately 20 furrows have developed into rills in different stages of<br />

development (Fig. 2). Average rill length was 10 m, thus we calculated a total rill length of<br />

200 m. The relationship between rill length and area (rill density) was 267 m ha -1 . This is<br />

lower than on the fields which were treated by the ‘ripping’ technique (a kind of deep<br />

137


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

loosening) where furrows were opened but higher than on a ploughed field in the study of<br />

Hagmann (1996). The informations about area size and rill lengths have been derived from<br />

self-made large scale aerial photographs. The camera used for acquiring the photos was a<br />

Canon 350D with a Canon EF-S 20 mm objective. The maximum resolution of the photos<br />

was 8 MP, the resulting stereoscopic images showed a ground resolution between 0.5 and 11<br />

cm, depending on the flying height. We used a low flight level, so the ground resolution of<br />

our photos is at about 1 cm. This is accurate enough for calculating rill lengths and area sizes.<br />

Other informations like cross sections or rill width are not calculated from aerial<br />

photographies but directly measured in the field. For these tasks, the ground resolution would<br />

not be appropriate.<br />

Four of these rills were randomly chosen for the experiments.<br />

Fig. 2 Field with the tested rills: The crawler as scale in the photo on the right has a length of<br />

about 1.5m. RE = rill experiment<br />

On our test site, texture class was at the threshold between poor silty sand and poor loamy<br />

sand, gravel content was 1%. Main texture class was fine and very fine sand, nearly half of<br />

the fine soil material showed a grain size between 0.2 and 0.063 mm. Grain density and dry<br />

bulk density have been estimated as 2.65 g cm -3 and 1.68 g cm -3 , respectively, following the<br />

138


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Ad-hoc-Arbeitsgruppe Boden (2005). Soil material contained 2.2% organic matter, and the<br />

starting soil moisture was 7.5%. The tested rills developed in this homogeneous substrate<br />

show neither steps nor plungepools but had different slopes and flow length. The rill<br />

descriptors with different values for each rill are summarized in Table 1, the constant setup,<br />

soil and climate factors are presented in Table 2.<br />

Table 1 Rill descriptors and values by test rill. MP = Measuring Point<br />

Factor Rill 1 Rill 2 Rill 3 Rill 4<br />

Tested flow length [m] 10 14 6 6.5<br />

Ø Slope [°] 4.3 3.1 2 2.9<br />

Max. Slope [°] 9.4 5.8 3.1 7.1<br />

Min. Slope [°] 1.8 1.2 0.7 1<br />

Standard deviation slope [°] 2.1 1.3 0.9 1.9<br />

MP 1 position [m] 3 4 1.5 2<br />

MP 1 slope [°] 1.8 3.5 2.8 3.7<br />

MP 1 depth [m] 0.109 0.115 0.075 0.105<br />

MP 1 width [m] ~ 0.52 ~ 0.56 ~ 0.25 ~ 0.28<br />

MP 2 position [m] 6 7 3 3.5<br />

MP 2 slope [°] 4.6 3 3.1 1.6<br />

MP 2 depth [m] 0.115 0.189 0.076 0.101<br />

MP 2 width [m] ~ 0.6 ~ 0.52 ~ 0.28 ~ 0.20<br />

MP 3 position [m] 9 12 4.5 5<br />

MP 3 slope [°] 3.3 2.1 1.5 1<br />

MP 3 depth [m] 0.108 0.161 0.085 0.093<br />

MP 3 width [m] ~ 0.6 ~ 0.42 ~ 0.28 ~ 0.24<br />

139


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Table 2 Constant parameters<br />

Parameter Factor Value<br />

Discharge intensity [ L min -1 ] 9<br />

Setup Discharge quantity [L] 72<br />

Discharge time [min] 8<br />

Soil texture<br />

Loamy sand<br />

Organic matter [%] ~ 2.2<br />

Land use<br />

Arable land<br />

Transport Coefficient K t [s 2 m 0.5 kg -0.5 ] 0.0107<br />

Soil Vegetation cover [%] ~ 1<br />

Rock fragment cover [%] ~ 2<br />

Starting soil moisture [%] ~ 7.5<br />

Grain density [g cm -3 ] 2.65<br />

Dry bulk density [g cm -3 ] 1.68<br />

Average precipitation [mm a -1 ] 380<br />

Climate<br />

Average annual temperature [°C] 19.2<br />

Evapotranspiration rate [mm a -1 ] 1084<br />

Characterisation<br />

Semi arid<br />

2.2 Rill experiment<br />

The rill experiment consisted of two runs: in the first run the rill was tested under dry<br />

conditions; in a second run, 15 min later, the same rill was tested under wet conditions,<br />

therefore the influence of the initial soil moisture could be considered.<br />

For each run, a motor driven pump was used to maintain a constant discharge of 9 L min -1 for<br />

8 minutes, resulting in a total water flow of 72 L (<strong>Wirtz</strong> et al., 2010, <strong>Wirtz</strong> et al., 2012). In the<br />

used range of water quantity and pressure the fluctuations were very low so we can assume<br />

that the inflow was stable.<br />

The flow velocity within the rill was characterized by the travel time of the waterfront and of<br />

two colour tracers (food colourings (E 124 red and E 13 blue started at 3 and 6 minutes after<br />

the start of the experiment) measured for every meter using a chronograph. By means of this<br />

procedure, three velocity curves with the maximum flow velocities were recorded and<br />

changes in flow dynamics could be detected. Following Govers (1992b) we assumed that the<br />

mean flow velocity is reached, where the waterfront reaches 80 – 90 % of its maximum width.<br />

In our experiments, the difference between the front and this point could not easily be<br />

140


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

distinguished and the time difference between these two points defined our measurement<br />

accuracy because of the relatively diffuse form of the waterfront and the flow velocity. The<br />

measured flow velocities were not equal to the flow velocities at sampling time. We assumed<br />

a linear increase or decrease between the three measured flow velocities so we could calculate<br />

the velocities between those points to get the velocities for the sampling time. This procedure<br />

is shown in Fig. 3.<br />

Fig 3 Calculation of the flow velocities<br />

At the end of the rill, the runoff was continuously measured by a pressure transducer (Ecotech<br />

DL/n, V2.35). For calibration of the discharge curve, runoff at the outflow was measured<br />

volumetrically at regular intervals. This allowed constant measurement of the discharge at<br />

high temporal resolution and throughout the whole experiment.<br />

The rill's slope was measured using dividers with a range of 1m and a digital spirit level as it<br />

was used in <strong>Wirtz</strong> et al. (2010). It is important to note that slope measurement provided only<br />

an average slope over the 1 meter.<br />

At three measuring points (MPs) along the rill four water samples were taken: the first as soon<br />

as the waterfront reached the sampling point, the second 30 seconds later, and the third and<br />

fourth after 1:30 min and 2:30 min total time from arrival of the water. The measurement of<br />

sediment passing given points is measurements of flux. Changes in flux at different points and<br />

at different times enable the identification of processes leading to spatial and temporal sources<br />

and sinks of sediment (Parsons et al., 2006). The sediment concentration was determined by<br />

filtration of the samples in the laboratory (<strong>Wirtz</strong> et al., 2010). The reason for this sampling<br />

141


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

scheme was that we could not collect the whole water at a MP, so we had to take aliquots. At<br />

the beginning we expected the highest fluctuations, hence the distance between sample 1 and<br />

2 was lower than between sample 2 and 3 and sample 3 and 4. Depending on the flow length<br />

to the MP it was possible that after 3 minutes the runoff stopped at this point, we had to take<br />

the samples before this occurred.<br />

At each MP, rill cross section was measured. With thin metal sticks, the distance between<br />

ground level and rill bottom was measured in 0.002 m steps. This allowed an accurate<br />

calculation of the rill’s cross section area and an estimation of the rill’s volume. Additionally,<br />

the water depth at a certain point in the rill was measured several times using a yardstick. By<br />

combining the rill cross section and the water depth it was possible to calculate flow area,<br />

wetted perimeter and hydraulic radius using the georeferencing tool of a GIS.<br />

2.3 Descriptors for soil detachment<br />

The relationship between detachment rate and detachment capacity and respectively the<br />

relationship between transport rate and transport capacity are important values for assessing<br />

different processes acting in the rill. These variables were calculated as follow:<br />

DC = K<br />

r<br />

(τ τcr<br />

) (Foster et al., 1995) Eq. (1)<br />

T<br />

1.5<br />

C<br />

= K<br />

t<br />

R τ<br />

(modified from Wagenbrenner et al., 2010) Eq. (2)<br />

SC<br />

V A<br />

DR =<br />

Eq. (3)<br />

L WP<br />

TR = SC<br />

V A<br />

Eq. (4)<br />

with D C = detachment capacity [kg m -2 s -1 ], K r = rill erodibility factor [s m -1 ], τ = shear stress<br />

[Pa], τ cr = critical shear stress [Pa], T C = transport capacity [kg s -1 ], K t = transport coefficient<br />

[s² m 0.5 kg -0.5 ], R = hydraulic radius [m], D R = detachment rate [kg m -2 s -1 ], S C = sediment<br />

concentration [g L -1 ], V = flow velocity [m s -1 ], A = flow area [m²], L = flow length [m], WP<br />

= wetted perimeter [m], T R = transport rate [kg s -1 ].<br />

The rill erodibility factor K r can be calculated for cropland or rangeland; in this case, we used<br />

the equation of Flanagan and Livingston (1995) for cropland. Parameters are very fine sand<br />

content (vfs) and organic material content (org.Mat.).<br />

K<br />

1.84(org.<br />

Mat.<br />

)<br />

r<br />

= 0.00197 + 0.0003 (vfs) + 0.03863<br />

Eq. (5)<br />

Depending on land use, the parameters for the critical shear stress are clay content and very<br />

fine sand content (Flanagan and Livingston, 1995).<br />

τ c<br />

= 2.67 + 0.065 (clay) 0.058 (vfs)<br />

Eq. (6)<br />

142


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Shear stress was calculated as follow:<br />

τ = ρ g R S (Giménez and Govers, 2002) Eq. (7)<br />

with ρ = fluid density [kg m -3 ], g = gravitational acceleration (9.81 m s -2 ), R = hydraulic<br />

radius [m] and S = effective slope (sin(slope angle)).<br />

The transport coefficient value was taken from the WEPP-database (Elliot et al., 1989). We<br />

used the value of the soil that is most similar to the soil in the test area. In this case, the<br />

location is Amarillo, the K t value is 0.0107 s 2 m 0.5 kg -0.5 .<br />

For comparing the variability of the different parameters, we calculated first the average of<br />

the relative measurement errors (RME) following the DIN 1319-1 (1995). This error is<br />

defined as<br />

xa<br />

xr<br />

RME = 100<br />

Eq. (8)<br />

x<br />

r<br />

with x a as measured value and x r as “correct” value, we used the mean of the measured values<br />

as x r . The variance and the standard deviation are still in the unit of each parameter, therefore<br />

the variances of different parameters are not comparable.<br />

The RME describes the difference between the average value of all experiments and the<br />

average values of one run. From our measuring values, we calculated the average value of<br />

each run. The average of these “run-values” is the x r in Eq. (8), the average of each run is<br />

presented by the x a . The given values in Table 4 are the average of the values of the 8 runs.<br />

The RME would be much higher if we would calculate the RME between average values of<br />

all 96 samples and the single, directly measured sample.<br />

As second value for describing the variation, we calculated the empirical variation coefficient<br />

EVC. It was calculated as follows:<br />

standard<br />

deviation<br />

EVC =<br />

average<br />

Eq. (9)<br />

The calculation for this factor is similar to the calculation of the RME values: We did not use<br />

the mean standard deviation and the average of all samples, because of the different<br />

conditions we calculated the average values for each run and averaged these results.<br />

3 Results<br />

Sediment concentrations at the MPs<br />

We use a number-letter-system to identify our runs. The first number is the number of the<br />

experiment, a or b means the first or the second run and the last number is the number of the<br />

measuring point.<br />

143


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Generally we could distinguish between seven different sediment concentration curves<br />

(Figure 4-7). (1) The first trend was defined by a decrease from the first to the last sample,<br />

only one sample with increasing value disturbed this trend. This form was found in 1a1, 1a2,<br />

1b1, 2a2, 2a3 and 3a3. (2) In the second form, the decreasing trend was not disturbed by any<br />

increasing value. This trend was shown by 1a3. (3) The next trend type was defined by an<br />

increase from the first to the second sample, after that, the values remained almost constant:<br />

1b2, 2b1, 3a1, 3b1, 3b2, 3b3, 4a2, 4b1, 4b2, and 4b3. (4) It was also possible that the values<br />

stayed about constant from the first to the last sample, this form was found in 1b3 and 4a3. (5)<br />

In 2a1, the values increased from sample 1 to sample 3, after that the value decreased at<br />

sample 4. (6) In version 6, we observed a decrease from sample 1 to sample 2 and an increase<br />

from sample 2 up to sample 4. This was found in 2b2, 3a2 and 4a1. (7) The last version found<br />

at 2b3 showed a decrease from sample 1 to sample 2 followed by a constant trend which was<br />

disturbed by one sample with increasing value.<br />

Flow velocities along the rill profile<br />

In experiment 1 and 4, the flow velocity curves showed similar forms: In the first runs, the<br />

two tracers are similar over the rill profile; the water front didn’t reach the same velocity. In<br />

run 2, the velocity of the waterfront increased and tracers and water front showed similar<br />

values. In experiment 3, the differences between the tracers and the front were also low in the<br />

second run, but the roughly parallel trend of the tracers in run 1 was disturbed in run 2. In the<br />

first run of experiment 2, the tracers showed similar values, the waterfront showed lower<br />

values again. But in run 2, the water front showed a very irregular trend with fewer<br />

similarities to the tracer curve. Because of technical problems, the velocity of tracer one in<br />

experiment 2b was not measured.<br />

Runoff values at the end of the tested rill part<br />

In most cases, the increase of the runoff curve is very fast, only run 1 in experiment 1 showed<br />

a slower increase. In all cases, the runoff in run 1 started later as in the second run and the<br />

values were lower, this difference was very noticeable in the first experiment. Also the runoff<br />

duration was longer in the second runs of all experiments. Because of technical problems, the<br />

runoff of run 1 in experiment 4 was not measured but the observations confirmed the<br />

measurements of the other experiments.<br />

144


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Fig. 4 Results of RE 1<br />

145


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Fig. 5 Results of RE 2<br />

146


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Fig. 6 Results of RE 3<br />

147


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Fig. 7 Results of RE 4<br />

Average values for the different runs:<br />

All factors for each run of the four REs are presented in Table 3. Whether values were<br />

measured or calculated is also noted. The average values in this table were calculated from 12<br />

samples (3 MPs x 4 samples in each run).<br />

148


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Table 3 Average values of all factors for all experiments. Shear stress 1 includes the sediment<br />

concentration and grain density in the liquid density calculation, shear stress 2 is calculated<br />

using a constant liquid density of 1 g cm -3 . c = value is calculated, equation number is given,<br />

m = value is measured. SSC = Sediment concentration. The r-l-factor describes the<br />

relationship between runoff and flow length: inflow/runoff * length.<br />

Factor Unit m/c 1a 1b 2a 2b 3a 3b 4a 4b<br />

Transport rate [kg s -1 ] c 4 0.0055 0.0057 0.0112 0.0152 0.0006 0.0004 0.0015 0.0013<br />

SSC [g L -1 ] m 16.32 13.53 44.02 37.51 2.56 1.78 10.19 9.41<br />

Detachment rate [kg s -1 m -2 ] c 3 0.0051 0.0054 0.0123 0.0146 0.0007 0.0006 0.0033 0.0037<br />

Shear stress 1 [Pa] c 7 7.17 7.15 5.88 6.24 3.64 4.75 2.77 3.5<br />

Shear stress 2 [Pa] c 7 7.09 7.09 5.73 6.12 3.64 4.75 2.76 3.49<br />

Hydraulic Radius [cm] m 1.28 1.28 1.16 1.19 0.90 1.13 0.74 0.87<br />

Slope [°] m 3.2 3.2 2.9 2.9 2.4 2.4 2.1 2.1<br />

Flow velocity [m s -1 ] m 0.15 0.18 0.16 0.21 0.14 0.17 0.16 0.19<br />

Liquid density [g cm -3 ] m 1.010 1.008 1.027 1.023 1.002 1.001 1.006 1.006<br />

r-l-factor [m] m 1.27 4.25 6.39 7.95 3.09 4.29 ND 3.34<br />

Runoff coefficient [%] m 12.68 42.47 45.61 56.78 51.56 71.54 ND 51.33<br />

RE1 showed the highest average values in both shear stress forms (constant liquid density and<br />

liquid density including sediment concentration and grain density), in the hydraulic radius<br />

values and in slope. Lowest values were measured for r-l-factor and runoff coefficient (1a).<br />

In RE2 the highest values for transport rate (2b), sediment concentration (2a), detachment rate<br />

(2b), flow velocity (2b), liquid density (2a) and r-l-factor (2b) were measured.<br />

RE 3 only showed the highest runoff coefficient (3b). Instead of highest values, many lowest<br />

values have been measured or calculated: transport rate, sediment concentration, detachment<br />

and liquid density in 3b and flow velocity in 3a.<br />

RE4 provided the lowest shear stress values, the lowest hydraulic radius (4a) and the lowest<br />

slope.<br />

In a model simulation, in most cases average values are used. The external parameters are<br />

constant on the whole field (see Table 2), so the 4 different rills would be described by one<br />

mean value. We calculated mean values of all samples (4 experiments x 2 runs x 3 MPs x 4<br />

samples = 96 samples) which are presented in Table 4. To compare the variability of the<br />

different parameters, we also calculated the relative measurement error (RME) and the<br />

empirical variation coefficient (EVC) for each of them. The RME values for the tested<br />

parameters are presented in Fig. 8, the EVC values in Fig. 9 and they are summarized in Table<br />

4.<br />

149


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Fig. 8 The relative measurement errors of the tested parameters: Shear stress 1 includes the<br />

sediment concentration and grain density in the liquid density calculation, shear stress 2 is<br />

calculated using a constant liquid density of 1 g cm –3<br />

Fig.9 The empirical variation coefficient of the tested parameters: Shear stress 1 includes the<br />

sediment concentration and grain density in the liquid density calculation, shear stress 2 is<br />

calculated using a constant liquid density of 1 g cm –3<br />

150


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Table 4 Variability of different runoff and erosion factors, hydraulic and rill parameters.<br />

RME is the relative measurement error. EVC is the empirical variation coefficient Shear<br />

stress 1 includes the sediment concentration and grain density in the liquid density<br />

calculation, shear stress 2 is calculated using a constant liquid density of 1 g cm –3 . Uncertainty<br />

= standard deviation * data quantity -1/2<br />

Factor Average RME [%] EVC Uncertainty<br />

Transport rate [kg s -1 ] 0.0052 81.7 0.69 0.009<br />

Sediment Concentration [g L -1 ] 16.9 70.5 0.54 1.793<br />

Detachment rate [kg s -1 m -2 ] 0.0057 67.5 0.74 0.001<br />

Shear stress 1 [Pa] 5.1 28.6 0.33 0.299<br />

Shear stress 2 [Pa] 5.1 28 0.33 0.295<br />

Hydraulic radius [cm] 1.1 16.4 0.09 0.035<br />

Slope [°] 2.7 14.7 0.28 0.104<br />

Flow velocity [m s -1 ] 0.2 10.2 0.27 0.005<br />

Liquid density [g cm -3 ] 1.01 0.7 0.006 0.001<br />

The highest average RME was calculated for the transport rate. RME values of more than<br />

60% were also calculated for sediment concentration and detachment rate. RME values<br />

between 20 and 40% were calculated for both shear stress variations. RME values below 20%<br />

were measured for flow velocity and for the input parameters of the shear stress equation<br />

hydraulic radius, slope and liquid density.<br />

The comparison of the EVC (empirical variation coefficient) showed similar results: High<br />

EVCs were calculated for transport and detachment rate and the sediment concentration. The<br />

shear stress values showed again middle values, the flow velocity and the slope were in the<br />

EVC-ranking in a middle position, in the RME-ranking they were in a back position. Low<br />

EVCs were determined for the hydraulic radius and the liquid density.<br />

Concerning the ranges of the RME values (Fig. 8) it is notable that the erosion parameters<br />

transport and detachment rate and sediment concentration showed the highest range between<br />

maximum and minimum value as well as between the 75 % and the 25 % quantile. Middle<br />

ranges (here defined as values between 20 and 40 %) between maximum and minimum values<br />

were determined in both shear stress variations, hydraulic radius and flow velocity, low<br />

ranges < 20 % in slope and liquid density. Regarding only the range between the 75 and the<br />

151


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

25 % quantile, nearly all parameters showed a low range, only the shear stress with constant<br />

liquid density showed a middle range.<br />

Regarding the ranges of the EVC (Fig. 9) it is notable that all parameters showed the same<br />

range class between maximum and minimum value as well as between the both quantiles. The<br />

detachment rate showed high ranges, transport rate, sediment concentration, hydraulic radius,<br />

slope and flow velocity middle ranges and low ranges were determined for both shear stress<br />

variations and the liquid density.<br />

4 Discussion<br />

Errors in measurement<br />

The statistical uncertainty of the different parameters (Table 4) showed the highest value in<br />

sediment concentration, contrasting, and the flow velocity showed a very low uncertainty.<br />

However, a formal statistical analysis of errors is problematic for this study, since extreme<br />

high or low values in sediment concentration would be regarded as outliers, and perhaps<br />

discarded or discounted. But, from our observations, these extreme values result from a<br />

process change and thus should certainly not be ignored. This problem is also found in the<br />

model concepts: The processes which cause the extreme values (rill bank failure for example)<br />

are not described by the used equations. Regarding the measurement methods for the<br />

sediment concentration and the flow velocity, the number of possible errors in velocity<br />

measurement is much higher than in sediment concentration measurement. The problem of<br />

flow velocity measurement was described for example by Govers (1992b) and Ali (2011). In<br />

some cases, a constant correction factor was used to calculate the mean flow velocity from the<br />

often measured maximum flow velocity of the waterfront. This factor changed depending on<br />

experimental setup, Govers (1992 b) determined 0.94. Ali (2011) analysed the variability of<br />

the correction factor depending on grain size, testing four different grain sizes in the sand<br />

group. The correction factors were 0.44, 0.77 and two times 0.82. Other authors found out that<br />

the correction factor depends on slope and Reynolds number (Li et al., 1996) or on slope,<br />

Reynolds number and sediment discharge (Zhang et al., 2010)<br />

Rate vs. Capacity<br />

Theoretically, the relationship between a rate and a capacity should not exceed 1 because<br />

sedimentation processes reduce this value when the rate is higher than the capacity (Scherer,<br />

2008). However, in our experiments, 67 of 96 samples (75%) showed higher transport rates<br />

than transport capacities, especially in higher sediment concentration ranges (Fig. 10).<br />

152


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

The transport rate exceeded the transport capacity in 77% of the samples (dry runs)<br />

respectively in 63% (wet runs) of the samples. But regarding the average values of the<br />

relationships between transport rate and transport capacity (only the values higher than 1) the<br />

second runs (b-runs) showed a higher value than the first runs (a-runs) (7 vs. 5.7). In the a-<br />

runs, the transport of loose material was the main process, in the b-runs the bank failure and<br />

the headcut retreat. Sediment supply was not limited due to the low flows during the<br />

experiment. Fig. 11 shows clearly that still at low runoff values the transport rate was much<br />

higher than the transport capacity. In most of the cases, the transport of loose material was<br />

responsible for the high transport rates.<br />

Fig.10 Relationship between transport rate and transport capacity vs. sediment concentration.<br />

The equilibrium line is shown<br />

153


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Fig. 11 Relationship between transport rate and transport capacity vs. runoff. The equilibrium<br />

line is shown<br />

Variability in soil erosion measurements<br />

The input parameters for calculating shear stress showed average RME values below 20%:<br />

hydraulic radius with 16.4%, slope with 14.7% and liquid density with 0.7%. The calculated<br />

shear stress values were also similar; the variability was 28.6% when liquid density with<br />

sediment concentration considered was used and 28% if the liquid density of clear water was<br />

assumed. In models, a commonly-used idea is that there is a linear relation between shear<br />

stress and soil detachment volumes. This means that the variability of soil loss parameters<br />

should also be in the same order of magnitude. For transport and detachment rate, input<br />

parameters are still flow velocity and sediment concentration. Flow velocity showed a low<br />

variability of 10.2%. On the other hand sediment concentration (70.5%), transport rate<br />

(81.7%) and detachment rate (67.5%) showed very high variability, all values being over<br />

60%. This variation in values contrasts with the approach used in models. In a model<br />

simulation the variations of the input parameters and the soil loss parameters would be<br />

similar. This means that in this case, there was no simple linear relation between shear stress<br />

and soil detachment.<br />

The results of the research reported by Govers (1991), Liu et al. (1996), Nearing (1998, 1999<br />

a), Risse et al. (1993), Ruttiman et al. (1995), Wendt et al. (1986) and Zhang et al. (1996)<br />

underscore the problems. Nearing (1998) tested the variability between replicated soil erosion<br />

154


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

field plots under natural rainfall to determine the principal factor or factors which correlate<br />

with the magnitude of variability. The coefficient of variation ranged in the order of 14% for a<br />

measured soil loss of 20 kg m -2 to greater than 150% for a measured soil loss of less than 0.01<br />

kg m -2 . Ruttiman et al. (1995) analysed statistically the data of four sites, each with five to six<br />

reported treatments; each of them with three replicates. The coefficients of variation in soil<br />

loss ranged from 3.4% to 173.2%, with an average value of 71%. In a classic study, Wendt et<br />

al. (1986) measured soil erosion rates on 40 experimental plots. All plots were cultivated and<br />

treated identically. The coefficients of variation for 25 storms ranged from 18% to 91%, with<br />

a clearly decreasing variability of soil loss with increasing erosivity of the storms: 15 storms<br />

with erosion rates higher than 0.1 kg m –2 were noted and 13 showed coefficients of variation<br />

less than 30%. Risse et al. (1993) applied the USLE to a large data-set of plot-years and<br />

natural runoff plots. Annual values of measured soil loss averaged 3.51 kg m –2 , the average<br />

magnitude of prediction error was 2.13 kg m –2 , that means 60%. Zhang et al. (1996) tested the<br />

WEPP-model in a similar way. The average soil loss was 2.18 kg m –2 with an average<br />

prediction error of 1.34 kg m -2 which is 61% of the mean. Liu et al. (1996) compared<br />

measured values with WEPP-calculated values. For one of the tested catchments, the<br />

sediment yield was under-predicted by approximately 50%. Govers (1991) tested the rill<br />

erosion rates on arable land in Central Belgium. The relevant characteristics of the selected<br />

fields were similar; the highest standard deviation was 65 m in length and 25.5% in sand<br />

content. Mean rill erosion rate from 156 measurements reached 0.36 kg m –2 , with a maximum<br />

of 3.5 kg m –2 , but also no erosion was observed. The cited results are mostly from plot<br />

measurements which often do not show any rills but the results reflected here show the high<br />

variability of soil erosion measurements, even under controlled, experimental conditions.<br />

Different processes<br />

At most MPs, the sediment concentration showed a decreasing trend while the flow velocity<br />

increased. That means the available loose material was washed out. This process has been<br />

called “bulldozer-effect” (Seeger et al., 2004, Regüés et al., 2000, <strong>Wirtz</strong> et al., 2012) and was<br />

detectable in 1a3, 1b1, 3a1, 3b1, 3b2, 3b3, 4b1 and 4b2 (The number after the experiment<br />

number is the MP again). This trend could be disturbed by some higher sediment<br />

concentration peaks which represent in most cases side wall failures or knickpoint<br />

undercutting and retreat (1a1, 1a2, 1b2, 2a1, 2a2, 2a3, 3a3). Another observed process was the<br />

combination of the bulldozer-effect and increasing incision into the bottom of the rill. This<br />

combination was detectable by a constant (starting with the first or also with the second<br />

155


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

sample at the MP) or even increasing sediment concentration (1b3, 2b1, 2b2, 3a2, 4a1, 4b3).<br />

The combination of all three processes is also possible, first the bulldozer-effect, followed by<br />

incision in the rill’s bottom disturbed by a sidewall failure (2b3, 4a2, 4a3).<br />

The very large variability is likely to be in part the result of non-homogeneous parameters<br />

used for soil characteristics and rainfall. On experimental plots, infiltration rates and soil<br />

aggregate stability can be highly variable (Ajayi and Horta, 2007). Even simulated rainfall<br />

shows a high spatial and temporal variability (Assouline, 2009, Dunkerley, 2008, Lascelles et<br />

al., 2000) on plot size thus, we can assume that natural rainfall events show at least a similar<br />

variability. The variability is likely higher under natural rainfall because rainfall simulators<br />

are mostly constructed to create spatially uniform rainfall conditions. Therefore, the input<br />

parameters for the different measurements reflected in the studies mentioned above were not<br />

really comparable. Nevertheless, the results also make clear that modelling soil erosion<br />

involves uncertainty in model input as well as in the data that can be used for model<br />

calibration and validation. In the field, the spatial and temporal variability of soil conditions<br />

cannot be avoided, and is, furthermore, part of the investigations. Therefore, additional input<br />

parameters such as rainfall or flow should be kept constant in the experiments to generate<br />

comparable data. There is a high variability in soil erosion processes that cannot be<br />

represented by a single factor e.g. shear stress. The shear stress equation implies that drag<br />

forces are the dominant forces controlling erosion. But rill erosion is the result of the<br />

combination of different processes including headcut erosion, sidewall sloughing, tunnelling,<br />

micro-piping, slaking, piping and sapping (Bryan et al., 1989, Bryan, 1990, Knapen et al.,<br />

2007, Owoputi and Stolte, 1995, Rapp, 1998, Zhu et al., 1995). These processes are not<br />

explicitly accounted for in shear stress equations, although there is ample evidence to support<br />

their importance. Zhu et al. (1995) concluded from laboratory experiments that the<br />

contribution of headcutting in detachment processes was four times higher than the<br />

contribution of bed scours. Kohl (1988) found that headcutting accounted for up to 60% of<br />

total rill erosion for the soil erodibility data of WEPP. <strong>Stefan</strong>ovic and Bryan (2009) tested, in<br />

laboratory experiments, the development of rills on a loamy sand and on a sandy loam and<br />

showed that concentrated flow causes sediment production primarily from knickpoints,<br />

chutes, meanders and bank failure. Govers (1987) distinguished between hydraulic erosion,<br />

mass wasting processes on rill sidewalls, gullying and piping. During his classic study in<br />

Huldenberg, the loamy hilly region of Flandres, the field was conventionally tilled and a<br />

seedbed was prepared. Hydraulic rill erosion mostly occurred during three observed runoff<br />

events with peak discharges between 70 and 90 L s –1 (4200 – 5400 L min –1 ). Runoff rates<br />

156


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

during other events were always below 25 L s –1 (1500 L min -1 ). We used runoff rates of 9 L<br />

min –1 which turned out to be too low to produce hydraulic rill erosion. However, it was<br />

sufficient to cause mass wasting processes on rill sidewalls and exceeding transport capacity.<br />

In the observed runoff events (Govers, 1987), mass wasting processes caused 37% of total<br />

erosion in rills. But in erosion modelling, rill erosion is considered to be only dependent on<br />

the erosive power of the flowing water, represented by shear stress, unit length shear force or<br />

stream power. The retreat erosion at knickpoints and headcuts (in Govers (1987) called<br />

“gullying”), is not considered in rill erosion equations. This process caused about 12% of rill<br />

erosion rates in the study of Govers (1987). In our experiments, we only cause mass wasting<br />

and headcut retreat, so the relations between hydraulic parameters and sediment concentration<br />

are mostly low. However, hydraulic rill erosion only occurs in extreme runoff events<br />

suggesting that, in most cases, runoff values are too low to cause this process (Govers, 1987,<br />

Nearing, 1991). All these observations agree with our own observations and measurements. In<br />

addition, there are several studies and models looking at headcut retreat (Bennett, 1999,<br />

Robinson et al., 2000, Alonso and Bennett, 2002, Bennett and Alonso, 2005, Bennett and<br />

Alonso, 2000, Flores-Cervantes et al., 2006, Gordon et al., 2007) and bank failure (Parker,<br />

1983, Kovacs and Parker, 1994). These results have not (yet) been applied in soil erosion<br />

models. Some of the interacting sub-processes in an eroding rill are summarized in Kinnell<br />

(2005) and especially in figure 1 of this paper. He distinguished between raindrop detachment<br />

with transport by raindrop splash, raindrop detachment with transport by raindrop-induced<br />

flow transport, raindrop detachment with transport by flow and flow detachment with<br />

transport by flow. In our experiments, we only activated the last process, all raindropdependant<br />

processes were not represented in our experimental setup. Following Kinnell<br />

(2005) the flow detachment does not occur unless a critical value is exceeded. As we just<br />

measured the effect of one sub-process, our sediment concentrations would have been higher<br />

in a real rainfall event. The hydraulic parameters and hence the results of a model simulation<br />

would not have been affected.<br />

Soil loss and the hydraulic parameters<br />

Knapen et al. (2007) calculated the correlation of shear stress, unit length shear force, stream<br />

power and Reynolds number with the detachment capacity using several WEPP datasets. The<br />

best average correlation was determined for stream power with R² = 0.59. The R² values for<br />

the shear stress variable used in the WEPP datasets were never strong. Knapen et al. (2007, p.<br />

80f.) describes the shear stress as follows: ”Although the use of flow shear stress as soil<br />

157


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

detachment predictor can be contested, critical shear stress (τ cr ) and concentrated flow<br />

erodibility KC (...) have been selected as the most universal parameters to describe soil<br />

erosion resistance to concentrated flow.“ The correlations between these factors and the soil<br />

detachment capacities in our study showed very different results. There was not a single<br />

parameter that always showed the best correlation.<br />

Other groups have found linear correlations between a hydraulic parameter and an erosion<br />

parameter as well as in laboratory flume experiments (Partheniades and Paaswell, 1970,<br />

Nearing et al., 1997, Ghebreiyessus et al., 1994, Torri et al., 1987, Giménez and Govers,<br />

2002) and in field research (Elliot et al., 1989, Nearing et al., 1997, Nearing et al., 1999 b).<br />

We could not confirm these observations, in our field experiments there was no linear<br />

relationship between any hydraulic parameter and any erosion parameter.<br />

In a laboratory study, Nearing et al. (1991) measured flow shear stresses ranging from 0.5 to 2<br />

Pa, while tensile strengths ranged from 1 to 2 kPa, a difference in magnitude of 1000. Despite<br />

this, detachment rates of nearly 300 g m –2 s –1 were measured. Nearing explained this as being<br />

the result of turbulent burst events which are much greater than the average flow shear<br />

stresses. Nearing and Parker (1994) further investigated the influence of turbulence on shear<br />

stress. They showed that under turbulent flow conditions the same shear stress value results in<br />

a clearly higher detachment rate. Differences between detachment rates caused by turbulent<br />

and laminar flow increased with increasing shear stress value.<br />

This means that the influence of turbulence on soil erosion will be higher when hydraulic<br />

conditions lead to a high shear stress value than when shear stress values are in the lower<br />

range. This can be the reason that soil erosion models often over-predict small soil losses and<br />

under-predict large soil losses. Nearing (1998) explained this by the presence of natural<br />

variations in model data, which the model is not capable of capturing. These variations will<br />

affect a bias in the erosion predictions relative to values on the higher end versus those on the<br />

lower end of the range of measured values (Nearing 1998).<br />

Origin of the used equations<br />

Another reason for wich the use of the shear stress equation as it is commonly used does not<br />

deliver satisfactory results is the origin of this equation. The shear stress equation is deduced<br />

from the Navier-Stokes equations which describe the motion of fluids. The equations arise<br />

from applying Newton’s second law to fluid motion (net force on a particle is equal to the<br />

time rate of change of its linear momentum in an inertial reference frame), combined with the<br />

assumption that the fluid stress is the sum of a diffusing viscous term plus a pressure term.<br />

158


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Using one of the Navier-Stokes equations, an incompressible fluid can be completely<br />

described; thus reducing hydrodynamic questions to a mathematical problem. But this<br />

problem consists of a system of second order nonlinear partial differential equations which<br />

require the most powerful computers to numerically solve even the easiest cases. For the<br />

general, 3-dimensional case, existence- uniqueness- and regularity statements are not yet<br />

proven. Indeed, the Clay Mathematics Institute (CMI) included this in their “Millennium<br />

Prize Problems” which represent the most important open problems in mathematics, and has<br />

offered a prize of US$ 1,000,000 for a solution or a counter example (Constantin, 2001,<br />

Fefferman, 2006, Seiler, 2002, Schneider, 2008, Temam, 2000, Wiegner, 1999).<br />

The inconsistencies between the experimental results and the process based model<br />

assumptions can be the consequence of several reasons, such as uncertainties in measurements<br />

on the one hand or, on the other hand, inconsistent and incomplete process representations<br />

within the models. Uncertainty in the measurement of soil erosion has been a strong point of<br />

discussion (Stroosnijder, 2005), which is certainly still not fully resolved. The experimental<br />

setup applied in this study aimed to minimize systematic errors and their propagation. The<br />

design of the inlet, the flume for runoff measurement and the monitoring of flow and<br />

sediment transport reduced disturbance to a minimum (<strong>Wirtz</strong> et al., 2010, <strong>Wirtz</strong> et al., 2012).<br />

The results of measurements were also within the range measured in other experiments (e.g.<br />

Knapen et al., 2007). This, in combination with qualitative process observations made during<br />

the experiments, allows us to draw the conclusion that the source of errors is found in the<br />

model concepts.<br />

In models, different parameters play an important role. It is to distinguish between the<br />

hydraulic parameter flow shear stress τ and the soil parameter critical shear stress τ c or τ cr .<br />

Shear stress exerted by flow must exceed the critical shear stress to cause erosion. In critical<br />

shear stress calculations, soil parameters like dry bulk density, grain size distribution and<br />

organic content are used, in shear stress calculation hydraulic parameters, roughness, flow<br />

velocity and fluid density are variables. Different research groups deleted or added different<br />

factors to adapt the equation to their research topic. As a consequence, the equations for<br />

calculating transport and detachment capacity, critical shear stress and shear stress have<br />

different forms in different publications.<br />

We assumed that shear stress or critical shear stress values are always defined in the unit [Pa<br />

= kg m -1 s -2 ]. Note that in each paper where the units of the variables were not always given,<br />

we assumed [Pa] when no other unit was defined by the author and calculated the unknown<br />

units of the used parameters considering this default. For a better comparison with the original<br />

159


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

literature, we did not homogenize the labelling of the different parameters, but used the ones<br />

given by the authors. That means different parameters can be labelled equally. As it can be<br />

seen in Table 5, for the calculation of shear stress exerted by flowing water, many different<br />

parameters have been taken into account. Even the physical definition of the parameter “shear<br />

stress” seems to be unclear. Some factors have been developed from empirical studies<br />

(equation 10, equation 11). In most cases, the theoretical basis of the equations is not clear.<br />

Equation 12 is derived from Landau and Lifchitz (1971). The critical shear stress is the force<br />

needed to detach a soil particle, so it corresponds to a soil parameter and therefore input for<br />

calculation should also be depending on soil characteristics. In the WEPP model (Table 6,<br />

equation 26-28), the critical shear stress is calculated using only soil parameters like texture,<br />

organic matter content and dry bulk density. However in some equations the so called<br />

“critical” shear stress consists of hydraulic parameters like water depth, water width or fluid<br />

density (Table 6, equations 17-20). In equation 21 hydraulic and soil parameters are used<br />

equally. In equation 22, 23 and 24 the empirical nature of the development is clear. According<br />

to equation 23 and 24 the critical shear stress is only dependent on a constant value and the<br />

relationship between given particle size and the subsurface d 50 . In equation 26, the typical<br />

parameters for calculating shear stress of flowing water are used to define critical shear stress.<br />

This method uses additionally the Manning friction factor, which is a descriptor of the soils<br />

surface roughness. Thus, the use of the term “critical shear stress” seems to be very unclear.<br />

The different equations for detachment and transport capacity were developed from data sets<br />

created from controlled laboratory experiments, field observations or field experiments. In<br />

equations 29-32 (Table 7), neither critical shear stress nor shear stress is used for calculation<br />

of the transport capacity. In equation 33, shear stress is used to calculate transport capacity, in<br />

equation 34 transport rate is calculated using shear stress and critical shear stress and in<br />

equation 35 the detachment capacity is calculated using critical shear stress and shear stress<br />

(see Table 7). Equation 33 is a modification of the Yalin (1963) equation from 1963 (Foster et<br />

al., 1995). In equation 34 and 35 it is clear to see that shear stress and critical shear stress are<br />

opponents, the important parameter is the difference between these two variables. If critical<br />

shear stress is higher than shear stress, no erosion can occur. A summary of these equations is<br />

given by Reid and Dunne (1996), on the EPA-homepage (2009) and in Hessel and Jetten<br />

(2007).It has been shown that the physical definition of the parameters is not clear and their<br />

mostly empirical foundation does not fit the high temporal and spatial variability of processes<br />

within a rill. Different processes dominate at different intensities and this fact causes high<br />

160


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

variability in sediment concentration, transport and detachment rates. But the spatial and<br />

temporal distribution of the different processes is apparently highly random.<br />

Table 5 Equations for calculating shear stress.<br />

Eq.<br />

Equation<br />

10<br />

τ = γ<br />

y<br />

b<br />

y<br />

(<br />

y<br />

a<br />

s b 2<br />

eff<br />

p<br />

i<br />

C<br />

it<br />

)<br />

τ = shear stress [Pa], γ = weight density of water (force/volume) [Pa], y = flow depth assuming<br />

laminar flow [m], b = time weighting factor in finite difference equation for continuity, s = sine<br />

of slope angle, y b /y p = the ratio of the flow depth on a smooth surface to that in the ponds from<br />

depressions and “dams” [m m -1 ], a = a coefficient to be estimated i eff = effective rainfall<br />

intensity [m s -1 ]<br />

C<br />

it<br />

y<br />

p<br />

= exp [ 0.21 ( 1)]<br />

y<br />

Foster (1982)<br />

11 = a (ρ ρ) g D<br />

12<br />

τ<br />

s<br />

b<br />

1.18<br />

ρ S = specific weight of the sediment [kg m -3 ], ρ = specific weight of the water [kg m -3 ] D = the<br />

particle size [m], a = an empirical factor between 0.039 and 0.09<br />

Shields (1936), Miller et al. (1977), Parker et al. (1982), Diplas (1987), Parker (1990), Komar<br />

(1987 a,b), Andrews (1983), Ashworth and Ferguson (1989 a,b), Komar and Carling (1991)<br />

τ = (σ<br />

g)<br />

2<br />

3<br />

( 3<br />

υ)<br />

1<br />

3<br />

*sin<br />

2<br />

3<br />

α<br />

q<br />

1<br />

3<br />

σ = fluid density [kg m -3 ], g = gravitation [9.81 m s -2 ], υ = kinematic viscosity [m 2 s -1 ]<br />

α = slope angle, q = runoff discharge rate per unit of width [kg m -1 s -1 ]<br />

Chisci et al. (1985)<br />

13 = σ g R* tan(γ)<br />

τ r<br />

τ r = runoff shear stress [Pa], σ = fluid density [kg m -3 ], g = acceleration of gravity [9.81 m s -2 ], R<br />

=hydraulic radius [m], γ = slope angle [°]<br />

Torri et al. (1987)<br />

γ<br />

τ =<br />

h<br />

L<br />

14 L<br />

R<br />

15<br />

γ = unit density of water [kg m -3 ], h L = head loss due to friction [m 2 s -2 ], R = hydraulic radius<br />

[m], L = channel length [m]<br />

Ghebreiyessus et al. (1994)<br />

τ<br />

s<br />

= ρ<br />

w<br />

g S R<br />

f<br />

f<br />

s<br />

tot<br />

ρ w = water density [kg m -3 ], g = gravitation factor [9.81 m s -2 ], S = slope, R = hydraulic radius<br />

[m], f s and f tot = Darcy-Weisbach friction factors for the bare soil and composite surface,<br />

161


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

respectively<br />

Nearing et al. (1997)<br />

16 τ = ρ g R S<br />

ρ = density of the fluid [kg m -3 ], g = gravitation factor [9.81 m s -2 ], R = hydraulic radius [m], S =<br />

sin(slope)<br />

Giménez and Govers (2002)<br />

Eq.<br />

17<br />

18<br />

19<br />

20<br />

21<br />

τ b<br />

τ b<br />

τ w<br />

τ w<br />

= γ<br />

= γ<br />

= γ<br />

= γ<br />

S<br />

S<br />

S<br />

S<br />

Table 6 Equations for calculating critical shear stress.<br />

D<br />

( 1<br />

B D<br />

for<br />

4 B<br />

D<br />

B<br />

for 0<br />

D<br />

) for 0<br />

B<br />

1<br />

2<br />

D<br />

B<br />

1<br />

2<br />

D<br />

B<br />

B B D<br />

( 1 ) for<br />

2 4 D B<br />

Equation<br />

1<br />

2<br />

1<br />

2<br />

τ b = critical bed shear stress, τ w = critical walls shear stress [Pa], γ = fluid density [kg m -<br />

3 ], S = slope [m m -1 ], D = water depth [m], B = water width [m] In Eq. 17, Ott and van<br />

Uchelen (1936) uses instead of D/2 the term D/B.<br />

Shields (1936), Ott and van Uchelen (1936)<br />

τ<br />

c<br />

D S<br />

γs<br />

γ<br />

( )<br />

γ<br />

d<br />

D = mean depth [m], S = water surface slope [m m -1 ], γ = specific weight of fluid [kg m -<br />

3 ], γ S = specific weight of sediment [kg m -3 ], d = particle diameter [m]<br />

22<br />

23<br />

Graf (1971)<br />

τ<br />

α n<br />

= ( )<br />

1.486<br />

2<br />

c<br />

γ w<br />

τ c = critical boundary shear stress [pounds per square foot], α = an empirical factor, n =<br />

Manning’s friction coefficient [m 1/3 s -1 ], γ w = unit weight of water [pounds per square<br />

foot]<br />

Partheniades and Paaswell (1970)<br />

τ<br />

c<br />

di<br />

= 0.0834 ( )<br />

d<br />

50<br />

0.872<br />

162


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

24 di<br />

0.887<br />

τc<br />

= 0.0384 ( )<br />

d50<br />

d i<br />

d 50 = relationship between given particle size and the subsurface d 50<br />

Andrews (1984), Andrews and Erman (1986)<br />

25<br />

26<br />

27<br />

28<br />

τ<br />

cr<br />

= ρ<br />

(ρ g<br />

u<br />

2<br />

cr<br />

n<br />

0.66<br />

= (ρ<br />

q<br />

g R<br />

0.66<br />

S<br />

0.7<br />

S)<br />

)<br />

cr<br />

cr<br />

ρ = density of the water [kg m -3 ], u cr = shear velocity [m s -1 ], g = gravitation factor [9.81<br />

m s -2 ], R = hydraulic radius [m], S = sin (slope), n = Manning friction factor [m 1/3 s -1 ], q<br />

= unit discharge [m 7/6 s]<br />

De Ploey (1990)<br />

τ c<br />

= 2.67 + 0.65 CLAY 0.058 VFS<br />

for cropland > 30 % sand<br />

τ c<br />

= 3.5<br />

for cropland < 30 % sand<br />

τ = 3.23 0.056 SAND 0.244 ORGMAT + 0.9 for rangeland<br />

c<br />

BD dry<br />

Clay = clay content [%], VFS = very fine sand content [%], SAND = sand content [%],<br />

ORGMAT= organic matter content [%] (1.724 times organic carbon content), BD dry =<br />

dry soil bulk density [g cm -3 ]<br />

Flanagan and Livingston (1995)<br />

Table 7 Equations for calculating transport and detachment capacity.<br />

No.<br />

Equation and Variables<br />

TC<br />

f<br />

TC =<br />

29<br />

TC<br />

( 1<br />

ρ<br />

S<br />

f<br />

)<br />

TC = transport capacity in clear water [g L -1 ], TC f = transport capacity in sedimentwater<br />

[g L -1 ], ρ S = density of solid material [kg m -3 ]<br />

Govers (1990)<br />

qb<br />

w<br />

TC =<br />

30 Q<br />

ρ<br />

S<br />

TC = transport capacity in clear water [g L -1 ], q b = volumetric bedload transport per<br />

unit width [m² s -1 ], w = flow width [m], ρ S = density of solid material [kg m -3 ], Q =<br />

runoff [m³ s -1 ]<br />

Abrahams et al. (2001), Low (1989), Rickenmann (1990)<br />

163


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

qS<br />

w<br />

TC =<br />

31 Q<br />

32<br />

33<br />

34<br />

35<br />

TC = transport capacity in clear water [g L -1 ], ρ S = density of solid material [kg m -3 ], w<br />

= flow width [m], Q = runoff [m³ s -1 ]<br />

Yalin (1993), Bagnold (1980)<br />

ρ<br />

f<br />

TC =<br />

( 1<br />

CP<br />

1E6<br />

CP<br />

)<br />

1E6<br />

TC = transport capacity in clear water [g L -1 ], ρ f = fluid density [kg m -3 ], C P =<br />

concentration [ppm]<br />

Yang (1973)<br />

1.5<br />

TC = k t<br />

τ<br />

f<br />

TC = transport capacity [kg s -1 ], k t = transport coefficient [m 0.5 s² kg -0.5 ], τ f = hydraulic<br />

shear acting on the soil [Pa]<br />

Foster et al. (1995)<br />

11.2 (τ τ<br />

QB<br />

=<br />

3<br />

(τ )<br />

c<br />

c<br />

)<br />

4.5<br />

Q B = transport rate per unit of width [kg 1.5 m -1.5 s -3 ], τ c = threshold value of τ required to<br />

initiate particle motion, τ = shear stress [Pa]<br />

Parker (1979)<br />

D<br />

C<br />

= K<br />

r<br />

(τ<br />

τ<br />

cr<br />

)<br />

D C = detachment capacity [kg m -2 s -1 ], K r = rill erodibility factor [s m -1 ], τ = shear stress<br />

[Pa], τ cr = critical shear stress [Pa]<br />

Foster et al. (1995)<br />

Constant soil parameters?<br />

Another important fact which cannot be handled by the process based models is the<br />

heterogeneity in critical shear stress. In this study, we performed our experiments on one field<br />

with uniform treatment, hence the soil parameters are as constant as possible. As a<br />

consequence, we assumed the critical shear stress to be constant in our experiments. In a<br />

typical model, the critical shear stress for cropland with a sand content > 30% (at our test side:<br />

about 80%, mainly middle, fine and very fine sand) is calculated using the clay and the very<br />

fine sand content and these values are constant. This assumes a constant critical shear stress<br />

but in reality, this homogeneity is not found. The critical shear stress can also change between<br />

164


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

experiments or within one experiment caused by wetting and drying or by sealing and<br />

crusting. These changes are not represented in models.<br />

Better ideas?<br />

The results reported here do not show a simple linear relation between a single hydraulic<br />

parameter and soil detachment rate. Depending on the model purpose and scale, these factors<br />

can be used to predict the magnitude of rill detachment but they are not applicable for the<br />

simulation of rill erosion with high-resolution spatial and temporal change in processes. A<br />

promising approach is to use probability density functions to predict soil detachment. The<br />

newest, although not yet operational, version of this approach has been developed by<br />

Sidorchuk (Sidorchuk, 2002, Sidorchuk et al., 2004, Sidorchuk, 2005 a, Sidorchuk, 2005 b,<br />

Sidorchuk et al., 2008, Sidorchuk, 2009). But two decades ago Nearing et al. (1991)<br />

demonstrated mathematically the stochastic nature of the „critical hydraulic shear stress“ and<br />

recognize that shear stress as well as local soil resistance have to be thought of and<br />

characterized in terms of probability density functions instead of singular values. A stochastic<br />

approach is not itself new: The idea of process-based stochastic modelling of erosion dates<br />

back to the fundamental work of H.A. Einstein (1937). According to Sidorchuk (2005b), the<br />

main ideas which made the stochastic approach attainable for the conditions of shallow,<br />

rough-bed flows and cohesive soils have been formulated by Mirtskhoulava (1988, cited in<br />

Sidorchuk, 2005b). Other approaches to stochastic soil erosion modelling are those published<br />

by Wilson (1993a,b) and Nearing (1991), who presented a simple probabilistically based<br />

mathematical model for the detachment of cohesive soil particles by shallow turbulent flow.<br />

Govindaraju & Kavvas (1994) also developed and tested a spectral stochastic theory for<br />

analysing rill distribution and profiles on a hillslope. Lisle et al. (1998) derive a stochastic<br />

model of sediment particle motion which, given suitable averaging, is consistent with the<br />

ideas of Einstein (1937) and with the model proposed by Hairsine and Rose (1991). The<br />

conceptualization of this model has been corroborated experimentally by Shaw et al. (2008).<br />

On the basis of the aforementioned conceptual work, Sidorchuk (2005b) developed a<br />

stochastic model of cohesive soil erosion and deposition which is denoted hereafter a 3 rd<br />

generation soil erosion model. It is noticeable that the 3 rd generation models are not obtained<br />

by merely imposing probability distributions on the input variables of 2 nd generation models<br />

but constitute a new, physical approach. The RillGrow 1 and 2 models use the stochastic<br />

approach (Favis-Mortlock et al., 1998, 2000). That approach is capable of explicitly<br />

specifying the spatial location of the initiation and subsequent development of rills.<br />

165


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Following the authors, RillGrow 1 has three main limitations: (1) many process descriptions<br />

(e.g. infiltration and deposition) are omitted, (2) the approach is very demanding of<br />

computational resources and (3) the model demands a great deal of data regarding<br />

microtopography (Favis-Mortlock et al., 1998). RillGrow 2 circumvents several but not (yet)<br />

all shortcomings (Favis-Mortlock et al., 2000). Following a personal communication from<br />

Favis-Mortlock (2012), RillGrow is now at version 6, and it is much more process-focused<br />

than earlier versions. The stochastic detachment relationship of Nearing (1991) is now used<br />

for flow detachment. Favis-Mortlock is currently writing up this version.<br />

5 Conclusions<br />

The results of this study clearly showed that the model concept of most process based soil<br />

erosion models is not suitable for modelling rill erosion processes. These models assume a<br />

linear relation between shear stress and soil detachment. That means, when input parameters<br />

for calculating shear stress are constant, or at least in a limited range, the erosion parameters<br />

should also be in a similar, limited range. In our experiments, hydraulic radius, flow velocity,<br />

slope and liquid density showed similar values but the resulting erosion parameters of<br />

sediment concentration, and detachment and transport rates showed variability values of more<br />

than 60%. The total measured rill erosion rates are the sum of erosion rates caused by a<br />

combination of different soil erosion processes with different spatial and temporal<br />

distribution. This combination cannot be described by one single equation.<br />

There are two different ways to address this conundrum. The first is to modify existing<br />

physical based model concepts so that different processes can be taken into account; the<br />

second is to head in a new direction, using the concept of stochastic soil erosion modelling. In<br />

both cases, much more field experimentation will be needed to provide the required data. A<br />

big question is whether or not the experimental setups currently employed can deliver the<br />

required data, or if totally new experimental setups will need to be developed.<br />

Following Rees (2002), there are three great frontiers in science: the very big, the very small,<br />

and the very complex. And erosion, which depends on turbulent behavior, is an example of<br />

the very complex.<br />

Acknowledgements<br />

This research was supported by the Internationales Graduiertenzentrum of Trier University.<br />

We thank all participants of the field trip to the Bardenas Reales in spring 2009 which<br />

supported the accomplishment of the experiments. A special thank goes (1) to our native<br />

166


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

speakers, which revised the complete paper and (2) to the reviewers whose comments helped<br />

to improve significantly the quality of the paper.<br />

References<br />

Abrahams, A.D., Li, G., 1996. Rill hydraulics on a semiarid hillslope, southern Arizona. Earth<br />

Surface Processes and Landforms, 21: 35-47.<br />

Abrahams, A.D., Gao, P., Aebly, F.A., 2000. Relation of sediment transport capacity to stone<br />

cover and size in rain-impacted interrill flow. Earth Surf Proc Land, 25: 497-504.<br />

Ad-hoc-Arbeitsgruppe Boden, 2005. Bodenkundliche Kartieranleitung. 5. Auflage,<br />

Bundesamt für <strong>Geowissenschaften</strong> und Rohstoff in Zusammenarbeit mit den staatlichen<br />

Geologischen Diensten der Bundesrepublik Deutschland.<br />

Ajayi, A.E., Horta, I.D.M.F., 2007. The effect of spatial variability of soil hydraulic properties<br />

on surface runoff processes. Anais XIII Simpósio Brasileiro de Sensoriamento Remoto,<br />

Florianópolis, Brasil, 21-26 April 2007, pp. 3243-3248.<br />

Ali, M., 2011. Sediment transport capacity for soil erosion modelling at hillslope scale: an<br />

experimental approach. Dissertation, Wageningen University<br />

Alonso, C.V., Bennett, S.J., Stein, O.R., 2002. Predicting headcut erosion and migration in<br />

upland flows. Water Resour Res, 38 (12): 1303-1317.<br />

Andrews, E.D., 1983. Entrainment of gravel from naturally sorted riverbed material. Geol Soc<br />

Am Bull, 94: 1225-1231.<br />

Andrews, E.D., 1984. Bed-material entrainment and hydraulic geometry of gravel-bed rivers<br />

in Colorado. Geol Soc Am Bull, 95(3): 371-378.<br />

Andrews, E.D., Erman, D.C., 1986. Persistence in the size distribution of superficial bed<br />

material during an extreme snowmelt flood. Water Resour Res, 22: 191-197.<br />

Ashworth, .PJ., Ferguson, R.I., 1989a. Size-selective entrainment of bed load in gravel bed<br />

streams. Water Resour Res, 25: 627-634.<br />

Ashworth, P.J., Ferguson, R.I., 1989b. Quantifying gravel deposition on river bars using<br />

flexible netting. J Sediment Res, 59(4): 623-624.<br />

Assouline, S., 2009. Drop size distributions and kinetic energy rates in variable intensity<br />

rainfall. Water Resour.Res, 45: W11501, doi: 10.1029/2009WR007927.<br />

Bagnold, R.A., 1963. Mechanics of marine sedimentation. In: M.N. Hill (Editor), The Sea:<br />

Ideas and Observations on Progress in the Study of the Seas, Vol. 3 – The Earth Beneath<br />

the Sea. Wiley-Interscience, London, pp. 507-528.<br />

167


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Bagnold, R.A., 1966. An approach to the sediment transport problem from general physics.<br />

Geological Survey Professional paper 422-1, 37 pp.<br />

Bagnold, R.A., 1977. Bed load transport by natural rivers. Water Resour Res, 13: 303-312,<br />

Bagnold, R.A., 1980. An empirical correlation of bedload transport rates in flumes and natural<br />

rivers. Proc. Royal Society Series, A372: 453-473.<br />

Bell, J.B., Garcia, A.L., Williams, S.S., 2007. Numerical methods for the stochastic Landau-<br />

Lifshitz Navier-Stokes equations. Phys Rev E 76.<br />

Bennett, S.J., 1999. An Experimental Study of Headcut Growth and development in upland<br />

concentrated flows. USDA and Agricultural Research service Research report No. 13,<br />

190 pp.<br />

Bennett, S.J., Alonso, C.V., 2005. Kinematics of flow within headcut scour holes on<br />

hillslopes. Water Resour Res 41:W09418, doi:10.1029/2004WR003752.<br />

Bennett, S.J., Alonso, C.V., Prasad, S.N., Rönkens, M.J.M., 2000. Experiments on headcut<br />

growth and migration in concentrated flows typical of upland areas. Water Resour Res,<br />

36: 1911-1922.<br />

Bruno, C., Di <strong>Stefan</strong>o, C., Ferro, V., 2008. Field investigation on rilling in the experimental<br />

Sparacia area, South Italy. Earth Surf Proc Land, 33: 263-279.<br />

Brunton, D.A., Bryan, R.B., 2000. Rill network development and sediment budgets. Earth<br />

Surf Proc Land, 25: 783-800.<br />

Bryan, R.B., 1990. Knickpoint evolution in rillwash. Catena, 17: 111-132.<br />

Bryan, R.B., Poesen, J., 1989. Laboratory experiment on the influence of slope length on<br />

runoff, percolation and rill development. Earth Surf Proc Land, 14: 211-231.<br />

Bryan, R.B., Govers, G., Poesen, J., 1989. The concept of soil erodibility and some problems<br />

of assessment and application. Catena, 16 (4-5): 393-412.<br />

Bryan, R.B., Hawke, R.M., Rockwell, D.L., 1998. The influence of subsurface moistureon rill<br />

system evolution. Earth Surf Proc Land, 23: 773-789.<br />

Casali, J., Loizu, J., Campo, M.A., De Santisteban, L.M., Álvarez-Mozos, J., 2006. Accuracy<br />

of methods for field assessment of rill and ephemeral gully erosion. Catena, 67: 128-138.<br />

Causapé, J., Quílet, D., Aragüés, R., 2004. Salt and nitrate concentrations in the surface<br />

waters of the CR-V irrigation district (Bardenas I, Spain): Diagnosis and prescriptions for<br />

reducing off-site contamination. J Hydrol, 295: 87-100.<br />

Causapé, J., Quílet, D., Aragüés, R., 2006. Groundwater quality in CR-V irrigation district<br />

(Bardenas I, Spain): Alternative scenarios to reduce off-site salt and nitrate<br />

contamination. Agr Water Manage, 84: 281-289.<br />

168


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Chisci, G., Sfalanga, M., Torri, D., 1985. An experimental model for evaluating soil erosion<br />

on a single-rainstorm basis. In: S.A. El Swaify, W.C. Moldenhauer, A. Lo (Editors), Soil<br />

Erosion and Conservation. Soil conservation Society of America, Ankeny, Iowa, pp. 558-<br />

565.<br />

Constantin, P., 2001. Some open problems and research directions in the mathematical study<br />

of fluid dynamics. Mathematics unlimited-2001 and beyond. pp 353-360.<br />

De Ploey, J., 1990. Threshold conditions for thalweg gullying with special reference to loess<br />

areas. Catena Supplement, 17: 147-151.<br />

De Roo, A.P.J., Wesseling, C.G., Jetten, V.G., Ritsema, C.J., 1996. LISEM – a physically<br />

based hydrological and erosion model incorporated in a GIS. IAHS Publ. 235: 395-403.<br />

De Santisteban, L.M., Casalì, J., Lòpez, J.J., Giràldez, J.V., Poesen, J., Nachtergaele, J., 2005.<br />

Exploring the role of topography in small channelerosion. Earth Surf Proc Land, 30: 591-<br />

599.<br />

Desir, G., Marín, C., 2007. Factors controlling the erosion rates in a semi-arid zone (Bardenas<br />

Reales, NE-Spain). Catena, 71: 31-40.<br />

DIN 1319-1., 1995. Grundlagen der Messtechnik – Teil 1: Grundbegriffe. Deutsches Institut<br />

für Normung e.V., Ausgabe: 1995-01 , Deutsch.<br />

Diplas, P., 1987. Bedload transport in gravel-bed streams. J Hydraul Eng ASCE, 113: 277-<br />

292.<br />

Dunkerley, D., 2008. Rain event properties in nature and in rainfall simulation experiments: a<br />

comparative review with recommendations for increasingly systematic study and<br />

reporting. Hydrol Process, 22: 4415-4435.<br />

Einstein, H.A., 1937. Der Geschiebetrieb als Wahrscheinlichkeitsproblem, Dissertation, ETH<br />

Zürich.<br />

Elliot, W.J., Laflen, J.M., 1993. A Process-based rill erosion model. T. ASAE, 36(1): 35-72.<br />

Elliot, W.J., Liebenow, A.M., Laflen, J.M., Kohl, K.D., 1989. A compendium of soil<br />

erodibility data from WEPP cropland soil field erodibility experiments 1987 and 1988.<br />

NSERL Report No. 3, The Ohio State University, and USDA Agricultural Research<br />

Service, pp. 186.<br />

EPA-homepage: Channel Processes: Bedload transport. United States Environmental<br />

protection agency http://water.epa.gov/scitech/datait/tools/warsss/bedload.cfm. Accessed<br />

14 September 2010.<br />

Fan, J., Gao, H., Guo, B., 2010. Regularity criteria for the Navier-Stokes-Landau-Lifshitz<br />

system. J Math Anal Appl, 363: 29-37.<br />

169


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Favis-Mortlock, D., Guerra, T., Boardman, J., 1998. A self-organizing dynamic systems<br />

approach to hillslope rill initiation and growth: model development and validation. IAHS<br />

Publ., 249: 53-61<br />

Favis-Mortlock, D.T., Boardman, J., Parsons, A.J., Lascelles, B., 2000. Emergence and<br />

erosion: a model for rill initiation and development. Hydrol. Process., 14: 2173-2205.<br />

Favis-Mortlock, D., Boardman, J., MacMillan, V., 2001. The limits of erosion modelling:<br />

why we should proceed with care. In: R.S. Harmon, W.W. Doe (Editors), Landscape<br />

erosion and evolution modeling. Kluuwer Academic/Plenum, New York, pp. 477-516.<br />

Favis-Mortlock, D., de Boer, D., 2003. Simple at heart? Landscapes as a self-organizing<br />

complex system. In: S. Trudgill, A. Roy (Editors), Contemporary meanings in Physical<br />

Geography: From what to why? Arnold Publishers, London.<br />

Fefferman, C.L., 2006. Existance and smoothness of the Navier-Stokes equation. The<br />

millennium prize problems. Clay Math. Inst., Cambridge, MA, pp. 57-67.<br />

Flanagan, D.C., Livingston, S.J., 1995. WEPP user summary, USDA-water erosion prediction<br />

project. NSERL Report No. 11. National soil erosion research Laboratory, pp. 141.<br />

Flanagan, D.C., Laflen, J.M., Meyer, L.D., 2003. Honoring the Universal Soil Loss Equation.<br />

American Society of Agricultural Engineers, Historic Landmark Dedication.<br />

Flores-Cervantes, J.H., Istanbulluoglu, E., Bras, R.L., 2006. Development of gullies on the<br />

landscape: A model of headcut retreat resulting from plunge pool erosion. J Geophys<br />

Res, 111:F01010, doi:10.1029/2004JF000226.<br />

Foster, G.R., 1982. Modelling the erosion process. In: C.T. Haan, H.P. Johnson, D.L.<br />

Brakensiek (Editors), Hydraulic Modelling of Small Watersheds. ASAE monograph 5,<br />

Hann CT, St. Joseph, MI, pp. 296-379.<br />

Foster, G.R., Huggins, L.F., Meyer, L.D., 1984. A laboratory study of rill hydraulics: I.<br />

Velocity relationships. T. ASAE, 27: 790-796.<br />

Foster, G.R., Flanagan, D.C., Nearing, M.A., Lane, L.J., Risse, L.M., Finkner, S.C., 1995.<br />

Hillslope erosion component. In: D.C. Flanagan, M.A. Nearing (Editors), USDA Water<br />

Erosion Prediction Project: Hillslope Profile and Watershed Model Documentation,<br />

Chapter 11, USDA-ARS NSERL Report 10.<br />

Ghebreiyessus, Y.T., Gantzer, C.J., Alberts, E.E., Lentz, R.W., 1994. Soil erosion by<br />

concentrated flow: shear stress and bulk density. T. ASAE, 37 (6): 1791-1797.<br />

Gilley, J.E., Kottwitz, E.R., Simanton, J.R., 1990. Hydraulic characteristics of rills. T. ASAE,<br />

33: 1900-1906.<br />

170


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Giménez, R., Govers, G., 2002. Flow Detachment by Concentrated Flow on Smooth and<br />

Irregular Beds. Soil Sci Soc Am J, 66(5): 1475-1483.<br />

Giménez, R., Planchon, O., Silvera, N., Govers, G., 2004. Longitudinal velocity patterns and<br />

bed morphology interaction in a rill. Earth Surf Proc Land, 29: 105-114.<br />

Gordon, L.M., Bennett, S.J., Wells, R.R., Alonso, C.V., 2007. Ephemeral gully headcut<br />

development and migration in stratified soil. In: J. Casalí, R. Giménez (Editors). Progress<br />

in Gully Erosion Research, IV International Symposium on Gully erosion, Pamplona,<br />

Spain, 17-19 September 2007, pp. 52-53.<br />

Govers, G., 1987. Spatial and temporal variability in rill development processes at the<br />

Huldenberg experimental site. Catena Supplement, 8: 17-34.<br />

Govers, G., 1990. Empirical relationships for transport capacity of overland flow. IAHS<br />

publication, 189: 45-63.<br />

Govers, G., 1991. Rill erosion on arable land in central Belgium: Rates, controls and<br />

predictability. Catena, 18: 133-155.<br />

Govers, G., 1992a. Evaluation of transport capacity formulae for overland flow. In: A.J.<br />

Parsons, A.D. Abrahams (Editors), Overland flow: hydraulics and erosion mechanics,<br />

UCL Press, London, pp. 243-273.<br />

Govers, G., 1992b. Relationship between discharge, velocity and flow area for rills eroding<br />

loose, non layered materials. Earth Surf Proc Land, 17: 515-528.<br />

Govers, G., Poesen, J., 1988. Assessment of the interrill and rill contributions to total soil loss<br />

from an upland field plot. Geomorphology, 1: 343-354.<br />

Govers, G., Giménez, R., van Oost, K., 2007. Exploring the relationship between<br />

experiments, modelling and field observations. Earth Sci Rev, 84 (3-4): 87-102.<br />

Govindaraju, R.S., Kavvas, M.L., 1994. A spectral approach for analysing the rill structure<br />

over hillslopes. Part I. Development of stochastic theory. J Hydrol, 158(3-4): 333-347.<br />

Graf, W.H., 1971. Hydraulics of sediment transport. McGraw-Hill, New York.<br />

Hairsine, P.B., Rose, C.W., 1991. Rainfall detachment and deposition: Sediment transport in<br />

the absence of flow-driven processes. Soil Sci Soc Am J, 55 (2): 320-324.<br />

Hairsine, P.B., Rose, C.W., 1992. Modeling water erosion due to overland flow using<br />

physical principles, 2. Rill flow. Water Resour Res, 28: 245-250.<br />

Hagmann, J., 1996. Mechanical soil conservation with contour ridges: cure for, or cause of,<br />

rill erosion? Land Degrad Dev, 7: 145-160.<br />

Helming, K., Römkens, M.J.M., Prasad, S.N., Somme, H., 1999. Erosional development of<br />

small scale drainage networks. In: S. Hergarten, H.J. Neugebauer (Editors), Process<br />

171


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Modelling and Landform Evolution, Lecture Notes in Earth Science, 78, Springer,<br />

Berlin, pp. 123-146.<br />

Hessel, R., Jetten, V., 2007. Suitability of transport equations in modelling soil erosion for a<br />

small Loess plateau catchment. Eng Geol, 91: 56-71.<br />

Huang, C., Bradford, J.M., Laflen, J.M., 1996. Evaluation of the detachment-transport<br />

coupling concept in the WEPP rill erosion equation. Soil Sci Soc Am J, 60: 734-739.<br />

Kinnell, P.I.A., 2005. Raindrop-impact-induced erosion processes and prediction: a review.<br />

Hydrol. Process., 19: 2815-2844.<br />

Kleinhans, M.G., Bierkens, M.F.P., van der Perk, M., 2010. HESS opinions. On the use of<br />

laboratory experimentation: „Hydrologists, bring out shovel and garden hoses and hit the<br />

dirt“. Hydrol Earth Syst Sc, 14: 369-382.<br />

Knapen, A., Poesen, J., Govers, G., Gyssels, G., Nachtergaele, J., 2007. Resistance of soils to<br />

concentrated flow erosion: A review. Earth Sci Rev, 80(1-2): 75-109.<br />

Kohl, K.D., 1988. Mechanics of rill headcutting. Dissertation, Iowa State University, Ames.<br />

Komer, P.D., 1987a. Selective grain entrainment by a current from a bed of mixed sizes: A<br />

reanalysis. J Sediment Petrol, 57: 203-211.<br />

Komar, P.D., 1987b. Selective gravel entrainment and the empirical evolution of flow<br />

competence. Sedimentology, 34: 1165-1176.<br />

Komar, P.D., Carling, P.A., 1991. Grain sorting in gravel-bed streams and the choice of<br />

particle sizes for flow-competence evaluations. Sedimentology, 38: 489-502.<br />

Kovacs, A., Parker, G., 1994. A new vectorial bedload formulation and its application to the<br />

time evolution of straight river channels. J Fluid Mec, 267: 153-183.<br />

Laflen, J.J., Elliott, W.J., Flanagan, D.C., Meyer, C.R., Nearing, M.A., 1997. WEPPpredicting<br />

water erosion using a process-based model. Journal of Soil and Water<br />

Conservation, 52: 96-102.<br />

Landau, L., Lifchitz, E., 1971. Mécaniques des fluids. MIR, Moscow.<br />

Lascelles, B., Favis-Mortlock, D.T., Parsons, A.J., Guerra, A.J.T., 2000. Spatial and temporal<br />

variation in two rainfall simulators: Implications for spatially explicit rainfall simulation<br />

experiments. Earth Surf. Process. Landforms, 25: 709-721.<br />

Li, G., Abrahams, A.D., Atkinson, J.F., 1996. Correction factors in the determination of mean<br />

velocity of overland flow. Earth Surface processes and Landforms, 21: 509-515.<br />

Lisle, I.G., Rose, C.W., Hogarth, W.L., Hairsine, P.B., Sander, G.C., Parlange, J.-Y., 1998.<br />

Stochastic sediment transport in soil erosion. J Hydrol, 204 (1-4): 217-230.<br />

172


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Liu, B.Y., Nearing, M.A., Baffaut, C., Ascough, II J.C., 1996. The WEPP watershed model:<br />

III. Comparisons to measured data from small watersheds. T. ASAE, 40: 945-951.<br />

Low, H.S., 1989. Effect of sediment density on bed-load transport. J. Hydraul. Eng. ASAE,<br />

115: 124-138.<br />

Lyle, W.M., Smerdon, E.T., 1965. Relation of compaction and other soil properties to erosion<br />

resistance of soils. T. ASAE, 8: 419-422.<br />

Mancilla, G.A., Chen, S., McCool, D.K., 2005. Rill density prediction and flow velocity<br />

distribution on agricultural areas in the Pacific Northwest. Soil Till Res, 84: 54-66.<br />

Miller, M., McCave, I.N., Komar, P.D., 1977. Threshold sediment motion under<br />

unidirectional currents. Sedimentology, 24: 507-527.<br />

Mirtskhoulava, T.Y., 1988. Principles of Physics and Mechanics of Channel Erosion.<br />

Gidrometeoizdat: Leningrad (in Russian).<br />

Moncoqut, F., 2003. Le desert de Bardenas Reales et sa region. Editions Lavielle.<br />

Moore, I.D., Burch, G.J., 1986. Sediment transport capacity of sheet and rill flow: Application<br />

to unit stream power theory. Water Resour Res, 22: 1350-1360.<br />

Morgan, R.P.C., Quinton, J.N., Smith, R.E., Govers, G., Poesen, J.W.A., Auerswald, K.,<br />

Chisci, G., Torri, D., Styczen, M.E., 1998. The European soil erosion model<br />

(EUROSEM): a dynamic approach for predicting sediment transport from field and small<br />

catchments. Earth Surface Processes and Landforms, 23: 527-544.<br />

Murelage, X., Suberbiola, X.P., De Lapparent de Broin, F., Rage, J.-C., Duffaud, S., Astibia,<br />

H., Badiola, A., 2002. Amphibians and Reptiles from the Early Miocene of the Bardenas<br />

Reales of Navarre (Ebro Basin, Iberian Peninsula). Geobis, 35: 347-365.<br />

Nearing, M.A., 1991. A probabilistic model of soil detachment by shallow turbulent flow. T.<br />

ASAE, 34 (1): 81-85.<br />

Nearing, M.A, 1998. Why soil erosion models over-predict small soil losses and under-predict<br />

large soil losses. Catena, 32: 15-22.<br />

Nearing, M.A., Parker, S.C., 1994. Detachment of soil by flowing water under turbulent and<br />

laminar conditions. Soil Sci Soc Am J, 58 (6): 1612-1614.<br />

Nearing, M.A., Foster, G.R., Lane, L.J., Finkner, S.C., 1989. A process-based soil erosion<br />

model for USDA - Water Erosion Predict Project technology. T. ASAE, 32: 1587-1593.<br />

Nearing, M.A., Bradford, J.M., Parker, S.C., 1991. Soil detachment by shallow flow at low<br />

slopes. Soil Sci Soc Am J, 55(2): 339-344.<br />

Nearing, M.A., Norton, L.D., Bulgakov, D.A., Larionov, G.A., West, L.T., Dontsova, K.M.,<br />

1997. Hydraulics and Erosion in Eroding Rills. Water Resour Res, 33 (4): 865-876.<br />

173


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Nearing, M.A., Govers, G., Norton, L.D., 1999a. Variability in soil erosion data from<br />

replicated plots. Soil Sci Soc Am J, 63: 1829-1835.<br />

Nearing, M.A., Simanton, J.R., Norton, L.D., Bulygin, S.J., Stone, J., 1999b. Soil erosion by<br />

surface water flow on a stony, semiarid hillslope. Earth Surf. Process. Landforms, 24:<br />

677-686.<br />

Ott, W.P., van Uchelen, J.C., 1936. Application of similarity principles and turbulence<br />

research to bed.load movement. Hydrodynamics Laboratory California Institute of<br />

Technology Publication No. 167, Soil conservation Service.<br />

Owoputi, L.O., Stolte, W.J., 1995, Soil detachment in the physically based soil erosion<br />

process: a review. T. ASAE, 38 (4): 1099-1110.<br />

Paola, C., Straub, K., Mohrig, D., Reinhardt, L., 2009. The "unreasonable effectiveness" of<br />

stratigraphic and geomorphic experiments. Earth Sci Rev, 97(1-4): 1-43.<br />

Parker, G., 1979. Hydraulic geometry of active gravel rivers. J. Hydr. Eng. Div. ASCE, 105:<br />

1185-1201.<br />

Parker, G., 1983. Discussion of “lateral bed load transport on side slopes” by S. Ikeda. J<br />

Hydraul Eng ASCE, 109: 197-199.<br />

Parker, G., 1990. Surface-based bedload transport relation for gravel rivers. J Hydraul Res,<br />

28: 417–436.<br />

Parker, G., Klingeman, P.C., McLean, D.G., 1982. Bedload and size distribution in paved<br />

gravel-bed streams. J. Hydr. Eng. Div. ASCE, 108: 544-571.<br />

Parsosn, A.J., Wainwright, J., Stone, P.M., Abrahams, A.D., 1999. Transmission losses in rills<br />

on dryland hillslopes. Hydrol. Process., 13: 2897-2905.<br />

Parsons, A.J., Wainwright, J., 2006. Depth distribution of interrill overland flow and the<br />

formation of rills. Hydrol. Process., 20: 1511-1523.<br />

Parsons, A.J., Wainwright, J., Brazier, R.E., Powell, D.M., 2006. Is sediment delivery a<br />

fallacy? Earth Surf. Process. Landforms, 31: 1325-1328.<br />

Partheniades, E., Paaswell, R.E., 1970. Erodibility of channels with cohesive boundary. J<br />

Hydr Eng Div ASCE: 755-771.<br />

Rapp, I., 1998. Effects of soil properties and experimental conditions on the rill erodibilities<br />

of selected soils. Dissertation, Faculty of Biological and Agricultural Sciences,<br />

University of Pretoria, South Africa.<br />

Rees, M., 2002. Our cosmic habitat. Princeton University Press, New Jersey.<br />

Regüés, D., Balasch, J.C., Castelltort, X., Soler, M., Gallart, F., 2000. Relación entre las<br />

tendencias temporales de producción y transporte de sedimentos y las condiciones<br />

174


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

climáticas en una pequeña cuenca de montaña mediterránea (Vallcebre, Pirineos<br />

Orientales). Cuadernos de Investigación Geográfica, 26: 41–65.<br />

Reid, D.M., Dunne, T., 1996. Rapid evaluation of sediment budgets. Catena Verlag,<br />

Reiskirchen, Germany.<br />

Rejman, J., Brodowski, R., 2005. Rill characteristics and sediment transport as a function of<br />

slope length during a storm event on loess soil. Earth Surf Proc Land, 30: 231-239.<br />

Rickenmann, D., 1991. Hyperconcentrated flow and sediment transport at steep slopes. J<br />

Hydraul Eng ASCE, 117: 1419-1439.<br />

Risse, L.M., Nearing, M.A., Nicks, A.D., Laflen, J.M., 1993. Assessment of error in the<br />

universal soil loss equation. Soil Sci Soc Am J, 57: 825-833.<br />

Robinson, K.M., Bennett, S.J., Casali, J., Hanson, G.J., 2000. Processes of headcut growth<br />

and migration in rills and gullies. Int J Sediment Res, 15(1): 69-82.<br />

Ruttimann, M., Schaub, D., Prasuhn, V., Ruegg, W., 1995. Measurement of runoff and soil<br />

erosion on regulary cultivated fields in Switzerland – some critical considerations.<br />

Catena, 25: 127-139.<br />

Sancho, C., Benito, G., Munoz, A., Pena, J.L., Longares, L.A., McDonald, E., Rhodes, E.,<br />

Saz, M.A., 2007. Actividad alluvial durante la pequena Edad del Hielo en Bardenas<br />

Reales de Navarra. Geogaceta, 42: 111-114.<br />

Sancho, C., Pena, J.L., Munzo, A., Bonito, G., McDonald, E., Rhodes, E.J., Longares, L.A.,<br />

2008. Holocene alluvial morphopedosedimentary record and environmental changes in<br />

the Bardenas Reales Natural Park (NE Spain). Catena, 73: 225-238.<br />

Scherer, U., 2008. Prozessbasierte Modellierung der Bodenerosion in einer Lösslandschaft.<br />

Dissertation, Fakultät für Bauingenieur-, Geo- und Umweltwissenschaften, Universität<br />

Fridericiana zu Karlsruhe (TH), Karlsruhe, Germany.<br />

Schneider, G., 2008. Ein Millenniumsproblem: Die globale Existenz und Eindeutigkeit glatter<br />

Lösungen der dreidimensionalen Navier-Stokes-Gleichungen. Mathematik-Online:<br />

Beiträge zu berühmten, gelösten und ungelösten Problemen, Nummer 2.<br />

Seeger, M., Errea, M.P., Beguería, S., Arnáez, J., Martí, C., García-Ruiz, J.M., 2004.<br />

Catchment soil moisture and rainfall characteristics as determinant factors for<br />

discharge/suspended sediment hysteretic loops in a small headwater catchment in the<br />

Spanish Pyrenees. Journal of Hydrology, 288: 299–311.<br />

Seiler, R., 2002. Die Navier-Stokes-Gleichung. Elemente der Mathematik, 57: 109-114.<br />

Shaw, S.B., Makhlouf, R., Walter, M.T., Parlange, J.-Y., 2008. Experimental testing of a<br />

stochastic sediment transport model. J Hydrol, 348(3-4): 425-430.<br />

175


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Shields, A., 1936. Anwendung der Ähnlichkeitsmechanik und der Turbulenzforschung auf die<br />

Geschiebebewegung. Dissertation, TU Berlin, Berlin, Germany.<br />

Sidorchuk, A., 2002. Stochastic Modelling of soil erosion and deposition. Proceedings of the<br />

12 th ISCO Conference, Beijing, China, 26-31 May 2002, pp. 136-142.<br />

Sidorchuk, A., 2005a. Stochastic components in the gully erosion modelling. Catena, 63: 299-<br />

317.<br />

Sidorchuk, A., 2005b. Stochastic modelling of erosion and deposition in cohesive soils.<br />

Hydrol Process, 19: 1399-1417.<br />

Sidorchuk, A., 2009. High-Frequency variability of aggregate transport under water erosion of<br />

well-structured soils. Eurasian Soil Sci, 42(5): 543-552.<br />

Sidorchuk, A., Smith, A., Nikora, V., 2004. Probability distribution function approach in<br />

stochastic modelling of soil erosion. In: V. Golosov, V. Belyaev, D.E. Walling (Editors),<br />

Sediment transfer trough the fluvial system, IHAS Publ. 288, pp. 345-353.<br />

Sidorchuk, A., Schmidt, J., Cooper, G., 2008. Variability of shallow overland flow velocity<br />

and soil aggregate transport observed with digital videography. Hydrol Process, 22:<br />

4035-4048.<br />

Slattery, M.C., Bryan, R.B., 1992. Hydraulic conditions for rill incision under simulated<br />

rainfall: a laboratory experiment. Earth Surf Proc Land, 8: 97-105.<br />

<strong>Stefan</strong>ovic, J.R., Bryan, R.B., 2009. Flow energy and channel adjustments in rills developed<br />

in loamy sand and sandy loam soils. Earth Surf Proc Land, 34: 133-144.<br />

Stroosnijder, L., 2005. Measurement of erosion: Is it possible? Catena, 64: 162-173.<br />

Temam, R., 2000. Some developments on Navier-Stokes equations in the second half of the<br />

20 th century. In: J.-P. Pier (Editor), Development of mathematics 1950-2000, Birkhäuser,<br />

Basel, Switzerland, pp. 1049-1106.<br />

Torri, D., Dfalanga, M., Chisci, G., 1987. Threshold conditions for incipient rilling. Catena<br />

Supplement, 8: 97-105.<br />

Wagenbrenner, J.W., Robichaud, P.R., Elliot, W.J., 2010. Rill erosion in natural and disturbed<br />

forests: 2. Modeling Approaches. Water Resour Res, 46:doi:10.1029/2009WR008315.<br />

Wendt, R.C., Alberts, E.E., Hjelmfelt, J.R., 1986. Variability of runoff and soil loss from<br />

fallow experimental plots. Soil Sci Soc Am J, 50: 730-736.<br />

Wiegner, M., 1999. The Navier-Stokes equations – a neverending challenge? Jahresbericht<br />

Deutsch. Math.-Verein, 101 (1): 1-25.<br />

Wilson, B.N., 1993a. Development of a fundamentally-based detachment model. T. ASAE,<br />

36 (4): 1105-1114.<br />

176


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Wilson, B.N., 1993b. Evaluation of a fundamentally-based detachment model. T. ASAE, 36<br />

(4): 1115-1122.<br />

<strong>Wirtz</strong>, S., Seeger, M., Ries, J.B., 2010. The rill experiment as a method to approach a<br />

quantification of rill erosion process activity. Z Geomorphol, 54 (1): 47-64.<br />

<strong>Wirtz</strong>, S., Seeger, M., Ries, J.B., 2012. Field experiments for understanding and quantification<br />

of rill erosion processes. Catena, 91: 21-34.<br />

Wishmeier, W.C., Smith, D.D., 1965. Predicting Rainfall Erosion Losses from Cropland East<br />

of the Rocky Mountains. Agricultural Handbook No. 282. US Department of<br />

Agriculture, Washington DC.<br />

Wishmeier, W.C., Smith, D.D., 1978. Predicting Rainfall Erosion Losses – A Guide to<br />

conservation Planning. Agricultural Handbook No. 537. US Department of Agriculture,<br />

Washington DC.<br />

Yalin, M.S., 1963. An expression for bedload transportation. J Hydr Eng Div ASCE, 89: 221-<br />

250.<br />

Yang, C.T., 1972. Unit stream power and sediment transport. J Hydr Eng Div ASCE, 98:<br />

1805-1825.<br />

Yang, C.T., 1973. Incipient motion and sediment transport. J Hydr Eng Div ASCE, 99: 1679-<br />

1703.<br />

Zhang, X.C., Nearing, M.A., Risse, L.M., McGregor, K.C., 1996. Evaluation of runoff and<br />

soil loss predictions using natural runoff plot data. T. ASAE, 39: 855-863.<br />

Zhang, G., Liu, B., Liu, G., He, X., Nearing, M.A., 2003. Detachment of undisturbed soil by<br />

shallow flow. Soil Sci Soc Am. J, 67: 713-719.<br />

Zhang, G., Luo, R., Cao, Y., Shen, R., Zhang, X.C., 2010. Correction factor to dye-measured<br />

flow velocity under varying water and sediment discharges. Journal of Hydrology, 389:<br />

205-213.<br />

Zhu, J.C., Gantzer, C.J., Peyton, R.L., Alberst, E.E., Anderson, S.H., 1995. Simulated smallchannel<br />

bed scour and head cut erosion rates compared. Soil Sci Soc Am J, 59 (1): 211-<br />

218.<br />

177


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Kapitel 6<br />

<strong>Wirtz</strong> et al. (subm.2012): Applicability of different hydraulic parameters to describe soil<br />

detachment in eroding rills. PLoSOne.<br />

178


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Applicability of different hydraulic parameters to describe soil detachment in eroding<br />

rills<br />

<strong>Wirtz</strong> S. a , Seeger M. a,b , Zell A. c , Wagner C. c , Wagner J.-F. d , Ries J.B. a<br />

a: Dep. of Physical Geography, Trier University, Germany; b: Dep. of Land Degradation and<br />

Development, Wageningen University, The Netherlands; c: Dep. 7.3- Technical Physics,<br />

Saarland University, Germany; d: Dep. of Geology, Trier University, Germany<br />

Abstract<br />

This paper presents the comparison of experimental results with assumptions used in<br />

numerical models. The aim of the field experiments is to test the linear relationship between<br />

different hydraulic parameters and soil detachment. The correlations between shear stress,<br />

unit length shear force, stream power, unit stream power and effective stream power and the<br />

detachment rate show that there is not a single parameter that continuously displays the best<br />

correlation. The best match not only changes from one experiment to another, but also from<br />

one measuring point to another. Different processes in rill erosion are responsible for the<br />

changing correlations. However, not all these procedures are considered in soil erosion<br />

models. Hence, hydraulic parameters alone are not sufficient to predict detachment rates.<br />

They predict the fluvial incising in the rill's bottom, but the main sediment sources are not<br />

considered sufficiently in its equations. The results of this study show that there is still a lack<br />

in understanding of the physical processes underlying soil erosion. Exerted forces, soil<br />

stability and its expression, the abstraction of the detachment and transport processes in<br />

shallow flowing water are still subject of unclear description and dependence.<br />

Keywords<br />

Soil erosion, rill hydraulics, shear stress, rill erosion, dynamic viscosity, field experiments<br />

1 Introduction<br />

Soil erosion models use different composite factors to describe and predict soil detachment<br />

and transport capacity. The factors used most frequently are average shear stress [1-4], unit<br />

length shear force [5], stream power [4, 6-9], unit stream power [10, 11] and effective stream<br />

power [12, 13]. Giménez and Govers [5] define the different hydraulic parameters as follows:<br />

179


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

(1) shear stress is the drag force exerted by the flow on the bed, (2) stream power is defined as<br />

the energy of the flow dissipated to the bed based on total discharge, (3) unit stream power is<br />

a product of slope and flow velocity, (4) the effective stream power can be seen as the power<br />

over a given rill cross section and (5) the unit length shear force is a hydraulic parameter<br />

calculated on a unit length basis [5].<br />

In most cases, a linear equation describes the relation between the hydraulic parameters<br />

mentioned above and the detachment rate. By exceeding a certain threshold, erosion by<br />

concentrated flow begins and detachment rate increases. This threshold has a positive x-axis<br />

intercept, which means, there is no detachment below this point.<br />

Another option is to consider concentrated flow erosion as a nonlinear threshold phenomenon<br />

or as a two-part linear threshold phenomenon: below the threshold soil detachment takes place<br />

(first linear relationship) but after exceeding the threshold, detachment rate increases much<br />

faster (second linear relationship) [14]. But is this linear relationship really suitable?<br />

Knapen et al. [14] calculated the correlation between shear stress, unit length shear force,<br />

stream power and Reynolds number and the detachment rate from several WEPP datasets.<br />

The best average correlation was determined for stream power with R² = 0.59. The WEPPused<br />

shear stress is a variable that reaches only low R² values for all of the tested data sets.<br />

Knapen et al. [14, p. 80f.] describes the shear stress as follows: “Although the use of flow<br />

shear stress as soil detachment predictor can be contested, critical shear stress (τ cr ) and<br />

concentrated flow erodibility KC (...) have been selected as the most universal parameters to<br />

describe soil erosion resistance to concentrated flow.” The correlations between these factors<br />

and the soil detachment rate show very different results. There is not a single parameter that<br />

always shows the best correlation. These considerations lead to 2 main questions:<br />

- Are soil erosion, detachment and transport, directly dependent on water flow<br />

characteristics?<br />

- Are these concepts, as implemented in soil erosion models, suitable to describe rill<br />

erosion?<br />

These questions have been tackled by many research groups that have been searching for the<br />

equation that suits their observations best [1-13, 15-42]. However, taking into consideration<br />

the numerous and variable results, a deeper insight into the rill erosion processes on hillslopes<br />

is needed. To get this insight, different strategies can be applied [43]: (I) Modelling, (II)<br />

laboratory experiments (III) field observations and (IV) field experiments. Each of these<br />

methods shows different advantages and disadvantages.<br />

180


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Due to difficulties to measure certain parameters, models have to be calibrated. During this<br />

process, the phenomenon of equifinality can appear: different parameter sets show the same<br />

result. Another point of criticism in model developing methods is that parameters for rill<br />

hydraulics are often adapted from equations made for describing flow behaviour in rivers.<br />

Govers and his colleagues [13, 44] showed that these equations are not suitable for rill erosion<br />

processes. Therefore, there is often a mismatch between model results and observed or<br />

measured “reality” [43]. Additionally, models only project the concepts of the designer, not<br />

necessarily the reality. Herein, there is a reproducibility of the modelling results, contrasting<br />

with the measurements, which often have a high inconsistency and can be modelled by a<br />

flexible set of input variables.<br />

In laboratory experiments, the initial and boundary conditions are well controlled. Soil<br />

parameters are well known and rill forms and slope can be adapted to the specific question.<br />

Thus, physical laws can be tested/ controlled in a well defined environment. However,<br />

Giménez and Govers [5] showed that parameters determined under laboratory conditions are<br />

not easily transformable to natural environments. One disadvantage of former laboratory<br />

experiments or field observations is that in most cases only total runoff and sediment output<br />

are measured while the relative contribution of the individual processes is not considered [45].<br />

Field data is as close as possible to reality. Nevertheless, observations as well as experiments<br />

show certain disadvantages: (I) Measurement techniques may disturb the observed processes,<br />

(II) time scale of human observations is shorter than that of the process under study, (III)<br />

some processes cannot be measured directly or indirectly and (IV) some processes are chaotic<br />

and the spatial and temporal variations are difficult to specify<br />

[43].<br />

The linear or two-part-linear relationship between soil detachment and hydraulic parameters<br />

used in soil erosion models is in most cases deduced from laboratory experiments. The<br />

question to be answered is: Are these relationships suitable for erosion processes in natural<br />

rills? Thus, the experiments have been accomplished in the field and not in artificial<br />

laboratory rills. The setup used enables to measure the input parameters for calculating<br />

hydraulic parameters. Therefore, it is no need to build on estimations as it is done in the small<br />

number of published field experiments. The advantages of laboratory experiments are<br />

combined with the advantage of testing natural rills. That means in all experiments, the same<br />

experimental setup was used (similar inflow intensity, same water quantity, same sampling<br />

scheme, same measurement of hydraulic parameters) and these experiments have been<br />

accomplished under field conditions (no artificial laboratory rills in bulk substrates but<br />

181


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

naturally developed rills in naturally developed soils). The purpose of the field experiments<br />

was to quantify in a detailed temporal and spatial resolution the soil erosion dynamics in<br />

natural rills under concentrated flow and to compare the measured sediment dynamics with<br />

those calculated by means of the most common detachment and transport equations.<br />

The aims of this paper are<br />

1) to elucidate the relationship between hydraulic parameters such as shear stress, unit<br />

length shear force, unit stream power, stream power, effective stream power and the<br />

Reynolds number and soil detachment in natural rills.<br />

2) An additional aim is to explain the reasons for the obvious problems of the physically<br />

based soil erosion models in reflecting the processes in rills.<br />

3) Finally a basic consideration is employed: Are the currently used model approaches<br />

able to describe rill erosion with all involved processes in general?<br />

The overall aim of this paper is to have a critical view on concepts for modeling rill erosion<br />

based on experiments performed in naturally developed rills.<br />

2 Material and Methods<br />

Ethics Statement:<br />

No specific permits were required for the described field studies. The mayors of the towns<br />

next to the study sites or the owners of the fields were informed about the intended activities<br />

and were asked for permission. The test sites Freila, Negratin and Salada are abandoned fields<br />

which are sporadically used as pasture for goats or sheep and in Belerda the experiment was<br />

accomplished on an almond field. The locations Freila, Negratin end Salada are not privatelyowned<br />

and permission was granted from the owner of the study site Belerda. None of the<br />

study sites are protected in any way and the field studies did not involve endangered or<br />

protected species.<br />

Study areas:<br />

The four study areas in Andalusia are located at Negratin, Freila, Salada and Belerda. UTM<br />

coordinates of the tested rills are given in Table 1.<br />

182


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Table 1 Temperature, precipitation, northing and easting of the four test areas.<br />

Tested Rill<br />

Meteorological<br />

station<br />

Average<br />

annual<br />

temperature<br />

Annual<br />

precipitation<br />

Northing of<br />

the rill<br />

Easting of<br />

the rill<br />

Freila 1+3 Baza 14.2 °C 368 mm 4154368 509860<br />

Freila 2 Baza 14.2 °C 368 mm 4154398 509826<br />

Negratin Baza 14.2 °C 368 mm 4156324 505710<br />

Salada<br />

Embalse<br />

Valdeinfierno<br />

13.4 °C 311 mm 4187266 595761<br />

Belerda Granada 15.6 °C 473 mm 4133440 478070<br />

Negratin and Freila: The areas are located within the Hoya de Baza sedimentary basin and<br />

composed of marls, in which calcareous Regosols have developed. The climate is semi-arid<br />

and vegetation is dominated by low shrubs and Stipa tenacissima grass tussocks. The land<br />

cover at the south side of the Negratin-dam is dominated by abandoned cereal fields, which<br />

are extensively grazed by sheep and agricultural land comprised mainly of cereal dry-farming<br />

and almond grooves [46].<br />

Salada: Located at the SE-margin of the Betic range (SE-Spain), inside the penibetic complex.<br />

The area is composed of conglomerates with a clayey to loamy matrix, in which Regosols as<br />

well as to fairly developed (Calcic) Cambisols have developed. Vegetation is similar to that<br />

found in the Freila and Negratin-area. The climate is semi-arid too, but less accentuated than<br />

in the previously mentioned area [46]. Here a mixed pattern of rainfed agricultural areas,<br />

mainly cereals, olives and almonds, and abandoned or uncultivated regions is to find.<br />

Belerda: This test area is located in the Guadix basin. The parent material consists of tertiary<br />

and quaternary conglomerates, sands, silts and clays. The soil texture class following the FAO<br />

[47] is a silty clay loam. The land use is separated into cultivated areas, with almond and olive<br />

groves, and abandoned agricultural fields [48]. The climate is, though still semi-arid,<br />

characterised by higher average annual temperatures and precipitations in comparison with<br />

the other test zones.<br />

The climatic parameters of the test fields are summarized in table 1.<br />

Tested rills:<br />

The main descriptors of the rills are summarized in table 2; photographies of the rills are<br />

presented in figure 1.<br />

183


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Figure 1 Photographies of the tested rills<br />

Table 2 Rill parameters: Grain size class limits are from [49], texture class is determined<br />

following [47].<br />

Freila 1 Freila 2 Freila 3 Negratin Salada Belerda<br />

Ø Slope<br />

[°]<br />

9.4 7.7 9.4 5.6 25.6 16.9<br />

Max. Slope<br />

[°]<br />

15.2 14.1 15.2 12.9 7.3 12.5<br />

Tested flow<br />

length [m]<br />

16 21 16 30 17 23<br />

Texture class SL SL SL SiL SiCL L<br />

Gravel > 2000 µm [%] 30 30 30 1 1 13<br />

Sand 2000-630 µm [%] 14 14 14 1 2 10<br />

630-200 µm [%] 14 14 14 5 2 10<br />

184


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

200-125 µm [%] 13 13 13 6 1 8<br />

125-63 µm [%] 16 16 16 11 7 17<br />

Silt 63-20 µm [%] 13 13 13 11 17 13<br />

20-6.3 µm [%] 10 10 10 20 17 13<br />

6.3-2 µm [%] 11 11 11 24 24 14<br />

Clay < 2 µm [%] 9 9 9 21 29 15<br />

Starting soil<br />

moisture [% w/w]<br />

3.1 3.5 3.1 3.1 5.8 2.4<br />

K t<br />

[s 2 m 0.5 kg -0.5 ]<br />

Location<br />

WEPP dataset<br />

Maximum<br />

width [m]<br />

Maximum<br />

depth [m]<br />

Vegetation<br />

cover [%]<br />

Rock fragment<br />

cover [%]<br />

Grain density<br />

[g cm -3 ]<br />

Dry bulk density<br />

[g cm -3 ]<br />

Org. material<br />

[%]<br />

Critical shear<br />

stress [Pa]<br />

0.0090 0.0090 0.0090 0.0095 0.0096 0.0093<br />

Academy Academy Academy Frederick Mexico Caribou<br />

~ 0.4 ~ 2.2 ~ 0.4 ~ 0.4 ~ 0.5 ~ 0.3<br />

~ 0.05 ~ 0.7 ~ 0.05 ~ 0.2 ~ 0.25 ~ 0.15<br />

~ 40 ~ 40 ~ 40 ~ 0 ~ 15 ~ 5<br />

~ 80 ~ 80 ~ 80 ~ 5 ~ 20 ~ 50<br />

2.69 2.69 2.69 2.65 2.66 2.61<br />

1.44 1.55 1.44 1.57 1.52 1.68<br />

1.29 1.29 1.29 1.75 2.97 1.34<br />

1.97 2.07 1.97 2.93 3.20 2.77<br />

Land use rangeland rangeland rangeland rangeland rangeland cropland<br />

The tested rills in Freila have developed on a sandy loam with high gravel content. Sand<br />

content is 57 % with a relatively homogeneous contribution between coarse, medium, fine and<br />

very fine sand. The same is true in the silt fraction, the 34 % are homogeneously contributed<br />

in the complete silt fraction between 63 and 2 µm. The rills show all a dense rock fragment<br />

cover and the highest vegetation cover of the four test sides.<br />

In Negratin, the soil material is nearly gravel free, coarse, medium and fine sand also show<br />

low amounts, most of the fine material is in the grain size class < 20 µm. The rock fragment<br />

185


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

cover in the rill is higher than the gravel content of the soil material thus it is possible that<br />

residual rock fragment accumulation has occurred.<br />

In Salada the grain size distribution is similar to Negratin. The highest account of the fine soil<br />

material is in the class < 63 µm. The residual rock fragment accumulation is formed even<br />

more clearly as in Negratin; the vegetation cover is relatively high compared to the other test<br />

sites.<br />

The rill in Salada is the only rill that has developed in an actually used agricultural field. The<br />

soil material is composed by a mixture of all particle size classes from gravel to clay. The<br />

rock fragment cover is high compared to the other test sites and the vegetation cover<br />

comparatively low. This test site shows the highest dry bulk density which can be declared by<br />

the actual agricultural use.<br />

Methods:<br />

Rill experiment (RE):<br />

The rill experiments consist of two runs: first the rill is tested under field conditions (run a); in<br />

a second run (run b), approximately 15 minutes later, the same rill is tested under almost<br />

saturated soil conditions. With a motor driven pump, a constant discharge of 250 or 330 L<br />

min -1 is maintained during 4 respectively 3 minutes, resulting in a total water inflow of 1,000<br />

L. Mobilisation of material at the inflow has been avoided. The flow velocity within the rill is<br />

characterized by the travel time of the waterfront and of two colour tracers (started at 1 and 2<br />

minutes of the experiment), measured for every meter using a chronograph. By means of this<br />

procedure, three velocity curves are recorded and changes in flow dynamics can be detected.<br />

As colour tracers, food colourings (E 124 (red) and E 13 (blue)) are used for reasons of safety.<br />

The rill's slope is characterized by measuring with a spring bow of 1m range and a digital<br />

spirit level. It must be considered that slope measuring provides only average slopes for 1<br />

meter. A step or a knick-point in the rill is not accounted, but its position and height are<br />

recorded.<br />

At three measuring points (MP1-3) always four water samples are taken: the first one as soon<br />

as the waterfront has reached the sampling point, the second 30 seconds later, the third after<br />

1:30 minutes and the fourth 2:30 minutes after the arrival of the water. The (suspended)<br />

sediment concentration SSC is determined by filtration of the samples in laboratory [50].<br />

At each measuring point, rill cross section was measured. With a laser rangefinder, the<br />

distance between sensor and rill bottom was measured in 0.002 m steps. This allows an<br />

accurate calculation of the rills cross section area and an estimation of the rills volume.<br />

186


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Water level was continuously measured by ultrasonic sensors at each measuring point.<br />

Descriptors for soil detachment:<br />

Soil detachment can be described by shear stress τ, unit length shear force Γ, stream power ω,<br />

unit stream power ω U and effective stream power ω eff .<br />

* g * R * S<br />

[Pa] Eq. (1)<br />

* g * A*<br />

S *<br />

[N m -1 ] Eq. (2)<br />

W P<br />

* g * R * S * v * v [W m -2 ] Eq. (3)<br />

U<br />

S * v<br />

[m s -1 ] Eq. (4)<br />

1.5 1.5<br />

( * v)<br />

eff<br />

[W m -1 ] Eq. (5)<br />

2<br />

2<br />

3<br />

3<br />

d d<br />

with ρ = liquid density [kg m -3 ], g the gravitational acceleration (9.81 m s -2 ), R the hydraulic<br />

radius [m], A the flow cross section area [m²], S the effective slope (sin(slope angle)), W P the<br />

wetted perimeter [m], v the flow velocity [m s -1 ] and d the water depth [m]; abbreviations of<br />

the units are Pa = Pascal, N = Newton, W = Watt.<br />

Reynolds number describes the balance between the inertial flow forces represented by the<br />

product in the numerator and the viscous forces as described by the dynamic viscosity in the<br />

denominator. It is a criterion for stability of a flowing medium. When Reynolds number is<br />

small, viscous forces dominate the motion and inertial ones can be ignored whereas at high<br />

Reynolds numbers inertial forces dominate and it is often possible to ignore viscosity [51].<br />

Reynolds Number Re is calculated as follows:<br />

*v * R<br />

Re Eq. (6)<br />

with ρ = liquid density [ kg m -3 ], v = flow velocity [m s -1 ], R = hydraulic radius [m] and η =<br />

dynamic viscosity [Pa s].<br />

Liquid density is calculated using sediment concentration and grain density. The use of<br />

water’s density is not practicable due to sediment concentrations of more than 400 g L -1 .<br />

Grain density was measured by a capillary pycnometer following DIN 18124 [52]. Flow<br />

velocity for each sample is interpolated between three measured velocities (arrival of the<br />

waterfront and arrival of the two colour tracers). Hydraulic radius and wetted cross section<br />

area can be calculated by measuring water level and the rill profile.<br />

The viscosity of the sediment suspensions was measured with a shear rate controlled<br />

rheometer (Haake MARS from Thermo Fisher Scientific, Karlsruhe, Germany) and a cone-<br />

187


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

plate geometry with an angle of 2° and a diameter of 60 mm [53]. The shear rate γ is defined<br />

as:<br />

dv<br />

dy<br />

Eq. (7)<br />

with v = fluid velocity and y = the gap between the cone and base plate. The rheomter<br />

controls the shear rate and measures the shear stress τ, from which the viscosity η is<br />

calculated via<br />

Eq. (8)<br />

The sample volume was always 2.0 ml and the cell was tempered to 20 °C +/- 0.01 °C. Data<br />

points were taken at shear rates between 150 s -1 and 1500 s -1 . The viscosity does not depend<br />

on the shear rate. This is according to theoretical considerations. For a suspension of<br />

monodisperse particles one expects a linear relation [54, 55] for volume concentrations up to<br />

approximately 10%.<br />

Detachment rate D R [kg s -1 m -2 ] is calculated from the measured sediment concentrations and<br />

different hydraulic parameters:<br />

D<br />

R<br />

SSC * v * A<br />

L * W<br />

P<br />

with SSC = sediment concentration [g L -1 = kg m -3 ] and L = flow length [m].<br />

Eq. (9)<br />

For the calculation of the critical shear stress, the equations from the WEPP model [34] is<br />

used. It is to separate between “cropland with sand content > 30%” and “rangeland”.<br />

cr<br />

( cropland ) 2.67 0.065*(% clay ) 0.058*(% very fine sand )<br />

Eq. (10)<br />

cr<br />

( rangeland ) 3.23 0.056*(% sand ) 0.244*(% org.<br />

mat.)<br />

0.9*( dry bulk density )<br />

Eq. (11)<br />

For quantification of the different processes in the rill, the transport rate T R [kg s -1 ] and the<br />

transport capacity T C [kg s -1 ] are to calculate:<br />

T R<br />

SSC * v*<br />

A<br />

Eq. (12)<br />

1.5<br />

T C<br />

R * K t<br />

*<br />

Eq. (13)<br />

Kt [s 2 m 0.5 kg -0.5 ] is a transport coefficient depending on soil substrate. The Kt value of the<br />

WEPP substrate which was most similar to the given test site conditions was used.<br />

3 Results<br />

3.1 Initial data<br />

188


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

The used parameters show a wide range of data. In most cases (12 of 19), the standard<br />

deviation is higher than the mean values, the highest standard deviation – mean - percentage<br />

reaches the transport capacity (224 %), the effective stream power (188.9 %), the sediment<br />

concentration (168.3 %) and detachment and transport rate (both 150 %). The lowest<br />

percentage is calculated for sample density (0.5 %). All initial data are presented in<br />

supporting information tables S1-S18 and the statistical values of the data in table 3.<br />

Table 3 Statistical values of the initial data.<br />

Maximum Minimum Mean Standard Deviation Percentage from Mean<br />

SSC [g L -1 ] 422.30 0.001 52.15 87.78 168.3<br />

D R [kg s -1 m -2 ] 0.96 0.001 0.10 0.15 150.0<br />

T R [lg s -1 ] 2.06 0.001 0.16 0.24 150.0<br />

р [g cm -3 ] 1.26 1.00 1.03 0.005 0.5<br />

Slope [°] 24.50 1.70 9.73 6.90 70.9<br />

T C [kg s -1 ] 3.38 0.001 0.25 0.56 224.0<br />

v [m s -1 ] 2.94 0.04 0.79 0.49 62.0<br />

η [kg s -1 m -1 ] 0.00311 0.00100 0.00126 0.00044 34.9<br />

Water depth [cm] 21.00 0.20 3.99 4.23 106.0<br />

A [cm²] 877.69 0.80 149.21 195.84 131.3<br />

W P [cm] 107.58 4.85 38.21 24.16 63.2<br />

R [cm] 9.65 0.10 2.92 2.12 72.6<br />

τ [Pa] 246.70 0.96 52.38 55.18 105.3<br />

Г [N m -1 ] 172.58 0.10 23.99 35.10 146.3<br />

ω [W m -2 ] 365.28 0.31 41.54 55.91 134.6<br />

ω U [m s -1 ] 0.88 0.001 0.14 0.17 121.4<br />

ω eff [W m -1 ] 37864.55 5.81 3807.14 7192.32 188.9<br />

Re [ ] 86918.88 237.00 19053.94 16226.56 85.2<br />

τ - τ cr [Pa] 244.73 -1.46 49.89 55.11 110.5<br />

3.2 Dynamic viscosity<br />

The measured liquid's dynamic viscosities show a clear correlation to sediment concentration<br />

(see figure 2), and increases with sediment concentration. Visible deviations from the trend<br />

line were noted in samples with low sediment concentrations, which were often rich on<br />

transported organic material. The small branchlets with low weight implicate a low sediment<br />

concentration, but in rheometer measurements, they tilt and a high shear stress is erroneously<br />

measured. The trend line equation has been calculated for samples from different test sites,<br />

the R²-value of 0.92 indicates that this equation can be used for further experiments.<br />

189


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Figure 2 Correlation between sediment concentration and the dynamic viscosity<br />

3.3 Correlations between detachment rate and hydraulic parameters<br />

The correlations between the detachment rate and different hydraulic parameters show the<br />

complete possible range from R² = 0 up to R² = 0.99 (see Table 4). Trend lines are increasing,<br />

decreasing and about constant, it is not possible to find any clear regularity. Only 40 of 252<br />

correlations (about 16 %) show an increasing trend line with an R² value ≥ 0.7. The highest<br />

average R²-value is calculated for the (τ-τ cr ) – detachment rate - relationship if all R² values<br />

are used (0.53), if only the R²- values with increasing trend line are considered in calculation,<br />

the τ – detachment rate relationship shows the highest average R² (0.55). Separating the<br />

experiments into two groups, Freila 1-3 with low sediment concentrations (LSSC) and<br />

Negratin, Salada, Belerda with high sediment concentrations (HSSC), it is noticeable that the<br />

LSSC-experiments clearly show higher R²-values than the HSSC-experiments, whether all R²values<br />

or only the R²-values of correlations with increasing trend line are used. Indicating<br />

that, the influence of hydraulic parameters is higher for low sediment concentrations or in<br />

other words, high sediment concentrations are not caused by hydraulic parameters.<br />

Table 4 R² - values between hydraulic parameter and the detachment rate.<br />

τ Г ω ω U ω eff Re τ - τ cr<br />

Freila 1 run a MP 1 0.29 0 0.12 0 0.02 - 0.02 - 0.00 0 0.02 - 0.29 0<br />

MP 2 0.49 0 0.55 0 0.49 + 0.01 0 0.48 + 0.49 + 0.49 0<br />

MP 3 0.18 + 0.18 + 0.01 + 0.94 - 0.01 - 0.02 - 0.18 +<br />

Freila 1 run b MP 1 0.03 0 0.06 0 0.41 - 0.47 - 0.33 - 0.51 - 0.03 0<br />

MP 2 0.95 0 0.98 0 0.95 + 0.30 0 0.94 + 0.95 + 0.95 0<br />

MP 3 0.16 + 0.15 + 0.00 0 0.15 - 0.02 - 0.00 0 0.16 +<br />

Freila 2 run a MP 1 0.81 + 0.91 + 0.17 + 0.45 - 0.32 - 0.01 + 0.81 +<br />

190


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

MP 2 0.08 - 0.08 - 0.58 - 0.61 - 0.37 - 0.61 - 0.08 0<br />

MP 3 0.8 - 0.79 - 0.97 - 0.78 - 0.94 - 0.98 - 0.80 -<br />

Freila 2 run b MP 1 0.84 - 0.85 - 0.99 - 0.84 - 0.92 - 0.99 - 0.84 -<br />

MP 2 0.99 - 0.99 - 0.58 - 0.53 - 0.19 - 0.62 - 0.99 -<br />

MP 3 0.92 - 0.92 - 0.22 - 0.08 - 0.13 - 0.27 - 0.92 -<br />

Freila 3 run a MP 1 0.97 + 0.98 + 0.14 - 0.69 - 0.43 - 0.36 - 0.97 -<br />

MP 2 0.88 - 0.82 - 0.89 - 0.92 + 0.89 - 0.94 - 0.88 -<br />

MP 3 0.80 + 0.81 + 0.62 - 0.99 - 0.86 - 0.78 - 0.80 +<br />

Freila 3 run b MP 1 0.96 + 0.96 + 0.93 - 0.99 - 0.94 - 0.94 - 0.96 +<br />

MP 2 0.64 - 0.60 - 0.68 + 0.72 + 0.74 + 0.79 + 0.64 -<br />

MP 3 0.93 + 0.93 + 0.26 + 0.03 + 0.09 + 0.19 + 0.93 +<br />

Negratin run a MP 1 0.03 - 0.03 - 0.16 - 0.23 - 0.01 + 0.20 - 0.03 -<br />

MP 2 0.60 + 0.53 + 0.22 + 0.05 + 0.06 + 0.14 + 0.60 +<br />

MP 3 0.31 - 0.09 + 0.19 + 0.96 + 0.15 0 0.23 0 0.05 0<br />

Negratin run b MP 1 0.86 - 0.86 - 0.53 - 0.51 - 0.00 0 0.58 - 0.86 -<br />

MP 2 0.14 + 0.19 + 0.16 + 0.08 + 0.17 0 0.14 + 0.14 +<br />

MP 3 0.31 + 0.27 + 0.39 + 0.24 + 0.99 0 0.79 0 0.31 0<br />

Salada run a MP 1 0.44 + 0.45 + 0.70 + 0.64 + 0.70 + 0.67 + 0.44 +<br />

MP 2 0.78 + 0.81 + 0.22 + 0.00 0 0.02 + 0.12 + 0.78 +<br />

MP 3 0.09 - 0.07 - 0.00 0 0.02 0 0.01 0 0.00 0 0.09 -<br />

Salada run b MP 1 0.23 - 0.18 - 0.19 - 0.27 - 0.17 - 0.20 - 0.23 -<br />

MP 2 0.86 + 0.87 + 0.01 - 0.18 - 0.05 - 0.03 - 0.86 +<br />

MP 3 0.00 - 0.00 - 0.03 0 0.02 0 0.05 0 0.02 0 0.00 0<br />

Belerda run a MP 1 0.35 - 0.35 - 0.25 + 0.28 + 0.26 + 0.08 + 0.34 -<br />

MP 2 0.71 + 0.55 + 0.02 0 0.85 - 0.92 - 0.23 - 0.71 +<br />

MP 3 0.27 + 0.19 + 0.47 + 0.31 + 0.78 + 0.00 0 0.27 +<br />

Belerda run b MP 1 0.12 + 0.07 + 0.05 - 0.46 - 0.59 - 0.37 - 0.12 +<br />

MP 2 0.06 + 0.20 + 0.00 0 0.01 0 0.02 - 0.16 - 0.06 +<br />

MP 3 0.71 + 0.74 + 0.88 + 0.86 + 0.83 + 0.72 + 0.71 +<br />

Mean 1 0.52 0.50 0.37 0.43 0.40 0.39 0.53<br />

Mean 2 0.55 0.52 0.40 0.46 0.45 0.39 0.53<br />

Mean 3 0.65 0.65 0.50 0.53 0.48 0.53 0.65<br />

Mean 4 0.69 0.70 0.43 0.56 0.56 0.49 0.64<br />

Mean 5 0.38 0.36 0.25 0.33 0.32 0.26 0.39<br />

Mean 6 0.44 0.41 0.39 0.43 0.52 0.35 0.45<br />

MP = measuring point, τ = shear stress, Г = unit length shear force, ω = stream power, ω U = unit stream power, ω eff =<br />

effective stream power, Re = Reynolds number, + indicates increasing trend line, - indicates decreasing trend line, 0 indicates<br />

a nearly constant trend line, Mean 1 = all values, Mean 2 = only values with increasing trendline, Mean 3 = all Freila<br />

experiments, Mean 4 = Negratin, Salada, Belerda all values, Mean 5 = Freila only values with increasing trend line, Mean 6 =<br />

Negratin, Salada, Belerda only values with increasing trend lines<br />

3.4 Quantification of different erosion processes<br />

The quantification of the different processes consists of several steps. One further model<br />

assumption is, that the transport rate cannot exceed the transport capacity, sedimentation<br />

191


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

processes would reduce the transport rate to the transport capacity [56]. Transport capacity<br />

calculation uses shear stress, meaning if transport rate is higher than the capacity, proceses<br />

without link to shear stress such as bank failure and headcut retreat provide the material.<br />

Therefore, it is possible to calculate the percentage of shear-stress-independent-providedmaterial.<br />

The transport of loose material is somewhere in between, it is caused by shear stress<br />

but dry bulk density and hence the critical shear stress which has do be exceeded is lower than<br />

on “non loosened” substrate. Figure 3 shows the relationships between the measured transport<br />

rates and the predicted transport capacities. From 144 samples, in 82 cases the transport rate<br />

exceeds the capacity, this is about 57 %. Table 5 presents the differences between transport<br />

rates and transport capacities as well as the percentage of transport rate exceeding the capacity<br />

and hence the percentage of processes which are not controlled by the influence of shear<br />

stress. The percentage of material which is transported by processes independent of shear<br />

stress is 41.5 % in mean. The distribution is uneven: in the three Freila-experiments, the mean<br />

is 24.3 % and in Negratin, Salada and Belerda, the value is 58.7 %. The second group shows<br />

clearly higher sediment concentrations that means, the processes independent of shear stress<br />

provide higher sediment concentrations than the shear stress-based processes.<br />

Figure 3 Transport rate vs. Transport capacity<br />

192


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Table 5 Transport rates vs. Transport capacities.<br />

run-MP-time Freila1 Freila 2 Freila 3 Negratin Salada Belerda<br />

Transport rate T R [kg s -1 ] - Transport capacity T C [kg s -1 ]<br />

a - 1 - 0:00 0.016496 0.208367 0.299736 0.356133 0.090648 0.449332<br />

a - 1 - 0:30 0.009216 -0.000042 -0.002630 0.388589 -0.068663 0.658430<br />

a - 1 - 1:30 0.028119 -0.009571 -0.011526 0.302982 -0.008596 0.347591<br />

a - 1 - 2:30 0.003356 0.008419 -0.018163 0.377529 0.129568 0.340157<br />

a - 2 - 0:00 0.000700 0.016861 -0.122188 0.080985 -0.049255 0.206128<br />

a - 2 - 0:30 0.000594 -0.003886 -1.661911 0.239944 -0.227338 0.075105<br />

a - 2 - 1:30 0.000078 -0.003910 -2.256325 0.167504 -0.360192 0.036567<br />

a - 2 - 2:30 -0.070511 -0.009096 -3.036197 0.191535 -0.060043 0.035613<br />

a - 3 - 0:00 0.218935 0.034645 0.337938 0.039447 0.126165 0.725246<br />

a - 3 - 0:30 0.036412 -0.594298 0.059926 0.039858 -0.158184 1.890184<br />

a - 3 - 1:30 0.012344 -0.618707 0.021373 0.035472 -0.289620 0.162081<br />

a - 3 - 2:30 0.005022 -0.637078 0.000277 0.040050 -0.421114 0.126305<br />

b - 1 - 0:00 0.058696 0.033665 -0.109527 0.544266 0.005208 0.093669<br />

b - 1 - 0:30 -0.000152 0.000836 -0.062758 0.311392 -0.161699 0.210084<br />

b - 1 - 1:30 0.000136 -0.005232 -0.033282 0.268657 -0.151983 0.007109<br />

b - 1 - 2:30 -0.001141 -0.007456 -0.038459 0.232281 -0.219765 0.006954<br />

b - 2 - 0:00 0.000120 0.027868 -1.587432 0.109482 -0.443133 0.478729<br />

b - 2 - 0:30 -0.000007 -0.011142 -2.043042 0.152084 -0.356465 0.023048<br />

b - 2 - 1:30 -0.068115 -0.014524 -2.539236 0.124697 -0.119275 0.165872<br />

b - 2 - 2:30 -0.001477 -0.013833 -3.356041 0.183860 -0.202332 0.071039<br />

b - 3 - 0:00 0.001240 0.188280 0.178267 0.038147 -0.132203 -0.044270<br />

b - 3 - 0:30 0.003910 -0.641102 0.033766 0.025673 -0.369358 0.379557<br />

b - 3 - 1:30 -0.198913 -0.656871 -0.020963 0.013890 -0.582478 0.584281<br />

b - 3 - 2:30 -0.230061 -0.693729 -0.041146 0.016264 -0.734992 0.764599<br />

Percentage of T R exceeding T C<br />

a - 1 - 0:00 92.9 78.0 78.9 94.0 55.4 96.8<br />

a - 1 - 0:30 87.8 0 0 93.0 0 97.6<br />

a - 1 - 1:30 95.6 0 0 91.3 0 95.4<br />

a - 1 - 2:30 72.7 42.6 0 91.7 41.8 87.9<br />

a - 2 - 0:00 96.5 49.9 0 97.1 0 85.1<br />

a - 2 - 0:30 65.0 0 0 92.7 0 89.0<br />

a - 2 - 1:30 4.8 0 0 91.4 0 94.1<br />

a - 2 - 2:30 0 0 0 87.4 0 86.1<br />

a - 3 - 0:00 91.1 12.0 81.4 99.7 40.0 82.7<br />

a - 3 - 0:30 86.8 0 68.8 99.8 0 91.8<br />

a - 3 - 1:30 43.5 0 44.1 99.7 0 54.5<br />

a - 3 - 2:30 20.4 0 0.6 99.6 0 42.6<br />

b - 1 - 0:00 97.8 90.3 0 95.9 6.8 78.8<br />

193


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

b - 1 - 0:30 0 5.8 0 91.5 0 92.9<br />

b - 1 - 1:30 9.7 0 0 90.4 0 58.9<br />

b - 1 - 2:30 0 0 0 87.5 0 22.9<br />

b - 2 - 0:00 94.9 62.1 0 97.3 0 92.6<br />

b - 2 - 0:30 0 0 0 88.7 0 59.4<br />

b - 2 - 1:30 0 0 0 84.5 0 96.4<br />

b - 2 - 2:30 0 0 0 83.9 0 82.3<br />

b - 3 - 0:00 0.5 42.2 59.0 99.7 0 0<br />

b - 3 - 0:30 32.5 0 44.1 99.4 0 72.4<br />

b - 3 - 1:30 0 0 0 99.4 0 80.9<br />

b - 3 - 2:30 0 0 0 99.3 0 84.1<br />

Mean 41.4 16.0 15.7 94.0 6.0 76.1<br />

Mean 24.3 58.7<br />

Mean 41.5<br />

4 Discussion<br />

A comparison with results of other research groups shows that the measured values are in a<br />

realistic range. Ghebreiyessus [3] measured shear stress values up to 40 Pa, in the experiments<br />

of Nearing et al. [4] Reynolds numbers up to 100000 and unit stream power values up to 10 m<br />

s -1 were reached. Giménez & Govers [5] found unit stream power values up to 0.4 m s -1 and<br />

unit length shear force values up to 6 N m -1 , in the paper of Zhang et al. [9], shear stress<br />

values up to 30 Pa and unit stream power values up to 0.5 m s -1 are published. Govers [13]<br />

measured shear stress values up to 100 Pa and effective stream power values up to 10000 W<br />

m -1 . The presented measurements are in the same order of magnitude and therefore realistic.<br />

But in contradiction with the cited papers, there are no clear linear correlations between<br />

hydraulic parameters and erosion parameters in the results of the field experiments. Therefore,<br />

these outcomes indicate that linear models are not applicable for description of experimental<br />

processes in natural rills. The question to be answered is: why doesn’t this concept that has<br />

been used for over 30 years work?<br />

In literature, four possible reasons can be found. (I) Unclear calculation and definition of the<br />

used parameters, (II) disregarding the influence of turbulence, (III) responsibility of processes<br />

which are not controlled by shear stress for material transport and (IV) a high spatial and<br />

temporal variability of those processes.<br />

The first reason could be the unclear calculation of shear stress. It is to differentiate between<br />

flow shear stress τ, a hydraulic parameter, and critical shear stress τ c or τ cr , a soil parameter<br />

which is similar to soil strength. Shear stress exerted by flow must exceed the critical shear<br />

stress to cause erosion.<br />

194


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

For calculation of shear stress exerted by flowing water, hydraulic parameters, roughness,<br />

flow velocity and fluid density have been taken into account. The actual version of the shear<br />

stress equation calculates the average shear stress by depth averaging of momentum equation<br />

for steady uniform flow per area and time, but this definition doesn’t fit to all definitions used<br />

in literature. Some factors have been developed from empirical studies [15-26]. In most cases,<br />

the theoretical basis of the equations is not clear. The formula used in Chisci et al. [27] is<br />

derived from Landau and Lifchitz [57]. Other versions of the Landau-Lifchitz equation are to<br />

find in literature [2-5].<br />

The critical shear stress is the force needed to detach a soil particle. So it corresponds to a soil<br />

parameter and therefore, input for calculations should also be depending on soil<br />

characteristics. Only in the WEPP model [34], the critical shear stress is calculated using only<br />

soil parameters like texture, organic matter content and dry bulk density. But in some<br />

equations the so called “critical” shear stress consists of hydraulic parameters like water<br />

depth, water width, Manning friction factor or fluid density [16, 28, 33], hydraulic and soil<br />

parameters are used equally [34]. In the studies of Partheniades & Paaswell [30], Andrews<br />

[31] and Andrews & Erman [32] the empirical nature of the development is clearly noticeable.<br />

That means the equations are not deduced from physical laws but from empirical studies.<br />

In many studies [12, 35, 37-41], neither critical shear stress nor shear stress are used for the<br />

calculation of the transport capacity. In other studies shear stress is used to calculate transport<br />

capacity and detachment capacity [36] or transport rate [42] and critical shear stress to<br />

calculate the detachment capacity [36]. In both cases it is clear that shear stress and critical<br />

shear stress operate against each other, the important parameter is the difference between<br />

these two variables.<br />

A summary of these equations can be found in Reid and Dunne [66], on the EPA-homepage<br />

[67] and in Hessel and Jetten [68].<br />

The second reason for the bad correlations between hydraulic parameter and soil detachment<br />

can be the lack of turbulence parameters in the equations.<br />

In the study of Knapen et al. [14] the Reynolds number shows very different correlations to<br />

the detachment rate, the same is to notice in the given results of this study. The reason could<br />

be that the turbulence, described by the Reynolds number, does not directly operate on<br />

substrate, it influences the acting shear stress, that means the calculated shear stress is much<br />

lower than the operating shear stress. This relation has been confirmed in several studies.<br />

Nearing et al. [69] found that turbulence can increase the active shear stress by a factor of<br />

several thousands. They measured flow shear stresses ranged from 0.5 to 2 Pa, while tensile<br />

195


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

strengths ranged from 1 to 2 kPa, a difference in magnitude of 1000. Despite this conflict,<br />

detachment rates of nearly 300 g m -2 s -1 were measured. He explained this result with<br />

turbulent burst events which are much greater than the average flow shear stresses. Another<br />

study about the influence of turbulence on detachment rates was published by Nearing &<br />

Parker [70]. They showed that under turbulent flow conditions the same shear stress value<br />

caused a clearly higher detachment rate. In their flume experiments the difference between<br />

detachment rate caused by turbulent and laminar flow increased with increasing shear stress<br />

value. That means, if given hydraulic conditions lead to a high shear stress value, the<br />

influence of turbulence on soil erosion is higher than in low shear stress value ranges.<br />

The shear stress equation as well as those related to the other hydraulic parameters assumes<br />

that drag forces are dominant for controlling erosion. But rill erosion is the result of the<br />

combination of different processes including headcut erosion, sidewall sloughing, tunnelling,<br />

micro-piping, slaking piping and sapping [14, 45, 71-75]. This is the third reason for the<br />

problems of the model equations. The percentage of headcutting in the different studies<br />

ranges between “four times higher than the contribution of bed scours” [75] to “60 % of total<br />

rill erosion” [76]. <strong>Stefan</strong>ovic and Bryan [77] showed that concentrated flow causes sediment<br />

production primarily from knickpoints, chutes, meanders and bank failure. Govers [45]<br />

distinguished between hydraulic erosion, mass wasting processes on rill sidewalls, gullying<br />

and piping. Hydraulic rill erosion mostly occurred during three extreme runoff events. Mass<br />

wasting processes caused 37 % of total erosion in rills. Gullying, the retreat erosion at<br />

knickpoints and headcuts caused about 12% of rill erosion rates. In the presented experiments,<br />

mainly mass wasting and gullying processes occurred, so the correlations between hydraulic<br />

parameters and detachment rate are mostly low. But the hydraulic rill erosion only occurs in<br />

extreme runoff events, in most cases, the runoff values are too low to cause this process. The<br />

percentage of material which is transported independent of shear stress is very high on the<br />

water front samples. Here the transport of loose material is probably more important than in<br />

the other samples. That means this process is mainly independent of shear stress. In these<br />

cases of transport rate vs. transport capacity < 1 the independence of shear stress cannot be<br />

excluded, in the other cases the processes controlled by shear stress can occur. It is to<br />

conclude that in the case of T R > T C , not only shear stress controlled processes provide the<br />

material; at least the difference between T R and T C is caused by processes independent of<br />

shear stress.<br />

The presented experiments show that the correlation between hydraulic parameters and<br />

detachment rate does neither change from one experiment to another nor from one run to<br />

196


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

another but from one measuring point and run to another. Thus, sediment producing processes<br />

change with high spatial and temporal variability. This is the fourth reason for model<br />

problems. It is very difficult to propose a single factor that always describes the soil<br />

detachment satisfactory. The high variability of erosion processes, even under controlled<br />

experimental conditions, has been shown in different studies, measured variability shows a<br />

wide range between 3.4 and 173.2 % [78-83]. This is partially the result of non-homogeneous<br />

parameters concerning soil characteristics and rainfall. On experimental plots, infiltration<br />

rates and soil aggregate stability can be highly variable [84] as well as the rainfall shows a<br />

high spatial and temporal variability [85]. Therefore, the input parameters to the different<br />

measurements reflected in the mentioned studies were not really comparable. Nevertheless,<br />

the results also make clear that modelling soil erosion has to include uncertainty in model<br />

input as well as in the data used for model calibration and validation.<br />

In the field, the spatial and temporal variability of soil conditions cannot be avoided, and is,<br />

furthermore, part of the investigations. Thus, additional input parameters as rainfall or flow<br />

should be maintained constant in the experiments to generate reproducible data. The high<br />

variability in soil erosion processes cannot be represented by a single factor like shear stress.<br />

The results show that there is not a simple linear correlation between a certain hydraulic<br />

parameter and soil detachment rate. Depending on model purpose and scale, the factors can be<br />

used to predict the magnitude of rill detachment but they are not applicable for the simulation<br />

of rill erosion with high-resolution spatial and temporal change in processes.<br />

A newer approach is the use of probability density functions to predict soil detachment [86,<br />

87]. Sidorchuk gives two sources of stochasticity in erosion modelling: (1) the necessity of<br />

spatial and temporal averaging when determining deterministic equations, which describe<br />

concentrated flow erosion and (2) the fact that the main erosion factors, if these can be<br />

determined anyway, can only be measured with limited accuracy. This is not the first attempt<br />

to model erosion by relating the probability of soil detachment with the excess of erosion<br />

driving forces over soil erosion resistance forces. Other papers were published by Nearing<br />

[88], Wilson [89] and Sidorchuk [90-95]. One of the earliest publications about stochastic in<br />

erosion processes has been published by Einstein [96]. These stochastic models reduce the<br />

number of empirical components, but this approach seems to be too complicated [14] – yet.<br />

5 Conclusion<br />

The results show that simple linear correlation between hydraulic parameter and soil<br />

detachment is not suitable, at least in natural rills. The reason for that is the combination of<br />

197


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

different processes that cause different amounts of soil erosion. Shear stress only describes<br />

one process. The results clearly show that there is not one fixed parameters that always<br />

predict soil detachment best. Applicability of one certain hydraulic parameter to predict the<br />

sediment concentration changes at a certain point in time within few minutes because the<br />

temporal and spatial distribution of the different erosion processes is highly randomly<br />

determined. Therefore, it might be more useful to formulate results in probalistic terms. This<br />

is required since the 1990s [89, 91] and implemented since the 2000s especially by the<br />

working group of Sidorchuk [90, 94-98]. But first, much more field experiments are needed to<br />

provide the required data.<br />

Acknowledgement<br />

The research was supported by the ‘Internationale Graduiertenzentrum’ of Trier University<br />

and the ‘Freundeskreis Trierer Universität e.V.’. Additionally, we thank all participants of the<br />

field trip to Andalusia in September 2009 who supported the performance of the experiments<br />

and Olli and Seta for revising the whole paper.<br />

References<br />

1. Lyle WM, Smerdon ET (1965) Relation of compaction and other soil properties to<br />

erosion resistance of soils. Transactions of the ASAE 8: 419-422.<br />

2. Torri D, Dfalanga M, Chisci G (1987) Threshold conditions for incipient rilling.<br />

Catena Supplement 8: 97-105.<br />

3. Ghebreiyessus YT, Gantzer CJ, Alberts EE, Lentz RW (1994) Soil erosion by<br />

concentrated flow: shear stress and bulk density. Transactions of the ASAE 37 (6):<br />

1791-1797.<br />

4. Nearing MA, Norton LD, Bulgakov DA, Larionov GA, West LT, Dontsova KM<br />

(1997) Hydraulics and Erosion in Eroding Rills. Water Resources Research 33 (4):<br />

865-876.<br />

5. Giménez R, Govers G (2002) Flow Detachment by Concentrated Flow on Smooth<br />

and Irregular Beds. Soil Sci Soc Am J 66(5): 1475-1483.<br />

6. Bagnold RA (1977) Bed load transport by natural rivers. Water resources Research<br />

13: 303-312.<br />

7. Hairsine PB, Rose CW (1992) Modeling water erosion due to overland flow using<br />

physical principles, 2. Rill flow. Water resources Research 28: 245-250.<br />

198


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

8. Elliot WJ, Laflen JM (1993) A Process-based rill erosion model. Transactions of<br />

the ASAE 36 (1): 35-72.<br />

9. Zhang G, Liu B, Liu G, He X, Nearing MA (2003): Detachment of undisturbed<br />

soil by shallow flow. Soil Science Society of America Journal 67: 713-719.<br />

10. Yang CT (1972) Unit stream power and sediment transport. J. Hydraulics Div.<br />

Am. Soc. Civil Eng. 98: 1805-1825.<br />

11. Moore ID, Burch GJ (1986) Sediment transport capacity of sheet and rill flow:<br />

Application to unit stream power theory. Water Resour. Res. 22: 1350-1360.<br />

12. Bagnold RA (1980): An empirical correlation of bedload transport rates in flumes<br />

and natural rivers. Proc. Royal Society Series A372: 453-473.<br />

13. Govers G (1992) Evaluation of transport capacity formulae for overland flow. In:<br />

Parsons AJ, Abrahams AD editors. Overland flow: hydraulics and erosion<br />

mechanics. London: UCL Press. pp. 243-273.<br />

14. Knapen A, Poesen J, Govers G, Gyssels G, Nachtergaele J (2007): Resistance of<br />

soils to concentrated flow erosion: A review. Earth-Science Reviews 80(1-2): 75-<br />

109.<br />

15. Foster GR (1982) Modelling the erosion process. In: Johnson HP, Brakensiek DL,<br />

editors. Hydraulic Modelling of Small Watersheds. St. Joseph, MI: ASAE<br />

monograph 5, Hann CT.<br />

16. Shields A (1936) Anwendung der Ähnlichkeitsmechanik und der<br />

Turbulenzforschung auf die Geschiebebewegung. Berlin: Dissertation, TU Berlin.<br />

17. Miller M, McCave IN, Komar PD (1977) Threshold sediment motion under<br />

unidirectional currents. Sedimentology 24: 507-527.<br />

18. Parker G, Klingeman PC, McLean DG (1982) Bedload and size distribution in<br />

paved gravel-bed streams. Journal of the Hydraulics Division, American Society of<br />

Civil Engineers 108: 544-571.<br />

19. Diplas P (1987) Bedload transport in gravel-bed streams. J. Hydraul. Eng. ASCE<br />

113: 277-292.<br />

20. Parker G (1990) Surface-based bedload transport relation for gravel rivers. J.<br />

Hydraul. Res. 28: 417–436.<br />

21. Komar PD (1987 a) Selective grain entrainment by a current from a bed of mixed<br />

sizes: A reanalysis. J. Sediment. Petrol. 57: 203-211.<br />

22. Komar PD (1987 b) Selective gravel entrainment and the empirical evaluation of<br />

flow competence. Sedimentology 34: 1165-1176.<br />

199


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

23. Andrews ED (1983) Entrainment of gravel from naturally sorted riverbed material.<br />

Geol. Soc. Am. Bull. 94: 1225-1231.<br />

24. Ashworth PJ, Ferguson RI (1989 a) Size-selective entrainment of bed load in<br />

gravel bed streams. Water Resources Research 25: 627-634.<br />

25. Ashworth PJ, Ferguson RI (1989 b) Quantifying gravel deposition on river bars<br />

using flexible netting. Journal of Sedimentary Research Vol. 59 No. 4: 623-624.<br />

26. Komar PD, Carling PA (1991) Grain sorting in gravel-bed streams and the choice<br />

of particle sizes for flow-competence evaluations. Sedimentology 38: 489-502.<br />

27. Chisci G, Sfalanga M, Torri D (1985) An experimental model for evaluating soil<br />

erosion on a single-rainstorm basis. In: El-Swaify SA, Moldenhauer WC, Lo A,<br />

editors. Soil Erosion and Conservation. Ankeny, Iowa: Soil conservation Society<br />

of America. pp. 558-565.<br />

28. Ott WP, van Uchelen JC (1936) Application of similarity principles and turbulence<br />

research to bedload movement. Hydrodynamics Laboratory California Institute of<br />

Technology Publication No. 167. Soil conservation Service.<br />

29. Graf WH (1971) Hydraulics of sediment transport. New York: McGraw-Hill.<br />

30. Partheniades E, Paaswell RE (1970) Erodibility of channels with cohesive<br />

boundary. Journal of the Hydraulic Division, Proceedings of the American Society<br />

of civil Engineers: 755-771.<br />

31. Andrews ED (1984) Bed-material entrainment and hydraulic geometry of gravelbed<br />

rivers in Colorado. Geological Society of America Bulletin 95(3): 371-378.<br />

32. Andrews ED, Erman DC (1986) Persistence in the size distribution of superficial<br />

bed material during an extreme snowmelt flood. Water Resources Research 22:<br />

191-197.<br />

33. De Ploey J (1990) Threshold conditions for thalweg gullying with special<br />

reference to loess areas. Catena Supplement 17: 147-151.<br />

34. Flanagan DC, Livingston SJ (1995) WEPP user summary, USDA-water erosion<br />

prediction project. NSERL Report No. 11. National soil erosion research<br />

Laboratory.<br />

35. Yalin MS (1963) An expression for bedload transportation. J. Hydr. Eng. Div.<br />

ASCE 89: 221-250.<br />

36. Foster GR, Flanagan DC, Nearing MA, Lane LJ, Risse LM, Finkner SC (1995)<br />

Hillslope erosion component. In: USDA Water Erosion Prediction Project:<br />

200


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Hillslope Profile and Watershed Model Documentation, Chapter 11, USDA-ARS<br />

NSERL Report 10.<br />

37. Govers G (1991) Rill erosion on arable land in central Belgium: Rates, controls<br />

and predictability. Catena 18: 133-155.<br />

38. Abrahams AD, Gao P, Aebly FA (2000) Relation of sediment transport capacity to<br />

stone cover and size in rain-impacted interrill flow. Earth Surf. Proc.Land. 25:<br />

497-504.<br />

39. Low HS (1989) Effect of sediment density on bed-load transport. J. Hydraul. Eng.<br />

ASAE 115: 124-138.<br />

40. Rickenmann D (1991) Hyperconcentrated flow and sediment transport at steep<br />

slopes. J. Hydraul. Eng. ASCE 117: 1419-1439.<br />

41. Yang CT (1973) Incipient motion and sediment transport. J. Hydr. Eng. Div.<br />

ASCE 99: 1679-1703.<br />

42. Parker G (1979) Hydraulic geometry of active gravel rivers. Journal of the<br />

Hydraulics Division, American Society of Civil Engineers 105: 1185-1201.<br />

43. Kleinhans MG, Bierkens MFP, van der Perk M (2010) HESS opinions. On the use<br />

of laboratory experimentation: „Hydrologists, bring out shovel and garden hoses<br />

and hit the dirt“. Hydrology and Earth System Sciences 14: 369-382.<br />

44. Govers G, Giménez R, van Oost K (2007) Exploring the relationsship between<br />

experiments, modelling and field observations. Earth-Science Reviews 84 (3-4):<br />

87-102.<br />

45. Govers G (1987) Spatial and temporal variability in rill development processes at<br />

the Huldenberg experimental site. Catena Supplement 8: 17-34.<br />

46. Seeger M (2007) Uncertainty of factors determining runoff and erosion processes<br />

as quantified by rainfall simulations. Catena 71 (1): 56-67.<br />

47. FAO (2006) Guidelines for soil description. Rome: 4. ed. Food and Agriculture<br />

Organization of the United Nations.<br />

48. Vandekerckhove L, Poesen J, Oostwoud Wijdenes D, de Figueiredo T (2003)<br />

Topographical thresholds for ephemeral gully initiation in intensively cultivated<br />

areas of the Mediterranean. Catena 33(3-4): 271-292.<br />

49. Ad-hoc-Arbeitsgruppe Boden (2005) Bodenkundliche Kartieranleitung. Hannover:<br />

5. Auflage, Bundesamt für <strong>Geowissenschaften</strong> und Rohstoff in Zusammenarbeit<br />

mit den staatlichen Geologischen Diensten der Bundesrepublik Deutschland. 438<br />

p.<br />

201


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

50. <strong>Wirtz</strong> S, Seeger M, Ries JB (2010) The rill experiment as a method to approach a<br />

quantification of rill erosion process activity. Zeitschrift für Geomorphologie Vol.<br />

54,1: 47-64.<br />

51. Allen JRL (1994) Fundamental properties of fluids and their relation to sediment<br />

transport processes. In: Pye K, editor. Sediment transport and depositional<br />

processes. Blackwell Scientific Publications. pp. 25-88.<br />

52. DIN 18124 (1997) Baugrund, Untersuchung von Bodenproben - Bestimmung der<br />

Korndichte - Kapillarpyknometer, Weithalspyknometer. Deutsches Institut für<br />

Normung e.V., Ausgabe : 1997-07 , Deutsch.<br />

53. Macosko CW (1994) Rheology: Principles, Measurements and Applications. New<br />

York: VCH: Wiley-VCH. New York.<br />

54. Einstein A (1906) Eine neue Bestimmung der Moleküldimensionen. Analen der<br />

Physik 19: 289-306.<br />

55. Einstein A (1911) Berichtigung zu meiner Arbeit: „Eine neue Bestimmung der<br />

Moleküldimensionen“. Analen der Physik 34: 591-593.<br />

56. Scherer U (2008): Prozessbasierte Modellierung der Bodenerosion in einer<br />

Lösslandschaft. Karlsruhe: Disserationsschrift, Fakultät für Bauingenieur-, Geound<br />

Umweltwissenschaften, Universität Fridericiana zu Karlsruhe (TH). 248 p.<br />

57. Landau L, Lifchitz E (1971) Mécaniques des fluids. Moscow: MIR<br />

58. Fan, J., Gao, H., Guo, B., 2010. Regularity criteria for the Navier-Stokes-Landau-<br />

Lifshitz system. J.Math. Anal. Appl. 363, 29-37<br />

59. Bell, J.B., Garcia, A.L., Williams, S.S. 2007. Numerical methods for the stochastic<br />

Landau-Lifshitz Navier-Stokes equations. Physical Review E 76<br />

60. Seiler R (2002) Die Navier-Stokes-Gleichung. Elemente der Mathematik 57: 109-<br />

114.<br />

61. Constantin P (2001) Some open problems and research directions in the<br />

mathematical study of fluid dynamics. Mathematics unlimited-2001 and beyond:<br />

353-360.<br />

62. Fefferman CL (2006) Existance and smoothness of the Navier-Stokes equation.<br />

The millennium prize problems. Cambridge, MA.: Clay Math. Inst. pp. 57-67.<br />

63. Temam R (2000) Some developments on Navier-Stokes equations in the second<br />

half of the 20 th century. Development of mathematics 1950-2000: 1049-1106.<br />

64. Wiegner M (1999) The Navier-Stokes equations – a neverending challenge?<br />

Jahresbericht Deutsch. Math.-Verein 101, No. 1: 1-25.<br />

202


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

65. Schneider G (2008) Ein Millenniumsproblem: Die globale Existenz und<br />

Eindeutigkeit glatter Lösungen der dreidimensionalen Navier-Stokes-Gleichungen.<br />

Mathematik-Online: Beiträge zu berühmten, gelösten und ungelösten Problemen,<br />

Nummer 2.<br />

66. Reid DM, Dunne T (1996) Rapid evaluation of sediment budgets. Reiskirchen:<br />

Catena Verlag.<br />

67. EPA-homepage (2009) Channel Processes: Bedload transport. United States<br />

Environmental protection agency. Available:<br />

http://water.epa.gov/scitech/datait/tools/warsss/bedload.cfm, access: 14.9.2010<br />

68. Hessel R, Jetten V (2007) Suitability of transport equations in modelling soil<br />

erosion for a small Loess plateau catchment. Eng. Geol. 91: 56-71.<br />

69. Nearing MA, Bradford JM, Parker SC (1991) Soil detachment by shallow flow at<br />

low slopes. Soil Science Society of America Journal 55 (2): 339-344.<br />

70. Nearing MA, Parker SC (1994) Detachment of soil by flowing water under<br />

turbulent and laminar conditions. Soil Science Society of America Journal 58 (6):<br />

1612-1614.<br />

71. Bryan RB, Govers G, Poesen J (1989) The concept of soil erodibility and some<br />

problems of assessment and application. Catena 16 (4-5): 393-412.<br />

72. Bryan RB (1990) Knickpoint evolution in rillwash. Catena 17: 111-132.<br />

73. Owoputi LO, Stolte WJ (1995) Soil detachment in the physically based soil<br />

erosion process: a review. Transactions of the ASAE 38 (4): 1099-1110.<br />

74. Rapp I. (1998) Effects of soil properties and experimental conditions on the rill<br />

erodibilities of selected soils. Pretoria: Ph. D. Thesis, Faculty of Biological and<br />

Agricultural Sciences, University of Pretoria, South Africa.<br />

75. Zhu JC, Gantzer CJ, Peyton RL, Alberst EE, Anderson SH (1995) Simulated<br />

small-channel bed scour and head cut erosion rates compared. Soil Science Society<br />

of America Journal 59 (1): 211-218.<br />

76. Kohl KD (1988) Mechanics of rill headcutting. Ames: Phd. Diss. Iowa State<br />

University.<br />

77. <strong>Stefan</strong>ovic JR, Bryan RB (2009) Flow energy and channel adjustments in rills<br />

developed in loamy sand and sandy loam soils. Earth Surface Processes and<br />

Landforms 34: 133-144.<br />

78. Nearing MA (1998) Why soil erosion models over-predict small soil losses and<br />

under-predict large soil losses. Catena 32: 15-22.<br />

203


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

79. Ruttimann M, Schaub D, Prasuhn V, Ruegg W (1995) Measurement of runoff and<br />

soil erosion on regulary cultivated fields in Switzerland – some critical<br />

considerations. Catena 25: 127-139.<br />

80. Wendt RC, Alberts EE, Hjelmfelt Jr (1986). Variability of runoff and soil loss<br />

from fallow experimental plots. Soil Sci. Soc. Am. J. 50: 730-736.<br />

81. Risse LM, Nearing MA, Nicks AD, Laflen JM (1993) Assessment of error in the<br />

universal soil loss equation. Soil Sci. Soc. Am. J. 57: 825-833.<br />

82. Zhang XC, Nearing MA, Risse LM, McGregor KC (1996) Evaluation of runoff<br />

and soil loss predictions using natural runoff plot data. Trans. ASAE 39: 855-863.<br />

83. Liu BY, Nearing MA, Baffaut C, Ascough II JC (1996) The WEPP watershed<br />

model: III. Comparisons to measured data from small watersheds. Transactions of<br />

the ASAE 40: 945-951.<br />

84. Ajayi AE, Horta IDMF (2007) The effect of spatial variability of soil hydraulic<br />

properties on surface runoff processes. Anais XIII Simpósio Brasileiro de<br />

Sensoriamento Remoto, Florianópolis, Brasil, 21-26 abril 2007, p. 3243-3248.<br />

85. Dunkerley D (2008) Rain event properties in nature and in rainfall simulation<br />

experiments: a comparative review with recommendations for increasingly<br />

systematic study and reporting. Hydrological Processes 22: 4415-4435.<br />

86. Sidorchuk A (2005 b) Stochastic modelling of erosion and deposition in cohesive<br />

soils. Hydrological Processes 19: 1399-1417.<br />

87. Sidorchuk A (2009 b) A third generation erosion model: The combination of<br />

probabilistic and deterministic components. Geomorphology 110 (1-2): 2-10.<br />

88. Nearing MA (1991) A probabilistic model of soil detachment by shallow turbulent<br />

flow. Transactions of the ASAE 34 (1): 81-85.<br />

89. Wilson BN (1993) Development of a fundamentally-based detachment model.<br />

Transactions of the ASAE 36 (4): 1105-1114.<br />

90. Sidorchuk A (2001) Calculation of the rate of erosion in soils and cohesive<br />

sediments. Eurasien Soil Science 34 (8): 893-900.<br />

91. Sidorchuk A (2002) Stochastic Modelling of soil erosion and deposition. 12 th<br />

ISCO Conference, Beijing 2002.<br />

92. Sidorchuk A, Smith A, Nikora V (2004) Probability distribution function approach<br />

in stochastic modelling of soil erosion. Sediment transfer trough the fluvial system<br />

Vol. 288. IHAS Publ., pp. 345-353.<br />

204


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

93. Sidorchuk A (2005 a) Stochastic components in the gully erosion modelling.<br />

Catena 63: 299-317.<br />

94. Sidorchuk A, Schmidt J, Cooper G (2008) Variability of shallow overland flow<br />

velocity and soil aggregate transport observed with digital videography.<br />

Hydrological Processes 22: 4035-4048.<br />

95. Sidorchuk A (2009 a) High-Frequency variability of aggregate transport under<br />

water erosion of well-structured soils. Eurasian Soil Science Vol. 42, No. 5: 543-<br />

552.<br />

96. Einstein HA (1936) Der Geschiebetrieb als Wahrscheinlichkeitsproblem. Zürich:<br />

Diss.-Druckerei A.-G. Gebr. Leemann & Co. 112 p.<br />

205


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Table S1 Freila 1 erosion data<br />

Run - MP - flow length [m]-<br />

sampling time [min:sec]<br />

Sediment<br />

Concentration<br />

[g L -1 ]<br />

Detachment<br />

rate<br />

[kg s -1 m -2 ]<br />

Transport<br />

rate<br />

[kg s-1]<br />

Sample<br />

density<br />

[g cm -3 ]<br />

Slope<br />

[°]<br />

Transport<br />

capacity<br />

[kg s-1]<br />

a-1-3.4-0:00 10.6 0.0156 0.017750523 1.01 2.4 0.00125<br />

a-1-3.4-0:30 2.1 0.0092 0.010500304 1.00 2.4 0.00128<br />

a-1-3.4-1:30 3.3 0.0257 0.029404312 1.00 2.4 0.00129<br />

a-1-3.4-2:30 0.4 0.0040 0.004617795 1.00 2.4 0.00126<br />

a-2-8.6-0:00 14.9 0.0008 0.000726251 1.01 7.4 0.00003<br />

a-2-8.6-0:30 2.4 0.0004 0.000914653 1.00 7.4 0.00032<br />

a-2-8.6-1:30 1.4 0.0005 0.001633948 1.00 7.4 0.00156<br />

a-2-8.6-2:30 0.6 0.0010 0.007309575 1.00 7.4 0.07782<br />

a-3-13.1-0:00 26.6 0.0407 0.240254717 1.02 6 0.02132<br />

a-3-13.1-0:30 8.0 0.0079 0.041958790 1.01 6 0.00555<br />

a-3-13.1-1:30 3.1 0.0049 0.028349123 1.00 6 0.01600<br />

a-3-13.1-2:30 2.5 0.0042 0.024596630 1.00 6 0.01957<br />

b-1-3.4-0:00 35.1 0.0524 0.060019839 1.02 2.4 0.00132<br />

b-1-3.4-0:30 0.5 0.0010 0.001109306 1.00 2.4 0.00126<br />

b-1-3.4-1:30 0.6 0.0012 0.001397038 1.00 2.4 0.00126<br />

b-1-3.4-2:30 0.1 0.0002 0.000302770 1.00 2.4 0.00144<br />

b-2-8.6-0:00 4.8 0.0001 0.000126626 1.00 7.4 0.00001<br />

b-2-8.6-0:30 0.6 0.0001 0.000260038 1.00 7.4 0.00027<br />

b-2-8.6-1:30 0.7 0.0013 0.010225583 1.00 7.4 0.07834<br />

b-2-8.6-2:30 0.1 0.0000 0.000132974 1.00 7.4 0.00161<br />

b-3-13.1-0:00 9.9 0.0336 0.247153306 1.01 6 0.24591<br />

b-3-13.1-0:30 2.6 0.0021 0.012025713 1.00 6 0.00812<br />

b-3-13.1-1:30 1.0 0.0042 0.030456868 1.00 6 0.22937<br />

b-3-13.1-2:30 0.8 0.0047 0.035191499 1.00 6 0.26525<br />

206


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Table S2 Freila 1 runoff data<br />

Run - MP - flow length [m]-<br />

sampling time [min:sec]<br />

Flow<br />

velocity<br />

[m s -1 ]<br />

Dynamic<br />

viscosity<br />

[kg s -1 m -1 ]<br />

Water<br />

depth<br />

[cm]<br />

Flow cross<br />

section<br />

[cm²]<br />

Wetted<br />

Perimeter<br />

[cm]<br />

Hydraulic<br />

radius<br />

[cm]<br />

a-1-3.4-0:00 0.41 0.001053 0.20 40.9 33.5 1.2<br />

a-1-3.4-0:30 1.20 0.001011 0.30 41.7 33.7 1.2<br />

a-1-3.4-1:30 2.12 0.001017 0.30 41.7 33.7 1.2<br />

a-1-3.4-2:30 2.61 0.001002 0.40 42.0 34.2 1.2<br />

a-2-8.6-0:00 0.34 0.001075 0.30 1.4 10.9 0.1<br />

a-2-8.6-0:30 0.39 0.001012 1.30 10.0 27.6 0.4<br />

a-2-8.6-1:30 0.41 0.001007 2.50 28.0 41.1 0.7<br />

a-2-8.6-2:30 0.41 0.001003 10.30 289.7 88.9 3.3<br />

a-3-13.1-0:00 0.92 0.001133 3.60 98.3 45.1 2.2<br />

a-3-13.1-0:30 1.01 0.001040 2.40 51.8 40.5 1.3<br />

a-3-13.1-1:30 1.05 0.001016 3.10 86.9 44.3 2.0<br />

a-3-13.1-2:30 1.06 0.001012 3.30 94.9 44.6 2.1<br />

b-1-3.4-0:00 0.41 0.001176 0.30 41.7 33.7 1.2<br />

b-1-3.4-0:30 0.57 0.001002 0.40 42.0 34.2 1.2<br />

b-1-3.4-1:30 0.59 0.001003 0.40 42.0 34.2 1.2<br />

b-1-3.4-2:30 0.91 0.001000 0.50 47.2 36.3 1.3<br />

b-2-8.6-0:00 0.32 0.001024 0.50 0.8 10.9 0.1<br />

b-2-8.6-0:30 0.49 0.001003 1.30 9.4 28.0 0.3<br />

b-2-8.6-1:30 0.53 0.001003 10.40 290.7 89.0 3.3<br />

b-2-8.6-2:30 0.46 0.001001 2.40 26.1 37.7 0.7<br />

b-3-13.1-0:00 0.76 0.001050 9.20 327.5 56.2 5.8<br />

b-3-13.1-0:30 0.72 0.001013 2.80 64.0 42.8 1.5<br />

b-3-13.1-1:30 0.98 0.001005 9.20 318.0 55.9 5.7<br />

b-3-13.1-2:30 1.32 0.001004 9.30 346.0 57.4 6.0<br />

207


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Table S3 Freila 1 hydraulic data<br />

Run - MP - flow length [m]- τ Г ω<br />

sampling time [min:sec] [Pa] [N m -1 ] [W m -2 ]<br />

ω U ω eff<br />

Re τ - τ cr<br />

[m s -1 ] [W m -1 ] [ ] [Pa]<br />

a-1-3.4-0:00 5.05 1.69 2.07 0.02 187.91 4791 3.08<br />

a-1-3.4-0:30 5.09 1.72 6.10 0.05 725.17 14707 3.12<br />

a-1-3.4-1:30 5.09 1.72 10.79 0.09 1703.78 25834 3.12<br />

a-1-3.4-2:30 5.05 1.73 13.18 0.11 1898.77 32015 3.08<br />

a-2-8.6-0:00 1.68 0.18 0.57 0.04 20.68 420 -0.30<br />

a-2-8.6-0:30 4.58 1.26 1.77 0.05 42.55 1384 2.61<br />

a-2-8.6-1:30 8.62 3.54 3.53 0.05 77.65 2776 6.64<br />

a-2-8.6-2:30 41.19 36.61 16.89 0.05 315.87 13326 39.22<br />

a-3-13.1-0:00 22.72 10.25 20.90 0.10 876.60 17996 20.75<br />

a-3-13.1-0:30 13.20 5.34 13.31 0.11 583.29 12474 11.23<br />

a-3-13.1-1:30 20.14 8.93 21.11 0.11 982.55 20267 18.17<br />

a-3-13.1-2:30 21.83 9.74 23.08 0.11 1077.88 22237 19.86<br />

b-1-3.4-0:00 5.20 1.75 2.13 0.02 149.50 4412 3.23<br />

b-1-3.4-0:30 5.05 1.73 2.90 0.02 196.40 7053 3.08<br />

b-1-3.4-1:30 5.05 1.73 2.98 0.02 204.64 7245 3.08<br />

b-1-3.4-2:30 5.33 1.94 4.84 0.04 363.78 11770 3.36<br />

b-2-8.6-0:00 0.96 0.10 0.31 0.04 5.81 237 -1.01<br />

b-2-8.6-0:30 4.25 1.19 2.08 0.06 54.38 1644 2.28<br />

b-2-8.6-1:30 41.30 36.75 21.83 0.07 461.19 17220 39.33<br />

b-2-8.6-2:30 8.73 3.29 4.00 0.06 96.25 3166 6.76<br />

b-3-13.1-0:00 60.18 33.79 45.74 0.08 1517.71 42492 58.21<br />

b-3-13.1-0:30 15.35 6.58 11.01 0.08 396.41 10602 13.38<br />

b-3-13.1-1:30 58.39 32.63 56.94 0.10 2108.18 55255 56.42<br />

b-3-13.1-2:30 61.89 35.49 81.83 0.14 3606.26 79497 59.92<br />

208


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Table S4 Freila 2 erosion data<br />

Run - MP - flow length [m]-<br />

sampling time [min:sec]<br />

Sediment<br />

Concentration<br />

[g L -1 ]<br />

Detachment<br />

rate<br />

[kg s -1 m -2 ]<br />

Transport<br />

rate<br />

[kg s-1]<br />

Sample<br />

density<br />

[g cm -3 ]<br />

Slope<br />

[°]<br />

Transport<br />

capacity<br />

[kg s-1]<br />

a-1-4-0:00 16.8 0.0668 0.267274644 1.01 5.7 0.05891<br />

a-1-4-0:30 2.4 0.0078 0.023207900 1.00 5.7 0.02325<br />

a-1-4-1:30 1.4 0.0058 0.014046683 1.00 5.7 0.02362<br />

a-1-4-2:30 2.3 0.0086 0.019756646 1.00 5.7 0.01134<br />

a-2-8.5-0:00 11.3 0.0079 0.033779926 1.01 3.4 0.01692<br />

a-2-8.5-0:30 3.9 0.0033 0.014486969 1.00 3.4 0.01837<br />

a-2-8.5-1:30 2.8 0.0028 0.011787666 1.00 3.4 0.01570<br />

a-2-8.5-2:30 1.8 0.0023 0.010364921 1.00 3.4 0.01946<br />

a-3-13.3-0:00 19.2 0.0239 0.287705065 1.01 7.4 0.25306<br />

a-3-13.3-0:30 6.1 0.0127 0.181365395 1.00 7.4 0.77566<br />

a-3-13.3-1:30 3.1 0.0070 0.098724460 1.00 7.4 0.71743<br />

a-3-13.3-2:30 2.3 0.0057 0.079795433 1.00 7.4 0.71687<br />

b-1-4-0:00 12.4 0.0193 0.037281867 1.01 5.7 0.00362<br />

b-1-4-0:30 2.0 0.0062 0.014347225 1.00 5.7 0.01351<br />

b-1-4-1:30 0.8 0.0027 0.006088968 1.00 5.7 0.01132<br />

b-1-4-2:30 0.4 0.0017 0.003861665 1.00 5.7 0.01132<br />

b-2-8.5-0:00 12.0 0.0103 0.044903275 1.01 3.4 0.01704<br />

b-2-8.5-0:30 1.6 0.0018 0.008409765 1.00 3.4 0.01955<br />

b-2-8.5-1:30 0.8 0.0010 0.005012019 1.00 3.4 0.01954<br />

b-2-8.5-2:30 0.7 0.0012 0.005702665 1.00 3.4 0.01954<br />

b-3-13.3-0:00 38.0 0.0371 0.445794531 1.02 7.4 0.25751<br />

b-3-13.3-0:30 2.9 0.0092 0.132228769 1.00 7.4 0.77333<br />

b-3-13.3-1:30 1.6 0.0042 0.059490319 1.00 7.4 0.71636<br />

b-3-13.3-2:30 1.4 0.0016 0.022511209 1.00 7.4 0.71624<br />

209


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Table S5 Freila 2 runoff data<br />

Run - MP - flow length [m]-<br />

sampling time [min:sec]<br />

Flow<br />

velocity<br />

[m s -1 ]<br />

Dynamic<br />

viscosity<br />

[kg s -1 m -1 ]<br />

Water<br />

depth<br />

[cm]<br />

Flow cross<br />

section<br />

[cm²]<br />

Wetted<br />

Perimeter<br />

[cm]<br />

Hydraulic<br />

radius<br />

[cm]<br />

a-1-4-0:00 0.47 0.001084 4.7 338.62 99.98 3.39<br />

a-1-4-0:30 0.54 0.001012 2.8 174.04 74.13 2.35<br />

a-1-4-1:30 0.69 0.001007 2.4 142.28 60.20 2.36<br />

a-1-4-2:30 0.84 0.001012 1.6 100.62 57.12 1.76<br />

a-2-8.5-0:00 0.21 0.001057 0.3 142.06 50.60 2.81<br />

a-2-8.5-0:30 0.24 0.001020 0.5 152.10 52.27 2.91<br />

a-2-8.5-1:30 0.31 0.001014 0.2 134.99 49.38 2.73<br />

a-2-8.5-2:30 0.38 0.001009 0.5 152.10 52.27 2.98<br />

a-3-13.3-0:00 0.32 0.001096 4.2 469.04 90.43 5.19<br />

a-3-13.3-0:30 0.34 0.001031 8.2 877.69 107.58 8.16<br />

a-3-13.3-1:30 0.37 0.001016 8 838.70 105.94 7.92<br />

a-3-13.3-2:30 0.41 0.001012 8 838.70 105.94 7.92<br />

b-1-4-0:00 0.56 0.001062 0.8 53.72 48.34 1.11<br />

b-1-4-0:30 0.67 0.001010 1.8 109.04 57.70 1.89<br />

b-1-4-1:30 0.80 0.001004 1.6 100.62 57.12 1.76<br />

b-1-4-2:30 0.87 0.001002 1.6 100.62 57.12 1.76<br />

b-2-8.5-0:00 0.26 0.001060 0.4 143.66 51.04 2.81<br />

b-2-8.5-0:30 0.31 0.001008 0.8 168.31 56.37 2.99<br />

b-2-8.5-1:30 0.39 0.001004 0.8 168.31 56.37 2.99<br />

b-2-8.5-2:30 0.48 0.001004 0.8 168.31 56.37 2.99<br />

b-3-13.3-0:00 0.25 0.001190 4.2 469.04 90.43 5.19<br />

b-3-13.3-0:30 0.52 0.001015 8.2 877.69 107.58 8.16<br />

b-3-13.3-1:30 0.45 0.001008 8 838.70 105.94 7.92<br />

b-3-13.3-2:30 0.19 0.001007 8 838.70 105.94 7.92<br />

210


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Table S6 Freila 2 hydraulic data<br />

Run - MP - flow length [m]- τ Г ω<br />

sampling time [min:sec] [Pa] [N m -1 ] [W m -2 ]<br />

ω U ω eff<br />

Re τ - τ cr<br />

[m s -1 ] [W m -1 ] [ ] [Pa]<br />

a-1-4-0:00 33.35 33.34 15.67 0.05 476.44 14840.13 31.28<br />

a-1-4-0:30 22.91 16.98 12.47 0.05 477.57 12643.85 20.84<br />

a-1-4-1:30 23.05 13.87 15.97 0.07 767.17 16276.69 20.98<br />

a-1-4-2:30 17.19 9.82 14.47 0.08 866.60 14676.93 15.12<br />

a-2-8.5-0:00 16.45 8.32 3.45 0.01 308.68 5619.61 14.38<br />

a-2-8.5-0:30 16.97 8.87 4.14 0.01 287.89 6976.46 14.90<br />

a-2-8.5-1:30 15.93 7.87 4.96 0.02 696.51 8412.52 13.86<br />

a-2-8.5-2:30 17.36 9.07 6.58 0.02 577.49 11212.34 15.29<br />

a-3-13.3-0:00 66.33 59.98 21.22 0.04 809.25 15329.03 64.26<br />

a-3-13.3-0:30 103.48 111.32 34.97 0.04 1095.81 26859.14 101.41<br />

a-3-13.3-1:30 100.22 106.18 37.48 0.05 1235.77 29202.70 98.15<br />

a-3-13.3-2:30 100.17 106.12 41.06 0.05 1417.16 32125.42 98.10<br />

b-1-4-0:00 10.91 5.27 6.11 0.06 377.59 5905.24 8.84<br />

b-1-4-0:30 18.44 10.64 12.30 0.07 628.24 12503.04 16.37<br />

b-1-4-1:30 17.17 9.81 13.73 0.08 801.27 14039.07 15.10<br />

b-1-4-2:30 17.17 9.81 14.92 0.09 908.07 15284.39 15.10<br />

b-2-8.5-0:00 16.50 8.42 4.29 0.02 352.57 6954.96 14.43<br />

b-2-8.5-0:30 17.39 9.80 5.34 0.02 308.90 9112.39 15.32<br />

b-2-8.5-1:30 17.38 9.80 6.84 0.02 447.03 11708.97 15.31<br />

b-2-8.5-2:30 17.38 9.80 8.25 0.03 592.94 14138.38 15.31<br />

b-3-13.3-0:00 67.10 60.68 16.78 0.03 568.65 11156.44 65.03<br />

b-3-13.3-0:30 103.27 111.10 53.49 0.07 2072.79 41729.93 101.20<br />

b-3-13.3-1:30 100.12 106.07 45.47 0.06 1651.29 35706.89 98.05<br />

b-3-13.3-2:30 100.11 106.06 19.44 0.03 461.72 15281.78 98.04<br />

211


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Table S7 Freila 3 erosion data<br />

Run - MP - flow length [m]-<br />

sampling time [min:sec]<br />

Sediment<br />

Concentration<br />

[g L -1 ]<br />

Detachment<br />

rate<br />

[kg s -1 m -2 ]<br />

Transport<br />

rate<br />

[kg s-1]<br />

Sample<br />

density<br />

[g cm -3 ]<br />

Slope<br />

[°]<br />

Transport<br />

capacity<br />

[kg s-1]<br />

a-1-2.9-0:00 45.0 0.2958 0.379845863 1.03 4.7 0.08011<br />

a-1-2.9-0:30 3.0 0.0174 0.017769517 1.00 4.7 0.02040<br />

a-1-2.9-1:30 1.6 0.0135 0.014059196 1.00 4.7 0.02558<br />

a-1-2.9-2:30 0.8 0.0081 0.008537001 1.00 4.7 0.02670<br />

a-2-11-0:00 43.2 0.1264 0.653849085 1.03 15.1 0.77604<br />

a-2-11-0:30 7.8 0.0298 0.189057846 1.00 15.1 1.85097<br />

a-2-11-1:30 4.3 0.0173 0.117480752 1.00 15.1 2.37381<br />

a-2-11-2:30 3.7 0.0161 0.120276643 1.00 15.1 3.15647<br />

a-3-13.8-0:00 56.3 0.0619 0.414999121 1.04 4.4 0.07706<br />

a-3-13.8-0:30 11.1 0.0144 0.087141739 1.01 4.4 0.02722<br />

a-3-13.8-1:30 5.5 0.0080 0.048446063 1.00 4.4 0.02707<br />

a-3-13.8-2:30 3.9 0.0068 0.043053290 1.00 4.4 0.04278<br />

b-1-2.9-0:00 3.5 0.0257 0.036744778 1.00 4.7 0.14627<br />

b-1-2.9-0:30 1.4 0.0178 0.023345189 1.00 4.7 0.08610<br />

b-1-2.9-1:30 0.4 0.0054 0.006318732 1.00 4.7 0.03960<br />

b-1-2.9-2:30 0.0 0.0000 0.000000000 1.00 4.7 0.03846<br />

b-2-11-0:00 10.2 0.0422 0.267669423 1.01 15.1 1.85510<br />

b-2-11-0:30 3.3 0.0122 0.080111940 1.00 15.1 2.12315<br />

b-2-11-1:30 1.9 0.0055 0.038627457 1.00 15.1 2.57786<br />

b-2-11-2:30 1.2 0.0026 0.019920464 1.00 15.1 3.37596<br />

b-3-13.8-0:00 26.1 0.0421 0.302333459 1.02 4.4 0.12407<br />

b-3-13.8-0:30 6.1 0.0121 0.076629194 1.00 4.4 0.04286<br />

b-3-13.8-1:30 2.5 0.0034 0.021755547 1.00 4.4 0.04272<br />

b-3-13.8-2:30 2.2 0.0002 0.001559311 1.00 4.4 0.04271<br />

212


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Table S8 Freila 3 runoff data<br />

Run - MP - flow length [m]-<br />

sampling time [min:sec]<br />

Flow<br />

velocity<br />

[m s -1 ]<br />

Dynamic<br />

viscosity<br />

[kg s -1 m -1 ]<br />

Water<br />

depth<br />

[cm]<br />

Flow cross<br />

section<br />

[cm²]<br />

Wetted<br />

Perimeter<br />

[cm]<br />

Hydraulic<br />

radius<br />

[cm]<br />

a-1-2.9-0:00 0.44 0.001225 7.6 191.68 45.06 4.25<br />

a-1-2.9-0:30 0.65 0.001015 4.3 89.43 35.77 2.50<br />

a-1-2.9-1:30 0.90 0.001008 4.8 100.07 36.54 2.74<br />

a-1-2.9-2:30 1.00 0.001004 4.9 103.52 37.15 2.79<br />

a-2-11-0:00 0.61 0.001216 10.00 248.00 47.02 5.27<br />

a-2-11-0:30 0.55 0.001039 15.00 436.45 57.68 7.57<br />

a-2-11-1:30 0.52 0.001022 17.00 517.84 61.88 8.37<br />

a-2-11-2:30 0.51 0.001018 20.00 638.72 68.08 9.38<br />

a-3-13.8-0:00 0.35 0.001281 7.00 210.67 48.55 4.34<br />

a-3-13.8-0:30 0.62 0.001055 5.00 128.02 44.00 2.91<br />

a-3-13.8-1:30 0.69 0.001027 5.00 128.02 44.00 2.91<br />

a-3-13.8-2:30 0.68 0.001020 6.00 160.05 45.78 3.50<br />

b-1-2.9-0:00 0.38 0.001018 9.7 276.10 50.23 5.50<br />

b-1-2.9-0:30 0.82 0.001007 5.9 204.80 46.02 4.45<br />

b-1-2.9-1:30 1.33 0.001002 5.7 134.37 41.18 3.26<br />

b-1-2.9-2:30 1.55 0.001000 5.6 123.89 38.41 3.23<br />

b-2-11-0:00 0.6 0.001051 15.00 436.45 57.68 7.57<br />

b-2-11-0:30 0.51 0.001016 16.00 479.77 59.92 8.01<br />

b-2-11-1:30 0.38 0.001009 18.00 550.13 63.54 8.66<br />

b-2-11-2:30 0.25 0.001006 21.00 674.83 69.96 9.65<br />

b-3-13.8-0:00 0.42 0.001130 9.00 276.30 52.05 5.31<br />

b-3-13.8-0:30 0.78 0.001031 6.00 160.05 45.78 3.50<br />

b-3-13.8-1:30 0.54 0.001013 6.00 160.05 45.78 3.50<br />

b-3-13.8-2:30 0.04 0.001011 6.00 160.05 45.78 3.50<br />

213


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Table S9 Freila 3 hydraulic data<br />

Run - MP - flow length [m]- τ Г ω<br />

sampling time [min:sec] [Pa] [N m -1 ] [W m -2 ]<br />

ω U ω eff<br />

Re τ - τ cr<br />

[m s -1 ] [W m -1 ] [ ] [Pa]<br />

a-1-2.9-0:00 35.16 15.84 15.47 0.04 339.16 15709.50 33.19<br />

a-1-2.9-0:30 20.13 7.20 13.13 0.05 387.49 16085.57 18.16<br />

a-1-2.9-1:30 22.03 8.05 19.74 0.07 664.13 24368.99 20.06<br />

a-1-2.9-2:30 22.41 8.33 22.33 0.08 788.02 27665.64 20.44<br />

a-2-11-0:00 138.45 65.10 84.45 0.16 3602.41 27174.42 136.48<br />

a-2-11-0:30 194.31 112.09 107.33 0.14 3938.82 40414.91 192.34<br />

a-2-11-1:30 214.45 132.70 111.95 0.14 3859.87 42875.13 212.48<br />

a-2-11-2:30 240.30 163.61 123.07 0.13 3992.08 47287.94 238.33<br />

a-3-13.8-0:00 33.81 16.42 11.83 0.03 239.70 12271.29 31.84<br />

a-3-13.8-0:30 22.05 9.70 13.56 0.05 368.01 17075.36 20.08<br />

a-3-13.8-1:30 21.97 9.67 15.20 0.05 436.48 19653.60 20.00<br />

a-3-13.8-2:30 26.38 12.08 17.97 0.05 497.22 23420.27 24.41<br />

b-1-2.9-0:00 44.28 22.24 16.83 0.03 326.93 20572.31 42.31<br />

b-1-2.9-0:30 35.80 16.48 29.31 0.07 1046.85 36207.56 33.83<br />

b-1-2.9-1:30 26.23 10.80 34.84 0.11 1388.43 43265.86 24.26<br />

b-1-2.9-2:30 25.93 9.96 40.30 0.13 1748.03 50139.66 23.96<br />

b-2-11-0:00 194.60 112.25 116.76 0.16 4468.98 43466.94 192.63<br />

b-2-11-0:30 205.03 122.86 105.19 0.13 3660.45 40501.41 203.06<br />

b-2-11-1:30 221.51 140.75 83.15 0.10 2378.54 32237.17 219.54<br />

b-2-11-2:30 246.70 172.58 61.60 0.07 1368.60 23964.67 244.73<br />

b-3-13.8-0:00 40.61 21.14 17.05 0.03 350.70 20049.07 38.64<br />

b-3-13.8-0:30 26.41 12.09 20.69 0.06 614.06 26675.66 24.44<br />

b-3-13.8-1:30 26.35 12.06 14.15 0.04 347.22 18563.41 24.38<br />

b-3-13.8-2:30 26.35 12.06 1.17 0.00 8.22 1533.53 24.38<br />

214


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Table S10 Negratin erosion data<br />

Run - MP - flow length [m]-<br />

sampling time [min:sec]<br />

Sediment<br />

Concentration<br />

[g L -1 ]<br />

Detachment<br />

rate<br />

[kg s -1 m -2 ]<br />

Transport<br />

rate<br />

[kg s-1]<br />

Sample<br />

density<br />

[g cm -3 ]<br />

Slope<br />

[°]<br />

Transport<br />

capacity<br />

[kg s-1]<br />

a-1-3.2-0:00 67.3 0.3509 0.378972891 1.04 3.2 0.02284<br />

a-1-3.2-0:30 55.7 0.3594 0.417851939 1.03 3.2 0.02926<br />

a-1-3.2-1:30 37.0 0.2853 0.331751054 1.02 3.2 0.02877<br />

a-1-3.2-2:30 35.2 0.3353 0.411477439 1.02 3.2 0.03395<br />

a-2-5.1-0:00 128.8 0.1191 0.083361100 1.08 7.7 0.00238<br />

a-2-5.1-0:30 102.6 0.2553 0.258820005 1.06 7.7 0.01888<br />

a-2-5.1-1:30 60.6 0.1840 0.183346450 1.04 7.7 0.01584<br />

a-2-5.1-2:30 41.3 0.1941 0.219213489 1.03 7.7 0.02768<br />

a-3-11.5-0:00 170.4 0.0370 0.039566213 1.11 1.7 0.00012<br />

a-3-11.5-0:30 155.3 0.0391 0.039947665 1.10 1.7 0.00009<br />

a-3-11.5-1:30 95.0 0.0333 0.035583905 1.06 1.7 0.00011<br />

a-3-11.5-2:30 81.0 0.0364 0.040214682 1.05 1.7 0.00016<br />

b-1-3.2-0:00 82.1 0.5253 0.567410676 1.05 3.2 0.02314<br />

b-1-3.2-0:30 35.9 0.2925 0.340134613 1.02 3.2 0.02874<br />

b-1-3.2-1:30 24.7 0.2555 0.297103809 1.02 3.2 0.02845<br />

b-1-3.2-2:30 12.4 0.2164 0.265524812 1.01 3.2 0.03324<br />

b-2-5.1-0:00 128.2 0.1551 0.112506225 1.08 7.7 0.00302<br />

b-2-5.1-0:30 48.3 0.1641 0.171502072 1.03 7.7 0.01942<br />

b-2-5.1-1:30 28.8 0.1376 0.147576853 1.02 7.7 0.02288<br />

b-2-5.1-2:30 28.3 0.1823 0.219256736 1.02 7.7 0.03540<br />

b-3-11.5-0:00 168.5 0.0358 0.038265903 1.10 1.7 0.00012<br />

b-3-11.5-0:30 79.8 0.0234 0.025837099 1.05 1.7 0.00016<br />

b-3-11.5-1:30 52.8 0.0137 0.013972868 1.03 1.7 0.00008<br />

b-3-11.5-2:30 51.4 0.0153 0.016371852 1.03 1.7 0.00011<br />

215


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Table S11 Negratin runoff data<br />

Run - MP - flow length [m]-<br />

sampling time [min:sec]<br />

Flow<br />

velocity<br />

[m s -1 ]<br />

Dynamic<br />

viscosity<br />

[kg s -1 m -1 ]<br />

Water<br />

depth<br />

[cm]<br />

Flow cross<br />

section<br />

[cm²]<br />

Wetted<br />

Perimeter<br />

[cm]<br />

Hydraulic<br />

radius<br />

[cm]<br />

a-1-3.2-0:00 0.53 0.001336 0.5 106.28 33.75 3.15<br />

a-1-3.2-0:30 0.59 0.001278 2 126.85 36.33 3.49<br />

a-1-3.2-1:30 0.71 0.001185 2 126.85 36.33 3.49<br />

a-1-3.2-2:30 0.82 0.001176 3 143.13 38.35 3.73<br />

a-2-5.1-0:00 0.64 0.001644 0.4 10.11 13.72 0.74<br />

a-2-5.1-0:30 0.75 0.001513 2.1 33.86 19.88 1.70<br />

a-2-5.1-1:30 0.96 0.001303 2 31.49 19.53 1.61<br />

a-2-5.1-2:30 1.18 0.001207 2.6 44.94 22.14 2.03<br />

a-3-11.5-0:00 0.46 0.001852 0.40 5.05 9.30 0.54<br />

a-3-11.5-0:30 0.59 0.001776 0.30 4.33 8.89 0.49<br />

a-3-11.5-1:30 0.74 0.001475 0.40 5.05 9.30 0.54<br />

a-3-11.5-2:30 0.81 0.001405 0.50 6.12 9.61 0.64<br />

b-1-3.2-0:00 0.65 0.001411 0.5 106.28 33.75 3.15<br />

b-1-3.2-0:30 0.75 0.001180 2 126.85 36.33 3.49<br />

b-1-3.2-1:30 0.95 0.001123 2 126.85 36.33 3.49<br />

b-1-3.2-2:30 1.50 0.001062 3 143.13 38.35 3.73<br />

b-2-5.1-0:00 0.76 0.001641 0.5 11.54 14.22 0.81<br />

b-2-5.1-0:30 0.99 0.001242 2.3 36.00 20.49 1.76<br />

b-2-5.1-1:30 1.29 0.001144 2.5 39.73 21.03 1.89<br />

b-2-5.1-2:30 1.46 0.001141 3 53.06 23.58 2.25<br />

b-3-11.5-0:00 0.45 0.001843 0.40 5.05 9.30 0.54<br />

b-3-11.5-0:30 0.53 0.001399 0.50 6.12 9.61 0.64<br />

b-3-11.5-1:30 0.61 0.001264 0.30 4.33 8.89 0.49<br />

b-3-11.5-2:30 0.63 0.001257 0.40 5.05 9.30 0.54<br />

216


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Table S12 Negratin hydraulic data<br />

Run - MP - flow length [m]- τ Г ω<br />

sampling time [min:sec] [Pa] [N m -1 ] [W m -2 ]<br />

ω U ω eff<br />

Re τ - τ cr<br />

[m s -1 ] [W m -1 ] [ ] [Pa]<br />

a-1-3.2-0:00 17.96 6.06 9.52 0.03 1004.67 13009.48 15.04<br />

a-1-3.2-0:30 19.78 7.19 11.70 0.03 543.29 16715.29 16.85<br />

a-1-3.2-1:30 19.56 7.11 13.84 0.04 698.89 21333.97 16.63<br />

a-1-3.2-2:30 20.89 8.01 17.08 0.05 730.82 26518.86 17.96<br />

a-2-5.1-0:00 10.46 1.44 6.70 0.09 687.70 3099.18 7.54<br />

a-2-5.1-0:30 23.82 4.74 17.75 0.10 982.41 8925.34 20.90<br />

a-2-5.1-1:30 21.99 4.30 21.12 0.13 1317.11 12329.08 19.06<br />

a-2-5.1-2:30 27.36 6.06 32.28 0.16 2090.11 20354.33 24.44<br />

a-3-11.5-0:00 1.75 0.16 0.80 0.01 28.58 1490.41 -1.18<br />

a-3-11.5-0:30 1.56 0.14 0.92 0.02 42.67 1786.38 -1.37<br />

a-3-11.5-1:30 1.67 0.16 1.24 0.02 54.86 2890.54 -1.26<br />

a-3-11.5-2:30 1.95 0.19 1.58 0.02 67.92 3863.99 -0.98<br />

b-1-3.2-0:00 18.12 6.12 11.78 0.04 1382.74 15249.04 15.20<br />

b-1-3.2-0:30 19.55 7.10 14.58 0.04 755.90 22575.90 16.62<br />

b-1-3.2-1:30 19.41 7.05 18.44 0.05 1074.74 29978.88 16.48<br />

b-1-3.2-2:30 20.60 7.90 30.90 0.08 1778.67 53133.31 17.67<br />

b-2-5.1-0:00 11.52 1.64 8.76 0.10 886.05 4058.91 8.59<br />

b-2-5.1-0:30 23.79 4.87 23.45 0.13 1404.17 14369.29 20.86<br />

b-2-5.1-1:30 25.28 5.32 32.62 0.17 2178.77 21693.14 22.35<br />

b-2-5.1-2:30 30.10 7.10 44.00 0.20 3022.79 29329.14 27.17<br />

b-3-11.5-0:00 1.74 0.16 0.79 0.01 27.61 1464.08 -1.18<br />

b-3-11.5-0:30 1.94 0.19 1.03 0.02 35.72 2528.39 -0.98<br />

b-3-11.5-1:30 1.47 0.13 0.90 0.02 40.73 2434.20 -1.46<br />

b-3-11.5-2:30 1.63 0.15 1.03 0.02 41.37 2810.22 -1.30<br />

217


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Table S13 Salada erosion data<br />

Run - MP - flow length [m]-<br />

sampling time [min:sec]<br />

Sediment<br />

Concentration<br />

[g L -1 ]<br />

Detachment<br />

rate<br />

[kg s -1 m -2 ]<br />

Transport<br />

rate<br />

[kg s-1]<br />

Sample<br />

density<br />

[g cm -3 ]<br />

Slope<br />

[°]<br />

Transport<br />

capacity<br />

[kg s-1]<br />

a-1-2.3-0:00 65.7 0.2744 0.163638470 1.04 14.9 0.07299<br />

a-1-2.3-0:30 21.9 0.0840 0.053634448 1.01 14.9 0.12230<br />

a-1-2.3-1:30 11.4 0.0767 0.044566797 1.01 14.9 0.05316<br />

a-1-2.3-2:30 15.4 0.4565 0.309609576 1.01 14.9 0.18004<br />

a-2-4.7-0:00 30.8 0.0812 0.094584461 1.02 24.5 0.14384<br />

a-2-4.7-0:30 18.3 0.0529 0.070550578 1.01 24.5 0.29789<br />

a-2-4.7-1:30 40.7 0.2666 0.437682105 1.03 24.5 0.79787<br />

a-2-4.7-2:30 11.4 0.0980 0.117807544 1.01 24.5 0.17785<br />

a-3-4.7-0:00 113.4 0.3238 0.315564275 1.07 24.5 0.18940<br />

a-3-4.7-0:30 42.2 0.1193 0.126580737 1.03 24.5 0.28476<br />

a-3-4.7-1:30 25.0 0.1295 0.149733833 1.02 24.5 0.43935<br />

a-3-4.7-2:30 14.7 0.2187 0.282982246 1.01 24.5 0.70410<br />

b-1-2.3-0:00 36.8 0.1280 0.076308115 1.02 14.9 0.07110<br />

b-1-2.3-0:30 8.1 0.0812 0.091376454 1.01 14.9 0.25308<br />

b-1-2.3-1:30 5.8 0.0783 0.081788592 1.00 14.9 0.23377<br />

b-1-2.3-2:30 7.0 0.1136 0.143162068 1.00 14.9 0.36293<br />

b-2-4.7-0:00 42.9 0.2433 0.404671007 1.03 24.5 0.84780<br />

b-2-4.7-0:30 19.7 0.1253 0.189952323 1.01 24.5 0.54642<br />

b-2-4.7-1:30 9.6 0.0684 0.084120916 1.01 24.5 0.20340<br />

b-2-4.7-2:30 10.9 0.1511 0.215740177 1.01 24.5 0.41807<br />

b-3-4.7-0:00 39.0 0.1429 0.151710680 1.02 24.5 0.28391<br />

b-3-4.7-0:30 9.5 0.0551 0.063738761 1.01 24.5 0.43310<br />

b-3-4.7-1:30 7.8 0.0905 0.117090916 1.00 24.5 0.69957<br />

b-3-4.7-2:30 6.6 0.1257 0.174190550 1.00 24.5 0.90918<br />

218


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Table S14 Salada runoff data<br />

Run - MP - flow length<br />

[m]-<br />

sampling time [min:sec]<br />

Flow<br />

velocity<br />

[m s -1 ]<br />

Dynamic<br />

viscosity<br />

[kg s -1 m -1 ]<br />

Water<br />

depth<br />

[cm]<br />

Flow cross<br />

section<br />

[cm²]<br />

Wetted<br />

Perimeter<br />

[cm]<br />

Hydraulic<br />

radius<br />

[cm]<br />

a-1-2.3-0:00 0.48 0.001328 2.5 51.91 25.92 2.00<br />

a-1-2.3-0:30 0.35 0.001110 3.2 69.41 27.75 2.50<br />

a-1-2.3-1:30 0.86 0.001057 2.2 45.47 25.27 1.80<br />

a-1-2.3-2:30 2.33 0.001077 4.1 86.29 29.49 2.93<br />

a-2-4.7-0:00 0.62 0.001154 4.7 49.50 24.79 2.00<br />

a-2-4.7-0:30 0.51 0.001091 3.7 76.10 28.35 2.68<br />

a-2-4.7-1:30 0.78 0.001204 7.7 137.92 34.93 3.95<br />

a-2-4.7-2:30 1.84 0.001057 4.1 56.01 25.58 2.19<br />

a-3-4.7-0:00 0.62 0.001567 5 44.87 20.73 2.16<br />

a-3-4.7-0:30 0.51 0.001211 6 59.01 22.58 2.61<br />

a-3-4.7-1:30 0.78 0.001125 7 76.97 24.61 3.13<br />

a-3-4.7-2:30 1.84 0.001073 8 104.38 27.53 3.79<br />

b-1-2.3-0:00 0.40 0.001184 2.5 51.91 25.92 2.00<br />

b-1-2.3-0:30 0.68 0.001041 6.3 164.46 48.91 3.36<br />

b-1-2.3-1:30 0.95 0.001029 5.9 148.03 45.41 3.26<br />

b-1-2.3-2:30 0.96 0.001035 7.5 212.97 54.81 3.89<br />

b-2-4.7-0:00 0.66 0.001214 8.0 143.01 35.38 4.04<br />

b-2-4.7-0:30 0.87 0.001098 6.6 110.25 32.24 3.42<br />

b-2-4.7-1:30 1.45 0.001048 4.2 60.45 26.15 2.31<br />

b-2-4.7-2:30 2.12 0.001054 5.7 93.66 30.39 3.08<br />

b-3-4.7-0:00 0.66 0.001195 6 59.01 22.58 2.61<br />

b-3-4.7-0:30 0.87 0.001047 7 76.97 24.61 3.13<br />

b-3-4.7-1:30 1.45 0.001039 8 104.38 27.53 3.79<br />

b-3-4.7-2:30 2.12 0.001033 9 124.17 29.48 4.21<br />

219


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Table S15 Salada hydraulic data<br />

Run - MP - flow length [m]- τ Г ω<br />

sampling time [min:sec] [Pa] [N m -1 ] [W m -2 ]<br />

ω U ω eff<br />

Re τ - τ cr<br />

[m s -1 ] [W m -1 ] [ ] [Pa]<br />

a-1-2.3-0:00 52.57 13.63 25.24 0.12 1482.75 7531.25 49.37<br />

a-1-2.3-0:30 63.95 17.75 22.53 0.09 1061.25 8050.56 60.75<br />

a-1-2.3-1:30 45.71 11.55 39.35 0.22 3144.11 14760.25 42.50<br />

a-1-2.3-2:30 74.52 21.98 173.86 0.60 19280.91 64004.64 71.32<br />

a-2-4.7-0:00 82.79 20.52 51.33 0.26 2823.76 10932.84 79.59<br />

a-2-4.7-0:30 110.44 31.31 56.09 0.21 3783.55 12635.30 107.23<br />

a-2-4.7-1:30 164.68 57.53 128.40 0.32 8038.43 26224.38 161.48<br />

a-2-4.7-2:30 89.70 22.95 165.43 0.76 17895.79 38472.08 86.50<br />

a-3-4.7-0:00 94.27 19.54 58.44 0.26 3292.10 9166.87 91.06<br />

a-3-4.7-0:30 109.10 24.64 55.42 0.21 2691.66 11247.11 105.90<br />

a-3-4.7-1:30 129.22 31.80 100.75 0.32 5953.66 22018.10 126.02<br />

a-3-4.7-2:30 155.65 42.85 287.09 0.76 26199.37 65737.82 152.45<br />

b-1-2.3-0:00 51.66 13.39 20.67 0.10 1098.76 6920.70 48.46<br />

b-1-2.3-0:30 85.25 41.70 58.17 0.18 2802.12 22158.58 82.04<br />

b-1-2.3-1:30 82.54 37.48 78.04 0.24 4548.92 30059.69 79.33<br />

b-1-2.3-2:30 98.44 53.96 94.08 0.25 5131.14 36030.09 95.24<br />

b-2-4.7-0:00 168.82 59.74 111.42 0.27 6334.72 22554.07 165.62<br />

b-2-4.7-0:30 140.82 45.40 123.21 0.36 8374.20 27572.20 137.61<br />

b-2-4.7-1:30 94.60 24.74 136.77 0.60 13237.98 32076.24 91.40<br />

b-2-4.7-2:30 126.24 38.36 267.99 0.88 29621.25 62486.08 123.04<br />

b-3-4.7-0:00 108.88 24.59 71.86 0.27 3974.97 14785.43 105.68<br />

b-3-4.7-0:30 127.99 31.50 111.99 0.36 6977.44 26284.46 124.79<br />

b-3-4.7-1:30 154.99 42.67 224.07 0.60 18065.59 53022.45 151.78<br />

b-3-4.7-2:30 172.07 50.72 365.28 0.88 34762.59 86918.88 168.87<br />

220


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Table S16 Belerda erosion data<br />

Run - MP - flow length [m]-<br />

sampling time [min:sec]<br />

Sediment<br />

Concentration<br />

[g L -1 ]<br />

Detachment<br />

rate<br />

[kg s -1 m -2 ]<br />

Transport<br />

rate<br />

[kg s-1]<br />

Sample<br />

density<br />

[g cm -3 ]<br />

Slope<br />

[°]<br />

Transport<br />

capacity<br />

[kg s-1]<br />

a-1-6-0:00 111.1 0.6584 0.463968090 1.07 11.7 0.01464<br />

a-1-6-0:30 250.8 0.9576 0.674872393 1.15 11.7 0.01644<br />

a-1-6-1:30 288.4 0.5173 0.364531283 1.18 11.7 0.01694<br />

a-1-6-2:30 176.1 0.4415 0.387089673 1.11 11.7 0.04693<br />

a-2-13-0:00 243.8 0.1124 0.242154919 1.15 15.1 0.03603<br />

a-2-13-0:30 359.6 0.1026 0.084365950 1.22 15.1 0.00926<br />

a-2-13-1:30 234.8 0.0617 0.038866537 1.14 15.1 0.00230<br />

a-2-13-2:30 88.0 0.0533 0.041378347 1.05 15.1 0.00577<br />

a-3-17-0:00 341.2 0.2734 0.876936459 1.21 15.4 0.15169<br />

a-3-17-0:30 422.3 0.6314 2.058803483 1.26 15.4 0.16862<br />

a-3-17-1:30 62.6 0.0887 0.297213079 1.04 15.4 0.13513<br />

a-3-17-2:30 383.4 0.0896 0.296150064 1.24 15.4 0.16984<br />

b-1-6-0:00 93.0 0.1522 0.118800584 1.06 11.7 0.02513<br />

b-1-6-0:30 212.6 0.3207 0.226026124 1.13 11.7 0.01594<br />

b-1-6-1:30 16.5 0.0211 0.012073263 1.01 11.7 0.00496<br />

b-1-6-2:30 13.5 0.0389 0.030357874 1.01 11.7 0.02340<br />

b-2-13-0:00 312.2 0.2399 0.516754731 1.19 15.1 0.03803<br />

b-2-13-0:30 97.3 0.0387 0.038785872 1.06 15.1 0.01574<br />

b-2-13-1:30 421.1 0.2325 0.171978951 1.26 15.1 0.00611<br />

b-2-13-2:30 61.8 0.0860 0.086292771 1.04 15.1 0.01525<br />

b-3-17-0:00 31.7 0.0240 0.078397448 1.02 15.4 0.12267<br />

b-3-17-0:30 141.3 0.1565 0.524266143 1.09 15.4 0.14471<br />

b-3-17-1:30 125.3 0.2185 0.722405305 1.08 15.4 0.13812<br />

b-3-17-2:30 109.2 0.2682 0.909694568 1.07 15.4 0.14510<br />

221


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Table S17 Belerda runoff data<br />

Run - MP - flow length<br />

[m]-<br />

sampling time [min:sec]<br />

Flow<br />

velocity<br />

[m s -1 ]<br />

Dynamic<br />

viscosity<br />

[kg s -1 m -1 ]<br />

Water<br />

depth<br />

[cm]<br />

Flow cross<br />

section<br />

[cm²]<br />

Wetted<br />

Perimeter<br />

[cm]<br />

Hydraulic<br />

radius<br />

[cm]<br />

a-1-6-0:00 2.94 0.001556 1.0 14.20 11.75 1.21<br />

a-1-6-0:30 1.89 0.002254 1.0 14.20 11.75 1.21<br />

a-1-6-1:30 0.89 0.002442 1.0 14.20 11.75 1.21<br />

a-1-6-2:30 0.80 0.001881 2.0 27.55 14.61 1.89<br />

a-2-13-0:00 0.42 0.002219 3.0 23.65 16.57 1.43<br />

a-2-13-0:30 0.46 0.002798 0.7 5.06 6.33 0.80<br />

a-2-13-1:30 0.72 0.002174 0.2 2.31 4.85 0.48<br />

a-2-13-2:30 1.09 0.001440 0.6 4.32 5.98 0.72<br />

a-3-17-0:00 0.56 0.002706 0.2 45.89 18.87 2.43<br />

a-3-17-0:30 1.03 0.003111 0.3 47.50 19.18 2.48<br />

a-3-17-1:30 0.95 0.001313 0.5 50.15 19.70 2.55<br />

a-3-17-2:30 0.16 0.002917 0.4 48.85 19.45 2.51<br />

b-1-6-0:00 0.65 0.001465 1.5 19.65 13.01 1.51<br />

b-1-6-0:30 0.75 0.002063 1.0 14.20 11.75 1.21<br />

b-1-6-1:30 0.94 0.001083 0.2 7.72 9.52 0.81<br />

b-1-6-2:30 1.14 0.001068 1.5 19.65 13.01 1.51<br />

b-2-13-0:00 0.70 0.002561 3.0 23.65 16.57 1.43<br />

b-2-13-0:30 0.48 0.001486 1.0 8.31 7.72 1.08<br />

b-2-13-1:30 1.08 0.003105 0.5 3.78 5.69 0.66<br />

b-2-13-2:30 1.68 0.001309 1.0 8.31 7.72 1.08<br />

b-3-17-0:00 0.52 0.001159 0.3 47.50 19.18 2.48<br />

b-3-17-0:30 0.74 0.001706 0.5 50.15 19.70 2.55<br />

b-3-17-1:30 1.18 0.001627 0.4 48.85 19.45 2.51<br />

b-3-17-2:30 1.62 0.001546 0.6 51.41 19.95 2.58<br />

222


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Table S18 Belerda hydraulic data<br />

Run - MP - flow length [m]- τ Г ω<br />

sampling time [min:sec] [Pa] [N m -1 ] [W m -2 ]<br />

ω U ω eff<br />

Re τ - τ cr<br />

[m s -1 ] [W m -1 ] [ ] [Pa]<br />

a-1-6-0:00 25.71 3.02 75.58 0.60 14155.26 24423.50 22.94<br />

a-1-6-0:30 27.78 3.26 52.62 0.38 8224.56 11735.83 25.01<br />

a-1-6-1:30 28.34 3.33 25.22 0.18 2728.19 5190.96 25.57<br />

a-1-6-2:30 41.58 6.08 33.17 0.16 2592.39 8864.96 38.81<br />

a-2-13-0:00 41.96 6.95 17.62 0.11 766.31 3107.72 39.19<br />

a-2-13-0:30 24.97 1.58 11.58 0.12 1076.71 1619.26 22.19<br />

a-2-13-1:30 13.93 0.67 10.00 0.19 1991.25 1799.58 11.16<br />

a-2-13-2:30 19.47 1.16 21.19 0.28 2954.62 5758.07 16.70<br />

a-3-17-0:00 76.70 14.47 42.95 0.15 17731.56 6092.17 73.92<br />

a-3-17-0:30 81.32 15.60 83.46 0.27 36654.75 10296.35 78.55<br />

a-3-17-1:30 68.88 13.57 65.17 0.25 17991.78 19049.17 66.11<br />

a-3-17-2:30 80.93 15.74 12.80 0.04 1816.39 1683.83 78.16<br />

b-1-6-0:00 31.78 4.13 20.66 0.13 1543.50 7087.10 29.01<br />

b-1-6-0:30 27.21 3.20 20.36 0.15 1979.84 4961.43 24.44<br />

b-1-6-1:30 16.31 1.55 15.41 0.19 3811.70 7155.59 13.54<br />

b-1-6-2:30 30.30 3.94 34.59 0.23 3345.39 16287.61 27.53<br />

b-2-13-0:00 43.50 7.21 30.45 0.18 1740.33 4652.71 40.73<br />

b-2-13-0:30 29.16 2.25 14.00 0.13 1128.08 3684.88 26.39<br />

b-2-13-1:30 21.39 1.22 23.11 0.28 3798.42 2911.66 18.62<br />

b-2-13-2:30 28.56 2.20 47.98 0.44 7159.25 14340.51 25.79<br />

b-3-17-0:00 65.77 12.62 34.20 0.14 9616.37 11330.97 63.00<br />

b-3-17-0:30 72.10 14.20 53.35 0.20 13326.36 12001.08 69.32<br />

b-3-17-1:30 70.51 13.71 83.19 0.31 30113.84 19632.98 67.74<br />

b-3-17-2:30 71.65 14.29 116.06 0.43 37864.55 28811.68 68.87<br />

223


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Kapitel 7<br />

Ries et al. (subm. 2011): Sheep and goat erosion – experimental geomorphology as an<br />

approach for the quantification of underestimated processes. Zeitschrift für<br />

Geomorphologie.<br />

224


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Sheep and goat erosion – experimental geomorphology as an approach for the<br />

quantification of underestimated processes<br />

Ries, J.B., Andres, K., <strong>Wirtz</strong>, S., Tumbrink, J., Wilms, T., Peter, K. D., Burczyk, M., Butzen,<br />

V. & Seeger, M.<br />

Dep. of Physical Geography, Trier University, Germany<br />

with 10 figures and 2 tables<br />

Abstract<br />

The grazing of goats and sheep is regarded as an important factor for soil degradation in semiarid<br />

landscapes. Nevertheless, hardly any data can be found in literature. In the presented<br />

study, the process dynamics of material disaggregation and translocation directly caused by<br />

trampling animals is quantified by means of experimental methods on test plots. Gerlach<br />

troughs are installed in order to quantify material mobilisation in different directions. The<br />

slope angle as well as the running speed of the animals is varied. Additionally, the amount of<br />

material that is loosened by the hooves of the goats is measured.<br />

The translocation rates are surprisingly high and slope as well as running speed turn out to be<br />

important influencing factors. In downslope direction, each goat can translocate between 0.6 g<br />

and 6.5 on each square meter, depending on slope. The maximum translocation rate in<br />

movement direction reaches 4.5 g m -2 per goat for fast running and 1.3 g m -2 per goat for slow<br />

motion. Additionally, each goat can loosen 14 g of soil material per square meter; this<br />

material can easily be removed by wind or water. Experiments with marked rock fragments<br />

on slopes of 4° and 11° showed that a flock of 45 goats translocates rock fragments (ø 3 cm)<br />

by a mean distance of 8.8 cm. While rock movement occurs in all directions, most often they<br />

are kicked forward-downslope. Net mean downslope translocation rates vary between 1.5 cm<br />

and 6.6 cm corresponding with slope and the total number of rock fragments moved.<br />

Keywords<br />

erosion, sheep and goats, experimental geomorphology, Mediterranean landscapes<br />

225


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

1 Introduction<br />

About 40 % of the earth's land-surface is used for grazing and 80 % of this grazing area is<br />

situated in landscapes with semi-arid and arid climates (e.g. BRANSON et al. 1981,<br />

MONFREDA et al. 2008, 2009, RAMANKUTTY et al. 2008). Over-grazing is considered the<br />

main cause of 35 % of world-wide soil degradation (OLDEMANN et al. 1991). On the<br />

African continent, 49 % of the soil surface is already degraded due to grazing and in Australia<br />

the affected area is 81 % (EVANS 1998, OLDEMANN et al. 1991). Accordingly, grazinginduced<br />

soil degradation is one of the most widespread land-degradation phenomena.<br />

Numerous field surveys in literature focus on the degradation of vegetation due to browsing<br />

damage by animals (e.g. ALADOS et al. 2004, ANDRESEN et al. 1990, MWENDERA &<br />

MOHAMED SALEEM 1997b, MWENDERA et al. 1997, NOY-MEIR 1975, 1978, 1995,<br />

PEARSON et al. 1990, PEGAU 1970, RIES et al. 2003, VALENTIN 1985, WANG et al.<br />

2002). There are also many studies about the formation of specific grazing-induced vegetation<br />

patterns (e.g. BARBIER et al. 2006, BOONKORKUEA et al. 2010, BORGOGNO et al. 2009,<br />

DEBLAUWE et al. 2008, GALLE et al. 2001, GOLODETS & BOEKEN 2006,<br />

HILLERISLAMBERS et al. 2001, MONTANA 1992, TONGWAY et al. 2001, TONGWAY<br />

& LUDWIG 2001, WORRALL 1959). Vegetation density and distributions patterns play a<br />

crucial role in soil erosion processes. According to basic literature (ELWELL & STOCKING<br />

1976), a vegetation cover of 30-40 % is sufficient to protect soil surfaces against degradation.<br />

SEUFFERT (1999) confirms these values. Other studies, which deal specifically with<br />

sediment yields in low open matorral shrublands in Mediterranean mountain regions, assume<br />

that a considerably higher degree of vegetation cover is needed to contain erosive processes<br />

(MOLINILLO et al. 1997, RIES 2005). They ascribe this to the patchy pattern of tussocks.<br />

Spots densely covered with Stipa tenacissima have high infiltration rates and low runoff<br />

coefficients, while the bare areas around them experience sheet flow and rill erosion. Grazing<br />

in these areas leads to a higher contrast between tussock patches and bare soil (CERDÀ<br />

1997).<br />

ROSTANGO et al. (1991) found a remarkable distinction in terms of infiltration rates and<br />

chemical and physical soil properties on comparing shrub mounds and inter-mound areas for<br />

soils in northeastern Patagonia. Similar results were presented by DUNKERLY & BROWN<br />

(1995) for banded vegetation covers in patterned chenopod shrublands in New South Wales,<br />

Australia. They divided their areas of investigation into source sections, characterized by high<br />

runoff rates, and sink sections with vital importance for the redistribution of water.<br />

226


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Furthermore, the article points out the vulnerability of these shrublands against grazinginduced<br />

erosion.<br />

Basically, two main theses on the impacts of grazing can be found in literature. The studies<br />

with a more ecological or botanical focus regard grazing animals as ‘ecosystem engineers’<br />

and emphasize the increased biodiversity of grazing areas with different biospheres for other<br />

biota, as compared with non-grazing areas (e.g. JONES et al. 1994, SCHAICH 2007, STA<strong>VI</strong><br />

et al. 2009). The more soil scientific, geomorphologic and hydrologic studies point out the<br />

physical compaction phenomena and, accordingly, the reduced infiltration rates of the soils<br />

(MULHOLLAND & FULLEN 1991, PARIENTE 2002, PIETOLA et al. 2005, PROFFITT et<br />

al. 1995, STA<strong>VI</strong> et al. 2008a, b, STA<strong>VI</strong> et al. 2009, STEWART & CAMERON 1992,<br />

TABOADA & LAVADO 1993, MWENDERA & SALEEM 1997a).<br />

In contrast to that, the immediate influence of the trampling of sheep and goat hooves on the<br />

soil surface are overlooked. The destruction of crusts and the loosening and mobilization of<br />

soil material for further transport through wind and water are possible effects. Correlating<br />

geomorphologic rates are also unknown, so far. Only few studies suggest that the impact of<br />

animal trampling is significant and should not be neglected any more (e.g. UNGAR et al.<br />

2009, RIES 2005, 2010).<br />

The issue of rock fragment translocation is treated in even fewer publications. GOVERS &<br />

POESEN (1998) investigate the translocation of rock fragments by sheep and goat trampling<br />

in a study area with steep, debris-covered slopes in Turkey, and discovered significant rock<br />

fragment movements. Particularly, the long-term effects of animal trampling on debriscovered<br />

slopes should be subject to further investigation (GOVERS & POESEN 1998).<br />

UNGAR et al. (2009) determine considerable rock fragment movement regardless of their<br />

size on grazed areas as opposed to the reference area without grazing influence in the Negev<br />

desert. The vegetation cover is regarded to be the most important factor influencing rockfragment<br />

movement by animal trampling. On existing trails more rock fragments are moved<br />

than on densely vegetation-covered areas. NYSSEN et al. (2006) emphasize rock fall and<br />

material movement as being the most important geomorphologic process induced by animal<br />

trampling in Ethiopia. On grazed test-areas, up to 95 % of the rock fragments on the soil<br />

surface were moved in the course of 20 months.<br />

Only a small number of experimental studies focus on the influence of grazing on vegetation<br />

cover and geomorphodynamics (RIES et al. 2003, ELDRIDGE 1998, MWENDERA &<br />

MOHAMED SALEEM 1997a). RIES et al. (2003) observe that sheep and goat trails in dense<br />

Matorral with intensive grazing can be regarded as the most active surfaces due to the<br />

227


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

concentrated trampling impact. They show a tendency to spatial progression, even when shrub<br />

density increases. By means of rainfall simulations, runoff coefficients of up to 98 % and<br />

sediment discharges of up to 328 g m -2 could be measured on trails using 40 mm h -1 rainfall<br />

intensity and a simulated rainfall duration of 30 minutes (RIES 2010). These values are very<br />

high compared to runoff coefficients and sediment discharges on other surfaces in the same<br />

study area (runoff up to 76 %, sediment discharge up to 68 g m -2 ). ELDRIDGE (1998)<br />

determine a higher erodibility, decreasing infiltration rates, increasing overland flow<br />

generation, a decrease in surface crust-cover and an increase of loose, bare soil, also using<br />

rainfall simulations on areas with crusted soil surfaces and simulated animal trampling<br />

(artificial sheep hooves).<br />

Nevertheless, livestock-induced erosion rates are hitherto practically unknown. The<br />

fundamental significance of trampling erosion is, however, undisputed. There are few<br />

quantitative field studies that acquire their measurements from animals in their natural<br />

surroundings. This lack of research is likely to derive from the complexity of the processes<br />

involved, being far from thoroughly understood, and non-existing adequate methods. In<br />

addition, handling live animals and using them as research objects can be difficult and should<br />

not be underestimated. Particularly, data are missing on: the impact of hooves on dry coherent<br />

crusted surfaces; the influence of trampling on the translocation of fine material in downslope<br />

as well as in and against the animals’ moving direction; the influence of slope and moving<br />

speed on those translocation rates; and the impact of trampling on the translocation of rock<br />

fragments.<br />

In this context there is need for comparative studies for the rock translocation experiments in<br />

the Taurus conducted by GOVERS & POESEN (1998). Even though this phenomenon is<br />

ubiquitous in nearly all semi-arid and arid regions, the impact of trampling livestock on stonecovered<br />

surfaces is clearly understudied. Above all, quantitative studies on the distance and<br />

direction of translocated rock fragments and fine material are missing. Finally, the impact of<br />

trampling livestock on fluvial erosive forms is of particular interest since little is known about<br />

how detached material is eroded when flushing occurs, for instance in rills or gullies.<br />

Following those points, the main objectives of this study are<br />

1. to develop methods for the quantification of translocation rates of soil-material and<br />

rock fragments by sheep and goat trampling on in-situ sites,<br />

2. to present first results concerning ranges of moved rock fragments and soil material<br />

under these specific experimental setups and<br />

3. to quantify the effect of animal trampling on sediment availability in rills.<br />

228


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

In this way, the directly measured translocation or loosening rates by goat trampling can be<br />

evaluated and compared with the rates of other erosion processes. In addition, the discussion<br />

on the methodological development is intended to be encouraged.<br />

2 Material and Methods<br />

2.1 Study Area<br />

The study area Freila is located in the Hoya de Baza sedimentary basin (see figure 1). The<br />

bedrock is composed of Pliocene marls and sandstones on which calcareous Lithosols in<br />

loamy sand are developed. The semi-arid climate is characterized by an average annual<br />

temperature of about 14.2 °C and an annual precipitation of 368 mm, with a high inter-annual<br />

variability. The vegetation is dominated by low shrub-land of Thymus vulgaris and Stipa<br />

tenacissima grassland. The land-cover on the southern lake-side of the Negratin reservoir is<br />

dominated by abandoned cereal fields, which are extensively grazed by flocks of sheep and<br />

goats. Agricultural land-use comprises mainly cereal dry-farming, almond plantations and<br />

olive orchards (SEEGER 2007). Generally, non-vegetated areas are crusted and rock<br />

fragments are embedded in the crusts. Only on recently used trails, the crusts are destroyed<br />

and rock fragments are covering the substrate.<br />

Figure 1: Localization of the study area in High Andalusia<br />

229


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

2.2 Methods:<br />

All experiments were run on test plots with similar footprints (see fig. 2), installed in contour<br />

line direction, but carried out on sites with differing surface properties and slopes. The test<br />

plots are representative for areas used as pasture.<br />

15 goats were guided over each test plot several times. Meanwhile, translocated material was<br />

collected and the transport direction recorded. Surface type, slope and running speed were<br />

varied in accordance with the observed way of natural flock movement patterns. The used<br />

‘Malaga-goats’ (DOPPELBAUER 2002) and sheep weighed between 45 and 50 kg each. In<br />

detail, we conducted the following experiments:<br />

Experiment 1: Translocation quantity in down-slope direction<br />

Experiment 1 was conducted on slopes of 8.5°, 12° and 20°. Vegetation cover was less than 5<br />

% on the plots. Plot size was 2 m x 1 m (see fig. 2a). A Gerlach trough with the length of 2 m<br />

was implemented as a sediment trap at the down-slope end of the plot. 15 goats were guided<br />

over the plot 4 times in a row (= 60 goats). Afterwards, the sediment traps were emptied and<br />

weighed. This process was repeated 10 times in order to simulate the impact of the passage of<br />

a flock of 600 animals. In experiment 1c (slope 20°) we increased the running speed of the<br />

animals after the 5 th run (= 5 x 60 goats), in the other 2 versions, the running speed was kept<br />

constant over all runs.<br />

Experiment 2: Translocation quantity parallel to contour lines<br />

Experiment 2 was conducted on a 8.5° steep sheep trail. The test plot was covered with 13 %<br />

vegetation and 18 % rock fragments. The footprint of the experiment was a combination of<br />

two separate plots (each 2 x 1 m) resulting in a total size of 4 x 1 m. The Gerlach troughs<br />

(each 1 m in length) were placed at a right angle to the moving direction of the goats at the<br />

end of those plots (see fig. 2b). Again, the material collected in the troughs was weighed after<br />

each run of 60 goats. In experiment 2a, the running speed of the goats was increased after the<br />

5 th run, in experiment 2b the speed was kept constant.<br />

Experiment 3: Translocation direction and distance of rock fragments<br />

Experiments 3 a, b and c were conducted on test plots of 2 x 0.6 m (1.2 m 2 ) on 4° (3 a & 3 b)<br />

and 11° slopes (3 c). On all plots 120, colored and numbered rock fragments (ø ca. 3 cm)<br />

230


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

were spread in a 10 x 10 cm pattern (see fig. 2c). After the passing of 3 x 15 goats (= 45<br />

goats), the translocation vector (distance and angle) of each rock fragment was measured.<br />

The original position of each fragment was marked by a nail that was driven into the ground.<br />

The resulting values were categorized into 4 segments according to the measured angles:<br />

- (1) 0-89° forward-downslope<br />

- (2) 90-179° backward-downslope<br />

- (3) 180-269° backward-upslope<br />

- (4) 270-359° forward-upslope<br />

Since the bulk of our rock fragments landed in segment 1, this segment was further divided,<br />

allowing for a more precise interpretation:<br />

- (1a) 0-44° forward-forward-downslope<br />

- (1b) 45-89° downslope-forward<br />

Later on, we applied the law of sines to calculate the vertical up- and downslope translocation<br />

distances of all moved rock fragments.<br />

Experiment 4: Loosening quantity<br />

The test plot of the 4 th experiment had a rectangular footprint of 4 m x 0.5 m (see fig. 2d) and<br />

was characterized by a flat, crusted surface. It was covered sparsely by dead vegetation (about<br />

5 %); rock fragments were not present. A flock of 15 goats was guided over the test plot 4<br />

times (= 60 goats) before it was swept with a regular hand broom. The loosened fine material<br />

was collected and weighed. This procedure was repeated 10 times in order to simulate the<br />

gradual loosening process of a soil crust caused by a flock of 600 animals.<br />

231


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Figure 2: Setup of the experiments. 2a: Translocation quantity in down-slope direction; 2b:<br />

Translocation quantity in running direction; 2c: Translocation direction and distance of rock<br />

fragments, black dots are the rock fragments; 2d: Loosening quantity<br />

Experiment 5: Flow detachment of the loosened material<br />

The rill experiments consist of two runs. First, the rill was tested under field conditions; in a<br />

second run, about 15 minutes later, the same rill was tested under wet conditions. Using a<br />

motor driven pump, a constant discharge of 250 L min -1 is maintained during 4 minutes,<br />

resulting in a total water inflow of 1000 L. The mobilization of material by the water jet at the<br />

inflow is prevented by covering the substrate with a doormat. For gathering intermediate data<br />

on suspended sediment transport, three adequate measuring points (MP) were selected. Small<br />

knick-points in the rill have proven to be the best sampling points because at these points, the<br />

bottles do not have to be pressed down to the bottom of the rill. Here, four water samples<br />

were taken. The first was taken immediately on the waterfront’s arrival at the sampling points.<br />

The second one was taken after 30 seconds, the third after 90 seconds and the fourth 150<br />

seconds after the arrival of the water. The sediment concentration was determined by filtration<br />

of the samples in the laboratory. Flow velocity was also measured for each meter flow length<br />

using a stopwatch (WIRTZ et al. 2010, 2012).<br />

232


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

The same rill was tested with and without goat trampling. The rill had an average slope of<br />

8.8° and a maximum slope of 15.2°. The tested flow length is 16 m. Gravel content was 30%,<br />

antecedent soil moisture 3 %, vegetation cover about 20 % and rock fragment cover was 60<br />

%. The maximum depth of the rill was 10 cm and the maximum width about 40 cm. Grain<br />

density was 2.69 Mg m -3 and (dry) bulk density was 1.44 Mg m -3 .<br />

Methodology for analysis:<br />

We used various methodological approaches to describe and illustrate our results:<br />

In order to describe the results, different methodological approaches were used:<br />

In experiment 1, the translocation quantity [g] is given. The translocated material is<br />

removed from a defined test site area and collected in Gerlach Troughs. This quantity refers to<br />

an area of 2 m² and 60 goats per run; the cumulative values represent 600 goats. As it is a<br />

downslope mobilization, we can additionally state the translocation flux, i.e. the<br />

translocation quantity per unit length in downslope direction. As the downslope length is 1 m,<br />

the translocation flux is equal to the translocation quantity. The translocation quantity is<br />

measured over the test plot length of 2 m, which leads to the following equation: translocation<br />

flux [g m -1 ] = translocation quantity / 2. The translocation rate can be calculated from the<br />

translocation quantity. It refers to 1 m² and one goat: translocation rate = translocation<br />

quantity / 120.<br />

In experiment 2 we use the translocation quantity and the translocation rate. The translocation<br />

flux is not used because the values are not the result of a downslope movement. In this<br />

experiment it is noteworthy that each trough collects the material from an area of 2 m², so<br />

cumulated material from both troughs refers to 4 m². Additionally, in experiment 2, we<br />

discuss a trail erosion rate. This value is given in Mg per ha and 600 goats, calculated from<br />

the translocation rate in g per m² and goat. This calculation is representative in this case,<br />

because on trails, the goats run similarily fast and similarily close as in our experiments.<br />

In experiment 3, we calculate a net mean downslope translocation rate for each individual<br />

rock fragment per experiment: Net mean downslope translocation rate = (downslope quantity<br />

x downslope mean distance) - (upslope quantity x upslope mean distance) / all spread out rock<br />

fragments.<br />

In experiment 4, the loosening quantity was measured by brushing off and collecting the<br />

loosened material from the flat test plot. The quantity is given in [g], referring to an area of 2<br />

m² and 60 goats in each run. The loosening rate refers to 1 m² and 1 goat, leading to:<br />

loosening rate = loosening quantity / 120.<br />

233


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

In experiment 5, the detachment rate [Kg s -1 m -2 ] and the transport rate [Kg s -1 ] are<br />

presented. They are calculated as follows:<br />

transport rate = sediment concentration * flow velocity * flow cross section<br />

detachment rate = transport rate / (flow length * wetted perimeter)<br />

The sediment concentration is given in [g L -1 ], flow velocity in [m s -1 ], the flow cross section<br />

in [m 2 ], flow length and wetted perimeter in [m].<br />

3 Results and interpretation<br />

3.1 Translocation quantities in downslope direction<br />

Figure 3: Translocation quantities in down-slope direction<br />

The first runs are not taken into account in order to eliminate possible errors due to the<br />

installation of the Gerlach troughs.<br />

On the 8.5° slope plot, rock fragments are translocated only in the 4 th run, in all other runs<br />

only fine material is affected by the trampling. The quantities are similar, they range between<br />

75 and 41 g, only run 4 with 282 g and run 8 with only 4 g are out of line. The quantities are<br />

presented in figure 3a. In total, 83 g rock fragments (11 %) and 662 g fine material was<br />

translocated (figure 3d). The translocation flux is equal to translocation quantity.<br />

234


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

The course of the translocation quantity of fine material and rock fragments in down-slope<br />

direction on a 12° slope is presented in figure 3b. From run 2 to 8, the quantity of fine<br />

material is higher than the quantity of rock fragments, the percentage of rock fragments on<br />

total translocated material is between 20 % and 25 % (in runs 4, 6 and 7, it is 0 %). Run 9 and<br />

10 show an increase in rock fragment quantity, in run 9 the percentage increases to 70 % and<br />

in run 10 even to 81 %. The translocation rate of fine material is similarily high in runs 2-5,<br />

while runs 6-10 show lower fine material amounts except for run 8 with a fine material<br />

amount in the range of the first group. The total translocated fine material is 1245 g, rock<br />

fragments with 496 g cause 28 % of total sampled material (figure 3d).<br />

On a 20° slope (figure 3c), the mass of translocated material increases up to a mean value of<br />

788 g (runs 2-5). From run 6 to 10 the running speed was increased, which can be regarded as<br />

typical for the more difficult slope segments. The amount of eroded material for the slow runs<br />

2 to 5 ranges between 475 g and 961 g. In the faster runs 6 to 10, values between 1 573 g and<br />

1 032 g are measured, which is equivalent to a medial increase of 66 %. The percentage of<br />

rock fragments of the total mass of eroded material increases from 10 % in the slow runs to 21<br />

% in the fast runs. In total 8096 g were translocated, 16 % of which were rock fragments. Due<br />

to comparability reasons, the 10 th runs of the experiments 12° and 20° were not considered in<br />

figure 3d.<br />

Interpretation:<br />

Experiment 1a (8.5° slope) only shows low material translocation rates. The given slope is too<br />

flat to transport the material to the troughs. The material seems to be loosened and after<br />

several intermediate storages, the material reaches the troughs in run 4.<br />

Also in experiment 1b (12° slope), only a low amount of material is translocated with 2.26 kg<br />

m -2 per goat. The fluctuating amounts of moved soil material can be explained by the fact that<br />

the material either reaches the Gerlach troughs directly, or it is temporally stored in the still<br />

existent micro-relief of ridges and furrows. The rock fragments are only mobilized after about<br />

480 goats have passed. The threshold was crossed this late because the rocks were embedded<br />

into the soil and re gradually detached by the trampling.<br />

Due to the increase of slope to 20° in experiment 1c, the translocation rates in the slow runs<br />

were 3 times higher and reach values of up to 6.6 g m -2 per goat. Again, the translocation rate<br />

increased with higher running speed up to 10 g m -2 per goat and thereby reaches the same<br />

range as the loosening rate in experiment 4. Due to the steep slope, the material reached the<br />

Gerlach trough directly, without intermediate storage. The amount of transported rock<br />

235


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

fragments increased considerably in later runs, because a certain number of crossing animals<br />

is necessary to mobilize the rock fragments from their embedded position. This could already<br />

be observed on the 12° slope. The increase in running speed also promotes the movement of<br />

rock fragments.<br />

3.2 Translocation quantities parallel to contour lines<br />

236


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Figure 4: Translocation quantities parallel to contour lines<br />

Again, the first run is not taken into account in order to avoid errors resulting from the<br />

installation of the Gerlach troughs.<br />

237


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Amounts of material which were moved on existing trails parallel to contour lines range<br />

between the values measured on the abandoned fields in downslope direction with 8.5° and<br />

12° slopes (experiment 1).<br />

In a first experiment, we changed the running speed after the 5 th run and measured fine<br />

material as well rock fragment translocation. In a second experiment with constant slow<br />

running speed, we wanted to ensure to have measured the influence of different speed in the<br />

first experiment, not the influence of a longer trampling time and higher animal quantity.<br />

In the first experiment (figure 4a) rock fragments were not moved in the first 5 slow runs.<br />

From run 6 onwards, the running speed of the animals was increased so that the goats were<br />

now really 'running' along the trail, just in a way that can be regarded as typical for faster<br />

movements of flocks. This faster movement lead to an increase of the translocation quantity<br />

by the factor five, but with a temporal lag, because it was recorded only in the 7 th run.<br />

Afterwards this trend decreased. Now, in addition to the higher fine material amount, rock<br />

fragments were moved in runs 7 to 10. Their proportion rose and fell from 31 % in the 7 th run<br />

over 35 % and 29 % down to 9 % in the last run. Altogether 4134 g of material was moved in<br />

this experiment, 1237 g or 30 % of this material were rock fragments (see figure 4a).<br />

In the second experiment, hardly any rock fragments were present on the plot. In mean, 145 g<br />

material was moved in each run, this is similar to the 155 g measured in the slow runs in<br />

experiment 2a.<br />

Interpretation:<br />

The similar mean values of the 4 slow runs in experiment 2a and the 9 runs in experiment 2b<br />

show that the differences between the slow and the fast runs in experiment 2a are caused by<br />

the different running speeds and not by trampling time and animal quantity.<br />

In the trail experiment, the measured translocation rates parallel to contour lines reached<br />

values around 1.3 g m -2 per goat for the slow runs which is comparable to the values measured<br />

on the abandoned field. At the higher running speed, as it is more typical for the trails, the<br />

translocation rate was five times higher with 5.9 g m -2 per goat. In the following experiment,<br />

it is clearly to see that the speed influences the translocation rate and not the longer trampling<br />

time and the higher number of animals. That means the running speed is an important<br />

parameter for the quantity of translocated material. The movement of rock fragments again<br />

can only be observed after the passing of about 300 goats.<br />

3.3 Translocation direction and distance of rock fragments<br />

238


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

239


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Figure 5: Translocation direction and distance of rock fragments. Length of the arrows<br />

represents the mean translocation distance, width of the arrows the quantity of the rock<br />

fragments in each segment.<br />

Figure 5 shows the previously introduced segments, the proportion of rock fragments that was<br />

translocated into each of them as well as the mean distance the fragments were moved. While<br />

the number of rock fragments is illustrated by the width of the arrows, their length represents<br />

the mean movement distance.<br />

Experiments 3a and b were both conducted on test plots with a slope of 4°. In experiment 3a,<br />

94% of the rock fragments were moved with a mean distance of 7,4 cm. Of these fragments<br />

44% were relocated forward downslope, 22% backward downslope, 12% backward upslope<br />

and 18% forward upslope. The mean transportation distances for the segments 1-4 were 8.4<br />

cm, 7.0 cm, 8.9 cm, and 5.3 cm, respectively. The values in the sub segments 1a and 1b<br />

resemble each other closely both in input proportion (22%, 22%) and movement distance (8.4<br />

cm, 8.3 cm).<br />

In experiment 3b, 71% of all spread out rock fragments were relocated with a mean distance<br />

of 5.0 cm. 56% of them were moved forward downslope, 14% backward downslope, 13%<br />

backward upslope and 16% forward upslope. The corresponding mean translocation distances<br />

were 7.2 cm, 4.4 cm, 3.6 cm, and 4.6 cm. Again, the sub segments 1a and b have equal input<br />

proportions (28%, 28%), but this time differ in terms of mean movement distance (8.7 cm, 5.6<br />

cm).<br />

Experiment 3c was conducted on a steeper slope (11°) than experiments 3a and b. Here, 90%<br />

of all spread out rock fragments were moved with a mean transportation distance of 14.1 cm.<br />

The dominating translocation direction was forward downslope with an input proportion of<br />

71% of all moved fragments. The input in segment 2 is 15%, in segment 3 it is lowest with<br />

4% and in segment 4 it is 10%. The mean transportation distances for the 4 segments are 16.8<br />

cm, 9.3 cm, 11.2 cm and 19.0 cm. The input in sub segment 1a is more than twice as high as<br />

in 1b (49%, 22%). The mean transportation distance, however, is higher in sub segment 1b<br />

(14.1 cm, 22.9 cm).<br />

240


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Figure 6: Net up- and downslope translocation<br />

In contrast to figures 5a-c, figures 6a and b do not distinguish between the four segments, but<br />

juxtapose the vertical up- and downslope translocation for the three experiments. In<br />

experiment 3a, 61% of all spread out rock fragments were moved downslope, as opposed to<br />

the 23% that were pushed upslope. Also, the mean translocation distance was higher<br />

downslope (5.4 cm) than upslope (3.2 cm). In experiment 3b, 49% of all rock fragments were<br />

translocated downslope with a mean distance of 3.7 cm. 16% were pushed upslope with a<br />

mean distance of 2.4 cm. The most contrasting results were generated in experiment 3c, where<br />

66% of all rock fragments were moved downslope with a mean distance of 11.4 cm, while<br />

11% were transported upslope with a mean distance of 7.9 cm.<br />

Interpretation:<br />

241


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Experiments 3a, b and c show that on average 85 % of all marked rock fragments were<br />

mobilized after the passing of 45 goats. As figure 5 illustrates, the main direction of<br />

translocation throughout all 3 experiments is forward downslope (segment 1) with an average<br />

proportion of 57 % of all moved rock fragments. Also true for all 3 experiments is the fact<br />

that significantly more rock fragments were pushed forward than backward (on average 62 %<br />

to 28 %). Hence, the goats’ moving direction and slope direction seem to correlate with the<br />

translocation direction of our rock fragments. As expected, the results of experiment 3c show<br />

that gravity plays an increasingly important role once slope is steeper. Here, 86 % of all<br />

moved rock fragments were relocated downslope, compared with 70 % in experiments 3a and<br />

b combined. Furthermore, the input proportion in the forward directions (segments 1 and 4) is<br />

heavily skewed in favor of the forward downslope compartment with a ratio of 71 % to 10 %,<br />

compared with 44 % to 18 % and 56 % to 16 % in experiments 3a and b, respectively. This<br />

indicates a change in the trajectory of the kicked rock fragments due to steeper slope. Rock<br />

fragments that would come to a halt in segment 4 on a flatter slope, now land in segment 1.<br />

This hypothesis is also nicely supported by the analysis of the sub segments 1a and b. Since<br />

the input quantities are similar for both sub segments in experiments 3a and b, they do not<br />

allow for a judgment of whether the goats’ moving direction or slope has a higher impact on<br />

the rock fragments relocation. In experiment 3c, however, where slope is much steeper (11°),<br />

the input in segment 1a is more than twice as high as in 1b. Likely, this result can be ascribed<br />

to an additional input of rock fragments in 1a, which on flatter slopes would have remained in<br />

segment 4.<br />

Finally, the distribution of rock fragments across all 4 segments, i.e. all directions, is<br />

noteworthy. The hooves’ impact cannot be easily translated into a forward-downslope kicking<br />

direction of rock fragments. While this is true for the lion’s share of all measured rock<br />

fragments, segments 2, 3 and 4 constantly contained at least some of the total quantity.<br />

Interestingly, the highest mean translocation distances were not always recorded in the<br />

forward-downslope direction. In fact, this was only the case in experiment 3b. In experiments<br />

3a and 3c, the upslope segments 3 and 4, respectively, show the highest values. However,<br />

since few rock fragments were moved into these segments and the mean is calculated, outliers<br />

can easily affect the result.<br />

As expected, the mean translocation distance increases with slope. Whereas it is 6.2 cm for<br />

experiments 3a and b put together, the impact of gravity clearly surfaces in experiment 3c,<br />

where it is 14.1 cm. On analyzing the results of the sub segments 1a (forward-forwarddownslope)<br />

and 1b (downslope-forward) in all 3 experiments, the impetus of the factor slope<br />

242


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

becomes evident once more. On the relatively flat surface in experiments 3a and b, the mean<br />

translocation distances of the sub segments is nearly identical in experiment 3a (8.4 cm, 8.3<br />

cm) and a little higher in sub segment 1a in experiment 3b (8.7 cm, 5.6 cm). On the steeper<br />

11° slope in experiment 3c, the mean translocation distance in the downslope-forward sub<br />

segment (1b) is significantly higher (14.1 cm, 22.9 cm). The difference in experiment 3c can<br />

probably be explained by the high overlap of the driving forces of initial kicking impulse and<br />

gravity in sub segment 1b. While we assume that 1a contains some rock fragments that were<br />

initially kicked in upslope direction, fragments that ended in 1b have likely been kicked in<br />

downslope direction by the goats right away. In that case, impulse and gravity would meet<br />

best in 1b, leading to the highest mean translocation distance (22.9 cm) here.<br />

The mean net downslope translocation rates for experiments 3a, b and c are 2.6 cm, 1.5 cm<br />

and 6.6 cm. Again, slope appears to be a dominating factor and explains why the highest<br />

distance is recorded in experiment 3c. The low net mean downslope translocation rate of 1.5<br />

cm in experiment 3b is caused by the relatively high proportion of unmoved stones.<br />

3.4 Loosening quantity<br />

Figure 7: Loosening rate of fine material on a highly crusted plain surface<br />

243


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Figure 7 shows that the loosening quantity increased from 305 g in the 1 st to 1538 g in the 3 rd<br />

run and stayed at this level until the total of 600 goats passed after the 10 th run. After the 5 th<br />

run, a slight decrease can be observed, but this is probably not significant.<br />

Interpretation:<br />

The increase in the amount of loosened material from run 1 to 3 can be traced back to a stepby-step<br />

cracking of the sedimentation crust on the whole test plot. After 180 goats, the crust is<br />

completely destroyed and the soil material is loosened with 1.5 kg 2 m -2 per 60 goats. The<br />

slightly decreasing trend can be explained by lower availability of detachable crust material<br />

and by increasing compaction. The measured loosening rate can be evaluated as high<br />

regarding the value of 12.5 g m - ² per goat.<br />

3.5 Flow detachment of loosened material<br />

Figure 8: Sediment concentrations, transport and detachment rates of the rill experiment with<br />

and without goat trampling (MP = measuring point)<br />

244


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

All three erosion parameters showed higher values in the experiment with goat trampling, but<br />

the difference is most obvious in the dry run. In the second (wet) run, the values were equal or<br />

at least similar to the values of the experiments without trampling. Transport and detachment<br />

rate were about 5 times higher in the dry run with goat trampling as in the runs without<br />

trampling. Sediment concentration in the dry run with trampling was about 2.5 higher<br />

compared to the experiment without trampling (see figure 8).<br />

Interpretation:<br />

The rill experiment clearly shows the short-term influence of goat trampling. Livestock’s<br />

hooves’ impact prepares the soil surface and a large quantity of loose material is made ready<br />

for transport (see experiment 4). Crucially, this material is removed very fast; the second run<br />

with trampling still shows nearly the same values as the runs without preceding trampling.<br />

The sediment concentrations at the waterfront were clearly higher in the experiment with<br />

trampling, whereas the other samples show similar concentrations. Accordingly, trampling<br />

provides a large quantity of loose material but its impact on mobilization is only detectable for<br />

a short time.<br />

3.6 Statistical analysis<br />

In table 1, the statistical values of the experiments 1, 2 and 4 are summarized. In experiment 1<br />

and 2, the first run was not used in the calculations to avoid mistakes caused by the<br />

installation of the troughs. The difference between mean and standard deviation increases<br />

with slope, meaning high quantities as well as low quantities of substrate can be translocated.<br />

The quantity of the mean increases with slope but the values become less stable. This is also<br />

detectable in figure 8. Regarding the values for the movement parallel to contour lines, the<br />

differences between mean and standard deviation for trough 1 and 2 in each version are very<br />

similar, they are in all three cases between 52 and 58 %. But in all three cases, the first trough<br />

shows clearly higher differences than the second trough. Because of the low n-values, we do<br />

not want to over-interpret the values and conclude that the high values for trough 1 are caused<br />

by the jumping-in and -out of the animals at the beginning of the test plot. Once they are<br />

inside the fenced course, the movements are smoother. Figure 9 shows that again, the fast<br />

running speed can cause very high values as well as low values, the mean values are clearly<br />

higher than under slow movement but the values are less stable. The high difference between<br />

245


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

mean and standard deviation in the loosening experiment can be explained by the slow<br />

increase at the beginning due to breaking the crust.<br />

Table 1: Statistical values of the different experiments. St.Dev. = standard deviation, T1 =<br />

Gerlach Trough A, T2 = Gerlach Trough B<br />

Mean [g] Median St.Dev. IMean-St.Dev.I % from mean<br />

[g] [g]<br />

Slope 8.5° 93.1 66.1 104.9 11.8 12.7<br />

Slope 12° 250.5 231.9 132.3 118.2 47.2<br />

Slope 20° 1047.0 1032.3 299.1 747.9 71.4<br />

Trail A slow T1 92.3 76.5 34.3 58.0 62.8<br />

Trail A slow T2 62.6 67.0 32.9 29.7 47.4<br />

Trail A fast T1 374.2 402.0 120.1 254.1 67.9<br />

Trail A fast T2 328.8 233.0 169.0 159.8 48.6<br />

Trail B slow T1 105.1 116.0 25.5 79.6 75.7<br />

Trail B slow T2 39.6 31.1 28.2 11.4 28.8<br />

Loosening 1322.2 1431.5 394.8 927.4 70.1<br />

Figure 9: Boxplots of the different experimental setups. T1 = Gerlach Trough A, T2 =<br />

Gerlach Trough B<br />

246


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

4 Discussion<br />

The calculated translocation rates increase with slope and with the running speed of the<br />

animals, thus conforming with theoretical expectations. All values are considered to be<br />

explicitly higher than a tolerable soil erosion rate for completely mechanized agriculture. The<br />

influence of slope has also been shown by OOSTWOUD WIJDENES et al. (2001). The<br />

measured transport rates on a 50 % slope were twice as high as on a 20 % slope. The total<br />

amount of loosened material by trampling of 6.6 Kg m -2 on the test plots has also to be<br />

evaluated as very high. The provided loose material can easily be eroded by wind and water<br />

erosion processes (cf. our rill experiments).<br />

The trails in the study area are distributed very irregularly (see figure 10). The total trail<br />

density is 247 m ha -1 ; 13 985 m ha -1 in the trail zone and 68 m ha -1 outside this zone.<br />

Assuming a mean trail width of 30 cm, the total trail area takes up 705 m²; 545 m² in the trail<br />

zone and 160 m² outside. Regarding the proportions of the total trail area, the values are 0.7 %<br />

for the whole study area, 3 % in the trail zone and 0.2 % outside the trail zone (see table 2).<br />

Table 2: Trails length and area sizes of different landscape units in the test area<br />

Value<br />

% of total area<br />

Trail length in Trail zone [m] 1818 -<br />

Trail length outside Trail zone[m] 532 -<br />

Total trail length [m] 2350 -<br />

Trail zone [m²] 17300 18<br />

Abandoned fields [m²] 49000 51<br />

Steep slopes [m²] 27830 29<br />

Untreated zones [m²] 1300 2<br />

Total area study site [m²] 95430 -<br />

On the trails, sheep run similarly fast and close as simulated in our experiments, so we can use<br />

our measured values to calculate trail erosion rates: Parallel to contour lines (in and against<br />

running direction), each slow moving goat can translocate 1.3 g m -2 , and, if speed is<br />

increased, the value can reach 5.9 g m -2 . For an extrapolation, we assume an average value of<br />

3.6 g m -2 per goat and the same slopes as in our experiments, since the plots were<br />

representative for trails in the study site. The total area is 9.5 ha, the trail area is 705 m²,<br />

meaning the trail erosion rate reaches 160 Kg ha -1 per 600 goats. This value should clearly be<br />

247


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

higher in trail zone (1.7 ha, trail area 545 m²), where the trail erosion rate reaches 692 Kg ha -1<br />

per 600 goats. Outside the trail zone, the value is 44 Kg ha -1 per 600 goats.<br />

Figure 10: Trail distribution in the study area<br />

The translocation of rock fragments results in an almost complete workup of the rock<br />

fragment cover. According to CERDÀ (2001) rock fragments retard ponding and surface<br />

runoff and give greater steady state infiltration rates and smaller interrill runoff discharges,<br />

sediment concentrations and interrill erosion rates. He performed rainfall simulations with 55<br />

mm h -1 for 120 min, in the middle of which he removed the rock fragment cover. The steadystate<br />

infiltration rate diminished from 44.5 to 27.5 mm h -1 , the runoff coefficient increased by<br />

the factor 3, sediment concentration by the factor 33 and the erosion rate by the factor 39.<br />

That means the rock fragments are an important erosion protection and if they are removed by<br />

animal trampling, the substrate below becomes vulnerable against other erosion processes.<br />

248


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

We measured distance and direction of the rock fragment translocation in experiment 3. The<br />

mean translocation distance of the 3 experiments only reaches 8.8 cm, thus being much lower<br />

than the mean displacement distance of 38 cm on Lesvos (Greece) that was determined by<br />

OOSTWOUD WIJDENES et al. (2001). However, this rather short distance is a consequence<br />

of the considerably flatter slopes of 4° - 11° in the present study vs. 2° - 24° in OOSTWOUD<br />

WIJDENES et al. (2001). It is important to notice that OOSTWOUD WIJDENES et al.<br />

(2001) measured the 38 cm after one year, the grazing peak is between April and June. It is<br />

not reconstructible how far the rock fragments were displaced after a certain number of sheep.<br />

Additionally, they corrected the negative distances (upslope movement) by adding the largest<br />

negative distance plus one centimeter to all negative distances. This transformation resulted in<br />

a slightly overestimated mean. Our smaller rock fragments are probably not the reason for<br />

these different results, as OOSTWOUD WIJDENES et al. (2001) showed that the rock<br />

fragment displacement distance is not depending on mass. The study started in May and in the<br />

first 7 months the displacement distances are lower than in the second 5 months<br />

(OOSTWOUD WIJDENES et al. (2000). In the 5-month period, only one of the grazing peak<br />

months is included; in the 7-month period two of them. That means despite longer measuring<br />

times and despite a higher trampling intensity, the values are higher in the 5-month period. A<br />

reason could be the winter months December and January, in which other erosive processes<br />

influence the displacement distances. This fact emphasizes the need for short-term<br />

experiments to isolate the influence of animal trampling.<br />

GOVERS & POESEN (1998) measured an average horizontal displacement distance of 38 cm<br />

for rock fragments on slopes between 27° and 31°. But the test sites had different widths<br />

(between 1.4 m and 4.1 m) and the number of sheep and goats ranged between 22 and 100. To<br />

make the measured values comparable, they calculated a unit displacement distance in cm per<br />

animal and meter. They presented values between 0.47 cm and 5.6 cm. The highest value was<br />

calculated for the smallest plot, the lowest value for the largest plot. Again, we used a plot<br />

width of 0.6 m and 45 goats and measured an average distance of 8.8 cm. Our test plot was<br />

thus smaller than the smallest one used by GOVERS & POESEN (1998). Calculating the<br />

same unit displacement distance (displacement distance / (number of animals / plot width))<br />

we receive a value of 0.1 cm. Once more, slope is probably responsible for this much lower<br />

value; its effect covers the fact that the plot width of the smallest Govers-Poesen-plot was<br />

twice as large as our plot width. In our experiments, the marked rocks were positioned on top<br />

of given overlying or embedded rock fragments. This means our rocks were probably easier to<br />

move than naturally occurring rock fragments. Even after considering this, our values are still<br />

249


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

much lower than those of GOVERS & POESEN (1998). It is hard to mark existing rock<br />

fragments and measure their translocation distances because their original position cannot be<br />

marked as easily. Knowing the exact original position of each rock fragment is paramount,<br />

however, for recording translocation distances and angels.<br />

The values of mobilized material indicates severe exceeding of the expectable tramplinginduced<br />

erosion rates for the whole study area, especially considering the fact that a mediumsized<br />

flock only crossed our test plots for a single time. The effects of the mobilization of soil<br />

material by the animals’ hooves has scarcely been regarded in literature until now. There is a<br />

plethora of studies, however, on the decrease in vegetation cover by grazing animals and the<br />

consequent increase in bare soil areas that are unprotected from erosion (e.g. ALADOS et al.<br />

2004, ANDRESEN et al. 1990, DUNKERLEY & BROWN 1995, MWENDERA &<br />

MOHAMED SALEEM 1997b, MWENDERA et al. 1997, NOY-MEIR 1975, 1978, 1995,<br />

PEARSON et al. 1990, PEGAU 1970, RIES et al. 2003, VALENTIN 1985, WANG et al.<br />

2002). CERDÀ (1997) showed that runoff and soil loss increase with decreasing vegetation<br />

cover. He used rainfall simulations on bare plots (0-3 % vegetation cover), on plots vegetated<br />

by herbs (30-85 % vegetation cover) and on plots vegetated by tussocks (75- 100 %<br />

vegetation cover). The mean runoff coefficient after 60 min on bare plots was 31 %, on herbs<br />

plots 2 % and on tussock plots 0.3 %. The mean erosion rate on bare plots was 37 g m -2 h -1 , on<br />

herbs plots 1.44 g m -2 h -1 and 0 g m -2 h -1 on tussock plots. Accordingly, the reported processes<br />

of mobilization of soil material and rock fragments by animal hooves need to be investigated<br />

in different areas in future studies because our preliminary findings show that the impacts of<br />

these processes on total erosion rates should be tremendous.<br />

Nevertheless, the experimental designs should be animatedly discussed and need to be further<br />

developed.<br />

5 Conclusions<br />

Independent from the discussible extrapolation of the measured values to a larger area of<br />

several hectares, the determined values have to be evaluated as high, regarding the land-use of<br />

extensive grazing in the test area. Also in comparison to erosion rates on agricultural fields,<br />

the measured values are considered to be too high for a sustainable land use of Mediterranean<br />

rangelands.<br />

Regarding our experimental results, the following conclusions can be drawn:<br />

- Slope and running speed influence the translocation rates by the factor 2.5 to 5.5.<br />

250


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

- Embedded rock fragments appear to be resistant against goat trampling until the<br />

passing of approx. 300 animals.<br />

- Rock fragments that rest loosely on top of the soil surface are prone to be moved by<br />

animal hooves even at low slope angles.<br />

- Crusted soil surfaces provide a high amount of detachable fine material once the crust<br />

is destroyed livestock trampling.<br />

- The provided fine material can be eroded by overland flow within a short time, and is<br />

immediately available for wind erosion processes, too.<br />

- The experimental methods have proven to be suitable for the determination of areaand<br />

animal-number-related rates that are surprising in their magnitude. Nevertheless,<br />

the methods require further validation and development.<br />

The results of the present study support the conclusion that sheep and goat erosion is a<br />

severely underestimated problem.<br />

Acknowledgement<br />

We want to thank all students and the voluntary assistants for their work during the field trip.<br />

Special thanks go to Guidobaldo Hernandez Vico, the herder of the “erosion-goats”. He<br />

provided the goats and he was the only person who was able to keep the 15 crazy critters<br />

under control.<br />

References<br />

ALADOS, L., ELAICH, A., PAPANASTASIS, V.P., OZBEK, H., NAVARRO, T.,<br />

FREITAS, H., VRAHNAKIS, M., LARROSI, D. & CABEZUDO, B. (2004): Change in<br />

plant spatial patterns and diversity along the successional gradient of Mediterranean<br />

grazing ecosystems. - Ecological Modelling 180: 523-535.<br />

ANDRESEN, H., BAKKER, J.P., BRONGERS, M., HEYDEMANN, B. & IRMLER, U.<br />

(1990): Long – term changes of salt marsh communities by cattle grazing. – Vegetatio 89:<br />

137-148.<br />

BARBIER, N., COUTERON, P., LEJOLY, J., DEBLAUWE, V. & LEJEUNE, O. (2006):<br />

Self-organized vegetation patterning as a fingerprint of climate and human impact on<br />

semi-arid ecosystems. – J. Ecol., 94: 537-547.<br />

BOONKORKUEA, N., LENBURY, Y., ALVARADO, F.J. & WOLLKIND, D.J. (2010):<br />

Nonlinear stability analyses of vegetation pattern formation in an arid environment. –<br />

Journal of Biological Dynamics 4, 4: 346-380.<br />

251


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

BORGOGNO, F., D’ODORICO, P., LAIO, F. & RIDOLFI, L. (2009): Mathematical Models<br />

of vegetation pattern formation in ecohydrolgy. – Reviews of Geophysics 47: RG1005,<br />

doi:10.1029/2007RG000256.<br />

BRANSON, F.A., GIFFORD, G.F., RENARD, K.G. & HADLEY, R.F. (1981): Rangeland<br />

hydrology, 2 nd edition. - Kendall/Hunt Pub Co, Dubuque, IA, 340 pp.<br />

CERDÀ, A. (1997): The effect of patchy distribution of stipa tenacissima L. on runoff and<br />

erosion. – Journal of Arid Environments 36: 37-51.<br />

CERDÀ, A. (2001): Effects of rock fragment cover on soil infiltration, interrill runoff and<br />

erosion. – European Journal of Soil Science 52: 59-68.<br />

DEBLAUWE, V., BARBIER, N., COUTERON, P., LEJEUNE, O. & BOGAERT, J. (2008):<br />

The global biogeography of semi-arid periodic vegetation patterns. – Global. Ecol.<br />

Biogeogr. 17: 715-723.<br />

DESIR, G. & MARIN, C. (2007): Factors controlling the erosion rates in a semi-arid zone<br />

(Bardenas Reales, NE Spain). Catena 71: 31-40.<br />

DOPPELBAUER, J.P. (2002): Ziegenzucht und Ziegenhaltung in der EU und den<br />

Beitrittsländern. - 1. Fachtagung für Ziegenzüchter und -halter, 12. - 13. November 2002,<br />

Bundesanstalt für alpenländische Landwirtschaft Gumpenstein, 4 pp.<br />

DUNKERLEY, D.L. & BROWN, K.J. (1995): Runoff and runon areas in a patterned<br />

chenopod shrubland, arid western New South Wales, Australa: characteristics and origin.<br />

– Journal of Arid Environments 30: 41-55.<br />

ELDRIDGE, D. J. (1998): Trampling of microphytic crusts on calcareous soils and its impact<br />

on erosion under rain-impacted flow. – Catena 33: 221-239.<br />

ELWELL, H.A. & STOCKING, M.A. (1976): Vegetal cover to estimate soil erosion hazard<br />

in Rhodesia. – Geoderma 15: 61–70.<br />

EVANS, R. (1998): The erosional impacts of grazing animals. – Progress in Physical<br />

Geography 22, 2: 251-268.<br />

GALLE, S., BROUWER, J. & DELHOUME, J. P. (2001): Soil water Balance. – In:<br />

Tongway, D.J., Valentin, C. & Seghieri, J. (eds.): Banded vegetation patterning in arid<br />

and semiarid environments: ecological processes and consequences for management:77-<br />

104; Springer-Verlag, New York.<br />

GOLODETS, C. & BOEKEN, B. (2006): Moderate sheep grazing in semiarid shrubland<br />

alters small-scale soil surface structure and patch properties. - Catena 65: 285-291.<br />

GOVERS, G. & POESEN, J. (1998): Field experiments on the transport of rock fragments by<br />

animal trampling on scree slopes. – Geomorphology 23: 193-203.<br />

252


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

HILLERISLAMBERS, R., RIETKERK, M., VAN DEN BOSCH, F., PRINS, H.H.T. & DE<br />

KROON, H. (2001): Vegetation pattern formation in semi-arid grazing systems. - Ecology<br />

82(1): 50-61.<br />

JONES, C.G., LAWTON, J.H. & SHACHAK, M. (1994): Organisms as ecosystem engineers.<br />

- OIKOS 69: 373-386.<br />

MOLINILLO, M., LASANTA, T., & GARCIA-RUIZ, J. (1997): Managing mountainous<br />

degraded landscapes after farmland abandonment in the Central Spanish Pyrenees. -<br />

Environmental Management 21(4): 587-98.<br />

MONFREDA, C., RAMANKUTTY, N. & FOLEY, J. A. (2008): Farming the planet: 2.<br />

Geographic distribution of crop areas, yields, physiological types, and net primary<br />

production in the year 2000. - Global Biogeochem. Cycles 22: GB1022,<br />

doi:10.1029/2007GB002947.<br />

MONFREDA, C., RAMANKUTTY, N. & HERTEL, T.W. (2009): Global Agricultural Land<br />

Use Data for Climate Change Analysis. – In: Hertel, T.W., Rose, S. & Tol, R.S.J. (eds):<br />

Economic Analysis of Land Use in Global Climate Change Policy: 33-48; Routledge.<br />

MONTANA, C. (1992): The colonization of bare areas in two-phase mosaics of an arid<br />

ecosystem. - Journal of Ecology 80: 315-327.<br />

MULHOLLAND, B. & FULLEN, M.A. (1991): Cattle trampling and soil compaction on<br />

loamy sands. - Soil Use and Management 7,4: 189-192.<br />

MWENDERA, E.J. & MOHAMED SALEEM, M.A. (1997a): Hydrologic response to cattle<br />

grazing in the Ethiopian highlands. - Agriculture, Ecosystems and Environment 64: 33-41.<br />

MWENDERA, E.J. & MOHAMED SALEEM, M.A. (1997b): Infiltration rates, surface<br />

runoff, and soil loss as influenced by grazing pressure in the Ethiopian highlands. - Soil<br />

Use and Management 13: 29-35.<br />

MWENDERA, E.J., MOHAMED SALEEM, M. A. & WOLDU, Z. (1997): Vegetation<br />

response to cattle grazing in the Ethiopian highlands. - Agriculture, Ecosystems and<br />

Environment 64: 43-51.<br />

NOY-MEIR, I. (1975): Stability of grazing systems: an application of predator-prey graphs. -<br />

Journal of Ecology 63: 459-481.<br />

NOY-MEIR, I. (1978): Grazing and production in seasonal pastures: analysis of a simple<br />

model. - Journal of Applied Ecology 15: 809-835.<br />

NOY-MEIR, I. (1995): Interactive effects of fire and grazing on structure and diversity of<br />

Mediterranean grasslands. - Journal of Vegetation Science 6: 701-710.<br />

253


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

NYSSEN, J., POESEN, J., MOEYERSONS, J., DECKERS, J. & HAILE, M. (2006):<br />

Processes and rates of rock fragment displacement on cliffs and scree slopes in an amba<br />

landscape, Ethiopia. – Geomorphology 81: 265-275.<br />

OLDEMANN, L.R., HAKKELING, R.T.A. & SOMBROEK, W.G. (1991): World map of the<br />

status of humaninduced soil degradation. An explanatory note. Wageningen: International<br />

Soil Reference and Information Centre/Nairobi: United Nations Environment Programme,<br />

41 pp.<br />

OOSTWOUD WIJDENES, D.J., POESEN, J., VANDEKERCKHOVE, L. & KOSMAS, C.<br />

(2000): The use of marked rock fragments as tracters to assess rock fragments transported<br />

by sheep trampling on Lesvos, Greece. – In: Foster, I.D.L. (ed): Tracers in<br />

Geomorphology. John Wiley & Sons, LTD: 201-220.<br />

OOSTWOUD WIJDENES, D.J., POESEN, J., VANDEKERCKHOVE, L. & KOSMAS, C.<br />

(2001): Measurements at one-year interval of rock-fragment fluxes by sheep trampling on<br />

degraded rocky slopes in the Mediterranean. – Zeitschrift für Geomorphologie 45,4: 477-<br />

500.<br />

PARIENTE, S. (2002): Spatial patterns of soil moisture as affected by shrubs, in different<br />

climate conditions. – Environmental Monitoring and Assessment 73: 237-251.<br />

PEARSON, H.A., BALDWIN, V.C. & BARNETT, J.P. (1990): Cattle grazing and pine<br />

survival and growth in subterranean clover pasture. – Agroforestry Systems 10: 161-168.<br />

PEGAU, R.E. (1970): Effect of reindeer trampling and grazing on lichens. – Journal of Range<br />

Management 23,2: 95-97.<br />

PIETOLA, L., HORN, R. & YLI-HALLA, M. (2005): Effects of trampling by cattle on the<br />

hydraulic and mechanical properties of soil. – Soil and tillage research 82: 99 – 108.<br />

PROFFITT, A.P.B., BENDOTTI, S. & MCGARRY, D. (1995): A comparison between<br />

continuous and controlled grazing on a red duplex soil. I. Effects on soil physical<br />

characteristics. - Soil & Tillage Research 35: 199-210.<br />

RAMANKUTTY, N., EVAN, A. T., MONFREDA, C. & FOLEY, J. A. (2008): Farming the<br />

planet: 1. Geographic distribution of global agricultural lands in the year 2000. - Global<br />

Biogeochem. Cycles 22: GB1022, doi:10.1029/2007GB002947.<br />

RIES, J.B. (2010): Methodologies for soil erosion and land degradation assessment in<br />

Mediterranean-type ecosystems. – Land Degradation & Development 21: 171-187.<br />

RIES, J.B., SEEGER, M., ISERLOH, T., WISTORF, S. & FISTER, W. (2009): Calibration of<br />

simulated rainfall characteristics for the study of soil erosion on agricultural land. - Soil &<br />

Tillage Research 106:109-116.<br />

254


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

RIES, J. B. (2005): Soil Erosion on Abandoned Fields in Mediterranean Mountains -<br />

Monitoring of Processes and Development. - Journal of Mediterranean Ecology 6,1: 43-<br />

52.<br />

RIES, J.B., MARZOLFF, I. & SEEGER, M. (2003): Einfluss der Beweidung auf<br />

Vegetationsbedeckung und Geomorphodynamik zwischen Ebrobecken und Pyrenäen. -<br />

Geographische Rundschau 55, 5: 52-59.<br />

ROSTANGO, C. M., DEL VALLE, H. F. & <strong>VI</strong>DELA, L. (1991): The influence of shrubs on<br />

some chemical and physical properties of an aridic soil in north-eastern Patagonia,<br />

Argentina. - Journal of Arid Environments 20: 179-188.<br />

SCHAICH, H. (2007): Analyse raum-zeitlicher Prozesse der Vegetationsentwicklung unter<br />

Beweideeinfluss in der renaurierten Syr-Aue (Luxemburg). – Cultera 51: 179-191.<br />

SEEGER, M. (2007): Uncertainty of factors determining runoff and erosion processes as<br />

quantified by rainfall simulations. - Catena, 71(1): 56-67.<br />

SEUFFERT, O., HERRIG K., OLLESCH G. & BUSCHE, D. (1999): REI - An integrated<br />

rainfall erosivity index for assessing and correlating rainfall structure, runoff and erosion.<br />

- Geoökodynamik 20: 1-54.<br />

STA<strong>VI</strong>, I., UNGAR, E.D., LAVEE, H. & SARAH, P. (2008a): Grazing-induced spatial<br />

variability of soil bulk density and content of moisture, organic carbon and calcium<br />

carbonate in a semi-arid rangeland. – Catena 75: 288-296.<br />

STA<strong>VI</strong>, I., UNGAR, E.D., LAVEE, H. & SARAH, P. (2008b): Surface microtopography and<br />

soil penetration resistance associated with shrub patches in a semiarid rangeland. –<br />

Geomorphology 94: 69-78.<br />

STA<strong>VI</strong>, I., UNGAR, E.D., LAVEE, H. & SARAH, P. (2009): Livestock modify ground<br />

surface microtopography and penetration resistance in a semi-arid shrubland. – Arid Land<br />

Research Management 23: 237-247.<br />

STEWART, D.P.C. & CAMERON, K.C. (1992): Effect of trampling on the soils of the St<br />

James Walkway, New Zealand. – Soil Use and Management 8,1: 30-36.<br />

TABOADO, M.A. & LAVADO, R.S. (1993): Influence of cattle trampling on soil porosity<br />

under alternate dry and ponded conditions. – Soil use and Management 9,4: 139-143.<br />

TOLLNER, E.W., CALVERT, G.V. & LANGDALE, G. (1990): Animal trampling effects on<br />

soil physical properties of two southeastern U.S. ultisols. – Agriculture, Ecosystems and<br />

Environments 33: 75-87.<br />

255


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

TONGWAY, D.J., VALENTIN, C. & SEGHIERI, J. (2001): Banded Vegetation Patterning in<br />

Arid and Semiarid Environments: Ecological Processes and Consequences for<br />

Management. – Springer, 1 st Edition, New York, 267 pp.<br />

TONGWAY, D. J. & LUDWIG, J. A. (2001): Theories on the origins, maintenance,<br />

dynamics, and functioning of banded landscapes. – In: Tongway, D.J., Valentin, C. &<br />

Seghieri, J. (eds.): Banded vegetation patterning in arid and semiarid environments:<br />

ecological processes and consequences for management: 20-31; Springer-Verlag, New<br />

York.<br />

UNGAR, E. D., STA<strong>VI</strong>, I., LAVEE, H. & SARAH, P. (2009): Effects of livestock traffic on<br />

rock fragment movement on hillsides in a semiarid patchy rangeland. - Land Degradation<br />

& Development 20: 1-8.<br />

VALENTIN, C. (1985): Effects of grazing and trampling on soil deterioration around recently<br />

drilled water holes in the Sahelian Zone. – In: El-Swaify, S.A., Moldenhauer, W.L. & Lo<br />

A. (eds): Soil Erosion and Conservation. Soil Cons. Soc. of America, Ankeny, USA: 51-<br />

65.<br />

WANG, Y., SHIYOMI, M., TSUIKI, M., TSUTSUMI, M., YU, X. & YI, R. (2002): Spatial<br />

heterogeneity of vegetation under different grazing intensities in the Northwest<br />

Heilongjiang Steppe of China. - Agriculture, Ecosystems and Environment 90: 217-229.<br />

WIRTZ, S., SEEGER, M. & RIES, J.B. (2010): The rill experiment as a method to approach a<br />

quantification of rill erosion process activity. – Zeitschrift für Geomorphologie 54,1: 47-<br />

64.<br />

WIRTZ, S., SEEGER, M. & RIES; J.B. (2012): Field experiments for understanding and<br />

quantification of rill erosion processes. – Catean 91:21-34.<br />

WORRALL, G. A. (1959): The Butana grass patterns. - Journal of Soil Science 10: 34-53.<br />

256


Experimentelle Rinnenerosionsforschung vs. Modellkonzepte – Quantifizierung der hydraulischen und erosiven Wirksamkeit von Rinnen<br />

Angaben zur Person<br />

Name, Vorname<br />

Geburtsdatum<br />

Familienstand<br />

Wissenschaftlicher Werdegang<br />

<strong>Wirtz</strong>, <strong>Stefan</strong><br />

17.01.1982, Saarbrücken<br />

ledig<br />

Berufserfahrung<br />

Seit Januar 2011<br />

Wissenschaftlicher Mitarbeiter im Fach Physische<br />

<strong>Geographie</strong><br />

Juli 2008 – Dezember 2010<br />

Promotions-Stipendiat an der Universität Trier,<br />

gefördert durch das Land Rheinland-Pfalz (GraFöG)<br />

Ausbildungsweg<br />

Juli 2008 -<br />

Doktorand an der Universität Trier, FB <strong>VI</strong><br />

<strong>Geographie</strong>/ <strong>Geowissenschaften</strong><br />

September 2002 bis Dezember 2007<br />

Studium der Angewandte Physische <strong>Geographie</strong> an<br />

der Universität Trier, Abschluss: Diplom-Geograph<br />

Nebenfächer Bodenkunde und Geologie<br />

Thema der Diplomarbeit: Spülversuche als<br />

experimentelle Methode zur Rinnenerosionsforschung<br />

Juli 2001-März 2002<br />

Wehrdienst Fallschirmjäger FschJgBtl. 261 Lebach<br />

1992-2001 Realgymnasium Völklingen, Abschluss: Abitur<br />

257


Dipl. Geogr. <strong>Stefan</strong> <strong>Wirtz</strong>; Physische <strong>Geographie</strong>, Universität Trier, Behringstraße, 54286 Trier<br />

Lebenslauf<br />

Persönliche Daten<br />

Name<br />

<strong>Wirtz</strong>, <strong>Stefan</strong><br />

Anschrift privat Thomasstraße 2<br />

54316 Franzenheim<br />

Telefon: +496588 982889<br />

E-Mail: stefanw-170182@t-online.de<br />

Geburtsdatum<br />

Geburtsort<br />

Staatsangehörigkeit<br />

Familienstand<br />

17.01.1982<br />

Saarbrücken<br />

Deutsch<br />

ledig<br />

Ausbildung<br />

1988-1992<br />

1992-2001<br />

Juli 2001-März 2002<br />

September 2002<br />

- Dezember 2007<br />

Juli 2008 – Dezember 2010<br />

Seit Januar 2011<br />

Grundschule Viktoria, Püttlingen<br />

Realgymnasium Völklingen, Abschluss: Abitur<br />

Wehrdienst Fallschirmjäger FschJgBtl. 261 Lebach<br />

Universität Trier<br />

Studiengang Angewandte Physische <strong>Geographie</strong><br />

Nebenfächer Bodenkunde und Geologie<br />

Abschluss mit Note „sehr gut“<br />

Thema der Diplomarbeit: Spülversuche als experimentelle<br />

Methode zur Rinnenerosionsforschung<br />

Promotionsstipendium des Landesgraduiertenzentrums der<br />

Universität Trier<br />

Wissenschaftlicher Mitarbeiter der Universität Trier,<br />

Physische <strong>Geographie</strong><br />

258


Dipl. Geogr. <strong>Stefan</strong> <strong>Wirtz</strong>; Physische <strong>Geographie</strong>, Universität Trier, Behringstraße, 54286 Trier<br />

Arbeitserfahrung<br />

März/April 2005<br />

März/April 2006<br />

Seit August 2012<br />

Praktikum beim Landesamt für Umweltschutz, Abteilung<br />

Geologie und Boden, Saarbrücken<br />

Praktikum bei der NABU Landesgeschäftsstelle Saarland,<br />

Lebach<br />

Aushilfe bei der EGP – Gesellschaft für urbane<br />

Projektentwicklung<br />

Sprachkenntnisse<br />

Englisch gut, Französisch Grundkenntnisse<br />

EDV-Kenntnisse<br />

MS Office, ArcGIS, Photoshop<br />

Interessen und Hobbys<br />

Leichtathletik, Kampfsport (Ju-Jutsu)<br />

Lehrveranstaltungen:<br />

Semester Veranstaltung Leitung<br />

SoSe 2006 2006 Geländepraktikum "Bodenphysik" Schneider, <strong>Wirtz</strong><br />

SoSe 2008 2008 Lehrforschungsprojekt Physische <strong>Geographie</strong> Geländeseminar Ries, Iserloh, <strong>Wirtz</strong><br />

SoSe 2008 2008 Lehrforschungsprojekt Physische <strong>Geographie</strong> Vorbereitungsseminar Ries, Iserloh, <strong>Wirtz</strong><br />

WiSe 2008 / 2009 Lehrforschungsprojekt Datenanalyse und Darstellung Ries, Iserloh, <strong>Wirtz</strong><br />

SoSe 2009 2009 Lehrforschungsprojekt Physische <strong>Geographie</strong> Geländeseminar Ries, Iserloh, <strong>Wirtz</strong><br />

SoSe 2009 2009 Lehrforschungsprojekt Physische <strong>Geographie</strong> Vorbereitungsseminar Ries, Iserloh, <strong>Wirtz</strong><br />

SoSe 2009 2009 Übung Grundlagen der Physischen <strong>Geographie</strong> II BSc <strong>Wirtz</strong><br />

WiSe 2009 / 2010 Lehrforschungsprojekt Datenanalyse und Darstellung Ries, Iserloh, <strong>Wirtz</strong><br />

SoSe 2010 2010 Geländepraktikum "Geomorphologische Kartierübung" Iserloh, <strong>Wirtz</strong><br />

SoSe 2010 2010 Tutorium für Bachelor- Master- Diplomarbeiten Iserloh, <strong>Wirtz</strong><br />

SoSe 2011 2011 Lehrforschungsprojekt Physische <strong>Geographie</strong> Geländeseminar Ries, Iserloh, <strong>Wirtz</strong><br />

SoSe 2011 2011 Lehrforschungsprojekt Physische <strong>Geographie</strong> Vorbereitungsseminar Ries, Iserloh, <strong>Wirtz</strong><br />

SoSe 2011 2011 Übung Grundlagen der Physischen <strong>Geographie</strong> II BEd <strong>Geographie</strong> <strong>Wirtz</strong><br />

SoSe 2011 2011 Übung Grundlagen der Physischen <strong>Geographie</strong> II BSc <strong>Wirtz</strong><br />

WiSe 2011 / 2012 Lehrforschungsprojekt Datenanalyse und Darstellung Ries, Iserloh, <strong>Wirtz</strong><br />

WiSe 2011 / 2012 Übung Grundlagen der Physischen <strong>Geographie</strong> I BEd <strong>Geographie</strong> <strong>Wirtz</strong><br />

WiSe 2011 / 2012 Übung Grundlagen der Physischen <strong>Geographie</strong> I Ökozonen der Erde BSc <strong>Wirtz</strong><br />

SoSe 2012 2012 Tutorium für Bachelor- Master- Diplomarbeiten Iserloh, <strong>Wirtz</strong><br />

SoSe 2012 2012 Übung Grundlagen der Physischen <strong>Geographie</strong> II BSc <strong>Wirtz</strong><br />

…………….……………………………………<br />

(Dipl. Geogr. <strong>Stefan</strong> <strong>Wirtz</strong>)<br />

259

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!