13.02.2013 Views

download "Translations" - MDC

download "Translations" - MDC

download "Translations" - MDC

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.

Note<br />

This is a story about people, science, and ideas, written for anyone who is curious about today’s research and its<br />

practitioners.<br />

Society is in the early stages of a revolution which is already changing our lives and will have a much greater impact in<br />

the years to come. Until recently it was happening quietly, in the cell cultures and laboratory flasks of research institutes<br />

all over the world. But now not a day goes by without important new discoveries that shed light on fundamental<br />

questions about life and give insights into the processes that lead to cardiovascular and neurodegenerative diseases,<br />

cancer, AIDS, malaria, and the other major afflictions of mankind. While these discoveries are as much a cultural treasure<br />

as great works of literature or art, they are harder to appreciate because they usually concern details of biological<br />

processes that can’t really be put into context in soundbytes or newspaper articles.<br />

Sometimes the best way to understand a whole is to observe a small part in great detail and show how it is related to<br />

everything else. The campus in Berlin-Buch is a microcosm of what is going on in experimental and clinical research<br />

laboratories all over the world. By focusing on the work being done there, this book aims to open a window on that much<br />

larger context. At the same time, the campus is home to unique people with a special relationship to history, culture and<br />

the arts. Their stories are fascinating in their own right.<br />

As big as this book is, it presents only a small part of the research on campus. Other groups have equally interesting<br />

stories that are waiting to be told. These stories are intertwined with those of many others who quietly work in the<br />

background to make the science possible: technicians, administrators, the teachers of the Gläsernes Labor, secretaries,<br />

architects, cooks, cleaners, drivers, etc. This book is dedicated to them. It is also dedicated to the dozens of scientists who<br />

directly participated in its production and permitted me to write about their lives in a very personal way. I apologize for<br />

the mistakes that remain despite their immense patience in explaining their work.<br />

I would particularly like to thank Walter Birchmeier, Jens Reich, Heinz Bielka, Walter Rosenthal, Friedrich Luft, Björn Maul,<br />

Gudrun Erzgräber, and Detlev Ganten for many stimulating discussions and help with the manuscript. Okan Toka very<br />

kindly spent a lot of time sharing his adventures in the Muskateer story, for which Silvia Bähring provided most of the<br />

photos; some of them were taken by Hakan Toka. Ewan St. J. Smith provided images of the naked mole rat, and Catherine<br />

Adamidi photographed the planaria. Peter Himsel and David Aussenhofer also generously provided photos. Elizabeth<br />

Kujawa-Schmeitzner, Christina Quensel, Maxine Saborowski and Wolfgang Groll provided invaluable organizational and<br />

moral help. Jane Holland helped with the interviews. Thanks also to Andreas Maetzold and Ulrich Scheller for many<br />

discussions about the campus and its future. Wolfgang Uckert had been saving some old laboratory bottles for many<br />

years, knowing that they would eventually be needed for something, and they finally found their destiny on the cover of<br />

this book. The Ivics family invited us into their home for pictures at the dinner table. Caroline Hadley did an excellent job<br />

of critically reading and editing the manuscript. Thanks to Dr. Engelsing and his colleagues at the museum of the City of<br />

Konstanz for a superb image of the Leiner collection. My appreciation to Hus No. 37 (the Swedish coffee shop in Pankow),<br />

and Star Coffee in Heidelberg, where large portions of this book were written. Thanks to Rainer Bender, Ingrid André, and<br />

the team of ColorDruck for their sharp eyes and good ideas and all of their input at many stages of this project. Maj Britt<br />

thanks in particular Manuela Brunner, Anita Glanz and Andreas Schmidt. She also thanks Klaus and the Vorgrimmler<br />

family for all their support and help. Nicola thanks her family: Jürgen, Lynn and Franka for their patience and support.<br />

I would like to thank my family - Gabi, Jesper, Sharon, and Lisa - for the many sacrifices they have made so that this book<br />

could exist.<br />

Russ Hodge<br />

May, 2008


translations<br />

from today’s science<br />

to tomorrow’s medicine<br />

in Berlin-Buch<br />

by Russ Hodge<br />

Maj Britt Hansen, photography<br />

Nicola Graf, design


Table of contents<br />

Science and the city: a drive to Berlin-Buch............................................................. 4<br />

Part One: The lives of nomads .................................................................................................. 20<br />

Songlines and shapeshifters ......................................................................................... 22<br />

The archeology of blood ................................................................................................. 40<br />

Breakouts in the brain..................................................................................................... 52<br />

Stem cells in a jar .............................................................................................................. 62<br />

Capturing a firefly............................................................................................................. 70<br />

A very quiet cure................................................................................................................ 78<br />

A brief incident of teleportation.................................................................................. 86<br />

Interlude: A glass heart.................................................................................................................. 88<br />

Part Two: Identity crisis.................................................................................................................. 100<br />

Where slow rivers meet.................................................................................................. 102<br />

Lost anchors and wrecked vessels............................................................................... 112<br />

A pact with the devil........................................................................................................ 118<br />

The decorators of Versailles........................................................................................... 132<br />

Waking a Sleeping Beauty and other tales of ancient genes............................ 140<br />

Table of contents<br />

2


Interlude: When rats don’t dance to the red-hot chili peppers......................... 150<br />

Part Three: Frontiers and ferrymen........................................................................................ 158<br />

The Swiss army knife of the cell .................................................................................. 160<br />

Bad origami......................................................................................................................... 174<br />

Trouble in the waterworks............................................................................................. 184<br />

Bribing the ferryman ....................................................................................................... 194<br />

The electrician’s toolbox................................................................................................. 200<br />

A kiss, interrupted ............................................................................................................. 210<br />

Interlude: The case of the short-fingered musketeer............................................... 218<br />

Part Four: From the protein village to the cosmopolitan cell ............................. 238<br />

Interviews:<br />

A hub in the city of the cell ........................................................................................... 240<br />

The pleasures and powers of green tea.................................................................... 250<br />

Playing the piano of planaria........................................................................................ 262<br />

A rat in a tree and a wolf on the loose ...................................................................... 274<br />

Walter Birchmeier, Scientific Director, <strong>MDC</strong> ............................................................ 284<br />

Walter Rosenthal, Director, FMP................................................................................... 288<br />

Detlev Ganten, Charité.................................................................................................... 292<br />

Gudrun Erzgräber, BBB .................................................................................................... 296<br />

Index.............................................................................................................................................................. 300<br />

Further reading.......................................................................................................................................... 302<br />

Credits........................................................................................................................................................... 305<br />

3 Table of contents


Science and the city:<br />

a drive to Buch<br />

This book was born during a ride through<br />

Berlin on a luminous autumn day in 2006.<br />

The train had come in from the west, gliding over a<br />

landscape of parks so spacious and green that it was<br />

often hard to believe we were entering the heart of a world<br />

capital. We passed above the zoo, where a long line of children<br />

waited eagerly for their first look at a newborn polar bear. “Knut”<br />

could be seen everywhere, in gift shops, the city’s newspapers and on<br />

the title pages of magazines. He had become a symbol of the zoo, of the<br />

city, of environmental campaigns. Global warming was high on everyone’s<br />

minds; people had been talking about it on the train. November<br />

had been so mild that the trees were only now changing color.<br />

The train was full of people on their way to Berlin – tourists, students,<br />

half an orchestra, scientists, business people talking loudly into cell<br />

phones. It was always interesting to strike up a conversation with a<br />

neighbor; people were being drawn to the city for a thousand different<br />

reasons. The conversations stopped as the train slid into Berlin’s new<br />

main station, a cathedral of glass and girders that decorated several levels<br />

of platforms in light and shadows.<br />

Science and the city a drive to Buch<br />

4<br />

Berlin main train station


At the top of the stairs a man in a leather jacket<br />

asked people for used S-Bahn tickets.<br />

Andreas Hollman, a driver from the Max<br />

Delbrück Center, was waiting at the bottom of<br />

a labyrinth of escalators. There would be a<br />

short walk to the car, for which he apologized.<br />

The capital city had built a fantastic Bahnhof,<br />

he said, but it was miserable for drivers.<br />

Hundreds of people needed to be dropped off<br />

and picked up every hour, and the one small<br />

street in front couldn’t handle the traffic.<br />

Leaving the station, we caught a glimpse of<br />

the glass dome of the Bundestag. Around the<br />

corner stood the tall brick buildings of the<br />

“Charité”, one of the world’s largest and oldest<br />

hospitals, founded in 1710 in anticipation of an<br />

outbreak of the bubonic plague that never<br />

came. Andreas pointed out the television<br />

tower that looms over the Alexanderplatz, in<br />

the distance. “The round part at the top<br />

Knut and Thomas Dörflein<br />

at the Berlin Zoo<br />

looked like a soccer ball for the World Cup.<br />

They just removed the decorations.” He<br />

paused. “They should have left it.”<br />

The scientific campus of Berlin-Buch lies on<br />

the northeastern outskirts of the city and is<br />

home to a number of important insitutions:<br />

the Max Delbrück Center for Molecular<br />

Medicine (<strong>MDC</strong>), the Leibniz Institute for<br />

Molecular Pharmacology (FMP), clinics of the<br />

Charité, and a number of biotech companies.<br />

If we had made this drive in 1988, the first<br />

time I had visited Berlin, we would have had<br />

to cross the border into East Germany. By the<br />

time of my next visit, in the mid-1990s, the<br />

Wall was gone but navigation through the<br />

city was hindered by hundreds of construction<br />

sites. Now there were few obvious signs<br />

of either phase of the city’s recent past. The<br />

drive took us past a huge building complex of<br />

the Bayer Schering Pharmaceutical company,<br />

through Wedding, along the Schönhauser<br />

Allee into Pankow, then into the quiet neighborhoods<br />

of the suburbs.<br />

Andreas often picked up new staff members<br />

like me or people coming for interviews and<br />

was happy to play tour guide. If most of Berlin<br />

didn’t have the monumental personality of<br />

London or Paris, he said, the reason lay with<br />

history. The city had not begun as a capital but<br />

rather as separate villages that grew together.<br />

That explained the low buildings, the green<br />

spaces. Now those villages had become<br />

neighborhoods with distinct personalities.<br />

Buch has its personality as well. In the late 19th century Berlin – like most of industrialized<br />

Europe – was afflicted by an epidemic of<br />

tuberculosis. Hospital wards were so full that<br />

patients spilled out into the hallways; beds<br />

were set up wherever space could be found.<br />

Science a nd the city a drive to Buch<br />

6<br />

Infectious diseases love a crowd, and health<br />

officials decided to build new hospitals outside<br />

the center to limit their further spread.<br />

One of Germany’s most distinguished architects,<br />

Ludwig Hoffmann, was hired to begin a<br />

massive construction project in Buch.<br />

Beautiful brick buildings with balconies and<br />

gables were erected on several sites around<br />

the village. Buch quickly became known as


the “Hospital city,” or “the city of the ill.” That<br />

tradition has played an important role in the<br />

current incarnation of Buch as one of<br />

Germany and Europe’s most important centers<br />

for biomedical research.<br />

The campus institutes are devoted to “molecular<br />

medicine,” reflecting the promise that<br />

what has been learned about the basic units<br />

of life – DNA, proteins, and the other mole-<br />

cules within our cells – can be transformed<br />

into new methods of diagnosing and treating<br />

diseases. Fifteen years ago the campus in<br />

Buch was one of the earliest centers in Europe<br />

to devote itself to this theme. At the time it<br />

was a bold move. Laboratories and clinics<br />

belonged to different scientific cultures, concerned<br />

with different questions, using different<br />

methods, and training their students very<br />

7 Science and the city a drive to Buch<br />

Roland Schmidt<br />

differently. To arrange a marriage between the<br />

cultures, they first had to be brought together.<br />

With its existing clinics and a strong biology<br />

institute, Buch seemed to be the ideal<br />

place to build bridges between the treatment<br />

of patients and research labs devoted to fundamental<br />

(and often arcane) questions about<br />

the functions of genes and the inner lives of<br />

cells.


In the meantime many institutes are defining<br />

(or redefining) themselves around this theme.<br />

That reflects progress in science, but also<br />

some realities about science politics. Most<br />

research in molecular biology is carried out in<br />

publicly-funded universities or other academic<br />

institutes. They regularly have to compete<br />

for funding, and one source is the large sums<br />

that politicians are willing to spend on medical<br />

research. So many institutes are redefining<br />

themselves as centers for molecular medi -<br />

cine, in so many contexts, that it is often hard<br />

to understand what the term means.<br />

I didn’t really know what it meant, although I<br />

had been working alongside scientists for<br />

more than a decade. Maybe the campus<br />

would provide an answer. If anyone could<br />

define molecular medicine, it ought to be the<br />

scientists of Berlin-Buch, who have been practicing<br />

it for 15 years. This book reflects what<br />

many of them have to say.<br />

Science and the city a drive to Buch<br />

8<br />

��<br />

During the drive Andreas Hollman talked<br />

about modern Berlin and its spectrum<br />

of cultural offerings: museums, theaters, concerts,<br />

lectures and literary events. He pointed<br />

out traces of history: high-rise buildings that<br />

have sprung up in the empty zone that used<br />

to border the Wall; houses still bear scars from<br />

shells fired during the Second World War. He<br />

didn’t say much about Berlin’s scientific histo-


y although many of the world’s greatest scientists<br />

have lived and worked in Berlin, and<br />

their tracks can be found throughout the city.<br />

Following their trail is worthwhile; one way to<br />

understand what is so unique about today’s<br />

science is to take a brief look back at its roots<br />

in the 19th century.<br />

Within just a few decades in the mid-1800s, a<br />

number of thinkers – almost entirely from<br />

Europe – completely revolutionized the life<br />

sciences. They created modern medicine, cell<br />

biology, evolution, genetics, embryology,<br />

chemistry, physics, and other fields. There is a<br />

similar revolution going on today. It also<br />

involves changes in our concepts about<br />

nature and life, but its main feature is how all<br />

of these fields are rapidly coming together in<br />

new ways.<br />

There was time to drive through town, so<br />

Andreas followed the Friedrichstrasse to<br />

Unter den Linden. Most tourists come here to<br />

see the Brandenburg Gate and to stroll down<br />

9<br />

Science and the city a drive to Buch<br />

the spacious, tree-lined boulevards. Farther<br />

down the street is a marble statue of<br />

Alexander von Humboldt, the namesake of<br />

Berlin’s largest university, on a pedestal outside<br />

some of the buildings where he gave lectures<br />

in 1827 and 1828. By that time Humboldt<br />

was rapidly becoming one of the most<br />

famous men in the world. He was a universal<br />

genius and author of articles on every conceivable<br />

scientific topic – from mineral formations<br />

on the Rhine to the vegetation that


grew in mines, the plants and animals of South<br />

America, the movement of continents, and the<br />

composition of the atmosphere. His accounts<br />

of his expeditions through South America combined<br />

exact observations with the romance of<br />

discovery and inspired generations of young<br />

naturalists, including Charles Darwin and<br />

Alfred Russel Wallace. Their own journeys in the<br />

tropics led to the theory of evolution.<br />

Humboldt’s generation and the next saw the<br />

birth of most of the fields that make up<br />

today’s biology and medicine. One of those<br />

disciplines, cell biology, began with a chance<br />

meeting at a Berlin train station. A biology<br />

student named Theodor Schwann struck up a<br />

conversation with Mathias Schleiden, an<br />

eccentric student from Heidelberg. There<br />

Schleiden had studied law, even opening a<br />

practice when he moved to Hamburg, but<br />

bouts of depression led to an unsuccessful<br />

attempt at suicide. As a result he carried a bullet<br />

in his brain for the rest of his life. Now he<br />

had moved to Berlin to start over, this time as<br />

a botanist.<br />

Alexander von Humboldt Hermann von Helmholtz<br />

Schwann had arrived from Bonn with one of<br />

his professors, Johann Müller, who pushed his<br />

students to use a new type of microscope just<br />

invented by the Englishman Joseph Lister. The<br />

instrument was made of two lenses mounted<br />

in a tube, a construction that solved some<br />

problems of distortion and dramatically<br />

increased its resolution. It gave scientists their<br />

sharpest view ever of the microscopic world.<br />

Schleiden’s observations led him to realize<br />

that plants were built of fundamental units –<br />

single cells – which somehow formed from<br />

the nuclei of other cells. He had developed a<br />

close friendship with Schwann and mentioned<br />

the discovery to him one night over<br />

dinner. Schwann immediately realized that<br />

the idea might explain what he had been<br />

observing in animal tissues. The two men left<br />

their meal half-eaten and rushed over to<br />

Schwann’s laboratory. Previously, anatomists<br />

had believed that animal tissues were made<br />

of fibers, grains, tubes, and other objects. They<br />

had been looking at cells, Schwann said, without<br />

knowing it.<br />

Science and the city a drive to Buch<br />

10<br />

Many of Müller’s students went on to become<br />

pioneers in biology and medicine. Today the<br />

Charité’s institute of pathology is named after<br />

one of the greatest of them, Rudolf Virchow.<br />

After his studies Virchow received a double<br />

professorship from the university and the<br />

Charité, where he treated patients and continued<br />

investigating basic questions about cells.<br />

He took Schleiden and Schwann’s observations<br />

a step further by proving that a cell<br />

could only arise from another cell. Previously<br />

many scientists had believed they could arise<br />

by themselves, somehow crystallizing from<br />

more basic substances. Virchow’s simple new<br />

idea had a huge impact because it changed<br />

the way scientists thought about all sorts of<br />

questions, from the growth of embryos to the<br />

nature of disease.<br />

For example, it gave Virchow a new view of<br />

cancer. He realized that tumors arose from<br />

small pools of cells that divided too often, in<br />

the wrong places. Removing the source might<br />

stop the spread of the disease. He developed<br />

new laboratory methods to diagnose the seri-


ousness of cancer and new surgical procedures<br />

to treat it. Some of his ideas were ahead<br />

of their time: his experience with patients<br />

suggested that the disease often arose at<br />

sites of injuries or infections. This made him<br />

think that tumors might be linked to inflammations<br />

– the body’s response to injuries. For<br />

over a century most scientists rejected this<br />

idea, but today scientists in Buch and elsewhere<br />

have gathered convincing evidence<br />

that he was right. There are clear links<br />

between some types of cancer and autoimmune<br />

diseases in which the body has trouble<br />

distinguishing between its own and foreign<br />

cells. Virchow’s discoveries likewise exposed<br />

the origins of other diseases. For example, he<br />

was the first to understand how blood vessels<br />

could become blocked by clots that formed<br />

elsewhere in the body, a process which he<br />

called embolism.<br />

One of Virchow’s colleagues, Robert Koch, laid<br />

another cornerstone of modern medicine. He<br />

discovered that a bacteria caused anthrax and<br />

– with Louis Pasteur – proposed the theory<br />

Old train station at<br />

Pankow-Heinersdorf<br />

that all infectious diseases were caused by<br />

microorganisms. Over the next few years<br />

Koch supported the theory with the discovery<br />

of the sources of tuberculosis and cholera,<br />

also bacteria. By 1891 his reputation earned<br />

him the directorship of a new Institute of<br />

Infectious Diseases in Berlin, which soon<br />

moved to a new building in Berlin-Wedding.<br />

Now called the Robert Koch Institute, it can<br />

still be found at Norduferstrasse 20. One of<br />

Koch’s first acts as director was to hire Paul<br />

Ehrlich, who discovered substances that could<br />

treat sleeping sickness and syphilis. His work<br />

helped lay the foundations of modern pharmacology.<br />

Today the development of new drugs lies<br />

mainly in the hands of chemists, and that was<br />

the subject chosen by a young Scotsman<br />

named Archibald Couper when he enrolled at<br />

the University of Berlin in 1854. Four years<br />

later he drew the first map of the positions of<br />

atoms inside a molecule. Prior to this time<br />

most chemists believed that molecules<br />

(assemblies of many atoms) could be under-<br />

Diagrams of molecules by Archibald Couper<br />

stood simply by listing the proportions of<br />

each element they contained. Couper showed<br />

that to understand their behavior, you also<br />

had to know how their atoms were arranged.<br />

(A German biochemist named Auguste<br />

Kekulé, working in Heidelberg and Ghent, had<br />

come to the same conclusions and published<br />

atomic models the same year.) This principle<br />

underlies the field of structural biology, which<br />

exposes the physical and chemical structure<br />

of molecules such as proteins and DNA to<br />

explain their functions.


12<br />

The same year that Couper published his<br />

drawings, Charles Darwin and Alfred Wallace<br />

announced a new theory to explain the origins<br />

of species and their relationships to each<br />

other. Evolution quickly began to change the<br />

way research was being done in many fields.<br />

Ernst Haeckel, who had studied medicine in<br />

Berlin under Virchow and Müller and then<br />

moved to the university in Jena, had decided<br />

that he preferred the quiet life of the laboratory<br />

to dealing with sick patients. His main<br />

interest was the grand mystery of how the<br />

bodies of animals developed from single cells.<br />

As he studied this process in a variety of<br />

organisms, he noticed that the embryos of<br />

different species went through phases in<br />

which they looked remarkably similar,<br />

although they ended up quite different as<br />

adults. Evolution offered a new perspective on<br />

the problem, and Haeckel devoted himself to<br />

understanding the links between an individual<br />

animal’s life history and the long evolutionary<br />

history of its species. One of his own<br />

students, Oskar Hertwig, eventually returned<br />

to Berlin to become professor of anatomy at<br />

the university. A main theme of his work was<br />

embryonic development. He was the first person<br />

to describe the fertilization of egg cells by<br />

sperm and some of the early steps of animal<br />

development.<br />

The work of Haeckel and Hertwig underlined<br />

the fact that creatures don’t only inherit features<br />

like eyes, brains, and hair color from their<br />

parents; they also inherit the processes by<br />

which those features form in an embryo.<br />

Understanding why this was so would require<br />

a solution to one of science’s greatest mysteries:<br />

heredity. In the 1850s and 1860s that problem,<br />

too, was being solved – in an unlikely<br />

place, a quiet abbey in Moravia. Here a monk<br />

named Gregor Mendel was carrying out<br />

breeding experiments with mice until a visiting<br />

bishop found the animals mating in his<br />

room. The mice had to go. Mendel moved to<br />

the monastery garden to work with peas<br />

and other plants. No one objected to that<br />

because, as Mendel remarked ironically, “The


ishop did not understand that plants also<br />

have sex.”<br />

The monk was an excellent statistician and<br />

became one of the first to apply mathematics<br />

to a biological problem, which is echoed in the<br />

huge role that mathematics and computers<br />

now play in the modeling of biological<br />

processes and the analysis of genomes.<br />

Mendel’s painstaking experiments revealed<br />

the basic laws of heredity. His work was published<br />

in a small journal, but few people<br />

noticed – including Charles Darwin, who was<br />

carrying out experiments on peas at the same<br />

time.. Bad luck prevented Mendel from publishing<br />

further important papers on the topic.<br />

He had been corresponding with the famous<br />

Swiss botanist Karl von Nägeli, who recommended<br />

that he repeat the experiments with<br />

a plant called the hawkweed. Cross-pollinating<br />

the plant was very difficult, and no one<br />

realized that there were some unusual factors<br />

in the way it reproduced. This meant that<br />

Mendel’s laws didn’t hold up, and the monk<br />

became unsure of his results. Soon afterwards<br />

he was appointed abbot of the monastery<br />

and became immersed in administration and<br />

political squabbles. As a result, his work was<br />

almost completely forgotten for nearly four<br />

decades.<br />

The link to Berlin lies in Mendel’s rediscovery,<br />

which happened simultaneously through the<br />

work of three scientists working independently<br />

in three different countries. One of the<br />

men, Carl Correns, was studying heredity in<br />

plants at the University of Tübingen. In<br />

January 1900 Correns published an article<br />

that appeared almost simultaneously with<br />

studies by Hugo de Vries and Erich von<br />

Tschernak. Taken together, they proved that<br />

Mendel’s laws were almost universal. As well<br />

as ensuring Mendel’s place in history, the<br />

work gave a boost to Correns’ career. In 1913 he<br />

was appointed director of the new Kaiser-<br />

Wilhelm-Institut für Biologie in Berlin-<br />

Dahlem. He stayed there until his retirement,<br />

continuing to work in the laboratory but<br />

rarely publishing his work. As a result, most of<br />

it was lost during a 1945 bombing of Berlin,<br />

when his manuscripts were destroyed. In<br />

Berlin, science and history have always<br />

been deeply connected. For better and for<br />

worse.<br />

��<br />

O ne neighborhood blurs into the next.<br />

Just as Berlin’s villages have grown<br />

together to produce a cosmopolitan city, the<br />

13 Science and the city a drive to Buch<br />

villages of science have been growing together<br />

to produce modern biomedicine. Cell biology,<br />

genetics, evolution, embryology, chemistry,<br />

physics, pathology, and medicine have steadily<br />

become intertwined to produce a set of<br />

powerful concepts and tools by which today’s<br />

scientists are investigating life. The process of<br />

fusion is not yet complete, but the closer it<br />

comes, the closer we will be to understanding<br />

the causes of the greatest health problems


facing our society today and finding ways to<br />

treat them.<br />

Until very recently, the greatest worldwide<br />

threats to human health were infectious diseases.<br />

Modern sanitation, antibiotics and vaccines<br />

have changed that situation in many<br />

places – but not everywhere. Malaria, AIDS,<br />

and other infectious diseases continue to decimate<br />

huge regions of the globe, and new<br />

plagues will continue to arise and cut a swath<br />

of destruction before cures are found. But in<br />

developed and developing countries, the<br />

major killers have become “old age” conditions<br />

like cardiovascular and metabolic diseases,<br />

cancer, and neurodegenerative diseases.<br />

The culprits in these cases are partly<br />

our own genes, partly the environment.<br />

The last revolution in medicine could take<br />

advantage of the body’s own defenses in com-<br />

bating disease. Vaccines stimulate our existing<br />

immune system, which arose under the<br />

pressures of evolution. Natural selection<br />

works to protect organisms as long as they<br />

reproduce and care for their young, and viruses<br />

and bacteria attack the young, so animals<br />

developed a sophisticated immune system to<br />

protect them from such diseases. But we do<br />

not have natural defenses against today’s<br />

major health threats, which arise late in life,<br />

often as a result of normal aging processes.<br />

Fighting these diseases – as well as very complex<br />

infectious diseases like AIDS – will require<br />

a new kind of medicine that reaches into cells<br />

and reprograms the machines that drive<br />

them. In doing so, it must avoid disrupting<br />

normal operations. Most of the machines are<br />

needed throughout our lives, in many different<br />

contexts. They have to continue to func-<br />

Science and the city a drive to Buch<br />

14<br />

tion in all of those situations in spite of the<br />

changes brought about by drugs. First all of<br />

their functions have to be known, and our<br />

knowledge of what goes on in each of our<br />

cells is still far from complete.<br />

��<br />

Athorough scientific tour of Berlin would<br />

take weeks; the snapshot given in the<br />

last few pages covers just a tiny fraction of<br />

discoveries made in the 19th century, and those<br />

traditions have continued to the present day.<br />

The much fuller story has been captured in<br />

other books, including several written by the<br />

scientists of Berlin-Buch. Heinz Bielka, who<br />

worked on the campus for over four decades,<br />

has devoted his retirement writing about its<br />

history. Helmut Kettenmann, whose ongoing<br />

research at the <strong>MDC</strong> is described later in this<br />

book, has also written extensively on the


topic. The two men collaborated on a Medical<br />

History of Berlin, and Helmut co-wrote a second<br />

book called Brains of Berlin, written to<br />

accompany a 1998 exhibition by the<br />

European Neurosciences Forum. These volumes<br />

give a much fuller account of some of<br />

the personalities of Berlin science, particularly<br />

in the area of brain research.<br />

I had the Brains of Berlin book open on my lap<br />

as we drove north, alongside the elevated S-<br />

Bahn tracks of the Schönhauser Allee. Ten<br />

minutes later came the ruined round train<br />

station at Pankow-Heindersdorf, ringed by<br />

shattered windows; any intact glass was<br />

cover ed in graffiti. The city gave way to countryside,<br />

still a lush green from the warm<br />

autumn. I leafed through the pages of the<br />

book and tried to get a feeling for Berlin-<br />

Buch.<br />

A few names jumped out because of their<br />

direct connection to the campus. If anyone<br />

could be counted as the heir of Alexander von<br />

Humboldt, the universal genius, it might be<br />

Hermann von Helmholtz, whose statue<br />

stands near that of Humboldt at the university.<br />

Helmholtz was another pioneer of interdisciplinarity,<br />

combining a career in medicine<br />

with research into physics, and the link to the<br />

campus in Buch lies in his name. The <strong>MDC</strong><br />

belongs to one of the largest research infrastructures<br />

in Germany, the Helmholtz<br />

Association of National Research Orga -<br />

nizations.<br />

As early as the 1850s Helmholtz was bringing<br />

physics and medicine together in new ways.<br />

One of his early projects focused on muscles,<br />

trying to show that tissues contained enough<br />

energy to drive their activity. This countered a<br />

15<br />

Science and the city a drive to Buch<br />

Heinz Bielka<br />

very old idea that substances required the<br />

addition of some sort of special “vital force” to<br />

become alive. In Europe vitalism was a strong<br />

philosophical tradition, bordering on mysticism;<br />

it was an inspiration for Frankenstein<br />

and later a number of bad films in which mad<br />

scientists zap dead bodies with electricity to<br />

bring them back to life. Vitalism began to give<br />

way in the late 19th century when chemists<br />

managed to create organic substances out of


Jeanne Mammen: Trompetender Hahn (ca. 1943) Tempera on cardboard<br />

inorganic matter. The philosophy was buried<br />

entirely (at least in biology) with the discovery<br />

of DNA’s role in heredity and cell chemistry.<br />

But it had its function: a number of important<br />

physicists became interested in biology, hoping<br />

to discover new physical forces. They<br />

brought along a new way of looking at life<br />

that inspired an entire new generation of<br />

young people to study science.<br />

Helmholtz’s later work focused increasingly<br />

on how the brain perceives sensory information<br />

– color, motion, and sound. He made<br />

groundbreaking measurements of the velocity<br />

of impulses as they traveled through nerves<br />

and even tried to develop theories of sart and<br />

music appreciation based on what he was<br />

learning about the nervous system.<br />

Helmholtz was a gifted pianist himself. It is<br />

not unusual for scientists to cultivate artistic<br />

skills – that, too, is a tradition in Buch. Several<br />

of the scientists working there are gifted<br />

musicians or artists, and the campus is<br />

adorned by sculptures and paintings.<br />

One of those works, a Cubist portrait of the<br />

great geneticist Max Delbrück, hangs in the<br />

office of <strong>MDC</strong> Scientific Director Walter<br />

Birchmeier. It was made by Jeanne Mammen,<br />

a lifelong friend of Delbrück, who supported<br />

her and amassed a large collection of her<br />

paintings. More than a dozen of her works can<br />

be found in the “Jeanne Mammen Saal,” a conference<br />

room located in the gatehouse on the<br />

Buch campus.<br />

This small building is the first that visitors see<br />

when they arrive at the site. It’s an unusual<br />

entrance for a scientific campus, the product<br />

of another curious bit of Berlin’s scientific history.<br />

Originally city planners intended to<br />

Science and the city a drive to Buch<br />

16<br />

develop a cemetery on the land. However,<br />

most of Berlin sits on a thin layer of soil atop<br />

a marshland. Any hole that is dug quickly fills<br />

with water. That ruled out a cemetery,<br />

although the gatehouse, a gardener’s house<br />

and a chapel had already been built.<br />

��<br />

Over the past 50 years a revolution has<br />

been happening in science, mostly a<br />

quiet one, carried out in the biology departments<br />

of hospitals, universities, and independent<br />

institutes like the <strong>MDC</strong> and FMP. A<br />

look at any day’s headlines shows that the<br />

revolution is quickly expanding far beyond the<br />

laboratory. Scientists have sequenced the<br />

genomes of humans and many other animals;<br />

they have learned to clone mammals,<br />

and tried new “gene therapies” on human<br />

patients. They hope to cure disease by training


the cells of the immune system to kill tumors,<br />

or by transplanting stem cells that will regenerate<br />

damaged tissues. These projects are still<br />

in their infancy, and it may be many more<br />

years before they become standard tools in a<br />

doctor’s black bag. But most scientists expect<br />

this to happen.<br />

Science is becoming both highly interdisciplinary<br />

and highly specialized. Most research<br />

involves detailed events within cells or organisms<br />

that are hard to understand without<br />

expert knowledge. While the biomedical revolution<br />

increasingly affects us all, few people<br />

have daily contact with the world of<br />

researchers. The picture of scientists presented<br />

by the media is often unbalanced and<br />

unfair. So it is no wonder that the rapid<br />

changes have raised concerns and even fears.<br />

In this environment, institutes have a respon-<br />

17


sibility to open their doors and explain their<br />

work in a public dialog.<br />

Ideally, everyone should have the chance to<br />

visit a place like Buch and talk directly to its<br />

scientists. Each year thousands of school children<br />

get that chance through trips to the<br />

campus’ teaching facility, called the<br />

“Transparent Laboratory,” and many more<br />

visit during Berlin’s “Long Night of the<br />

Sciences.” This book has been written for<br />

those who have not yet made the trip and for<br />

those who would like to know more.<br />

��<br />

In the novel Life, a User’s Manual the French<br />

author Georges Pérec meticulously examined<br />

an apartment building from top to bottom<br />

as if it were made of glass, describing<br />

every room, every stairwell, every occupant. If<br />

we could achieve a complete understanding<br />

of one place, he implied, we would gain<br />

insights into the entire world.<br />

Pérec’s book came to mind the first time I visited<br />

the campus. The exterior of the new<br />

Center for Medical Genomics consists entirely<br />

of windows; from across the street, you can<br />

see which scientists are in their offices, which<br />

are meeting with students, and who is<br />

thumbing through the newspaper in the coffee<br />

room. The same is true of the <strong>MDC</strong> administration<br />

building. You don’t need to call to see<br />

if a colleague is in his or her office; you can<br />

simply go outside.<br />

This book does not visit nearly every laboratory<br />

of the Buch campus, nor does it explore all<br />

the stairwells (although some of them, such<br />

as the spiral staircase in the Medical<br />

Genomics building, are worth seeing). And if<br />

you were to walk through the hallways and<br />

knock on doors, you would encounter many<br />

people who do not appear here – the secretaries,<br />

the Hausmeister, the conference office,<br />

the computer group, the people who order<br />

and deliver and inventory supplies... The list is<br />

much longer. All of these people play a vital<br />

role in the life of the institutes, and without<br />

their contributions none of the stories in this<br />

book could have happened.<br />

So even in a work of this size, there are<br />

inevitably many gaps. But the intent is similar<br />

to that of Pérec. There are stories to be found<br />

everywhere. The institutes of Berlin-Buch are<br />

among thousands across the globe devoted<br />

to biomedical themes. The work being done<br />

here has some unique features because of the<br />

individuals who are doing it, the history of the<br />

campus, and the character of the institutes<br />

that have been assembled. Some of that work<br />

is among the best in the world, but the purpose<br />

of this book is not to boast or advertise.<br />

Instead, a deep look at the work being done in<br />

Buch offers a window onto this unique<br />

moment in the history of science, and a<br />

glimpse of what it is likely to produce in the<br />

future.<br />

Science and the city a drive to Buch<br />

18


Part One: The lives of nomads


Songlines and<br />

shapeshifters<br />

Most days in the early afternoon, Walter Birchmeier leaves his<br />

research lab in the Max Delbrück Center and traipses across<br />

the street past the bus stop, a statue of a blue bear, and a bronze bust<br />

of Hermann von Helmholtz. Arriving at the administration building, he’s<br />

about to enter his second life as Scientific Director of the <strong>MDC</strong>, and<br />

sometimes he sighs as he opens the door.<br />

It’s a stunningly beautiful building. The front door opens onto a cathedral-like<br />

space that rises five stories. The ceiling is reminiscent of the<br />

bottom of a ship – if the ship were made of glass – it even has large, propeller-like<br />

fans on either end. The far wall is a vast surface of white<br />

brick, decorated with tall paintings. Behind that wall is one of<br />

Germany’s most advanced facilities for the care and study of laboratory<br />

animals. Sometimes there is a faint,<br />

zoo-like scent (which you don’t notice after<br />

the first month) and the fans kick in to draw<br />

it out. A series of bridges link the offices to<br />

long flights of stairs, and colorfully<br />

painted alcoves look down from<br />

the upper floors. It would be<br />

a perfect place for theater<br />

22


Walter Birchmeier


performances; a few years ago one of the<br />

German television stations filmed part of a<br />

murder mystery here. The acoustics are magnificent,<br />

and short concerts are sometimes<br />

given here by some of the musicians on the<br />

staff.<br />

Walter is one of them, as evidenced by three<br />

tall organ pipes mounted on the wall just<br />

inside his office. He inherited the pipes from<br />

the former director, Detlev Ganten, but they<br />

are appropriate decorations for Walter’s office<br />

as well. Long ago, in what seems like another<br />

life, he studied church music. “I have the diploma<br />

here somewhere,” he says, rummaging<br />

through a stack of personal papers. “If I can<br />

find it, I’ll show it to you.”<br />

Before entering the office he has to get past<br />

his secretary, Elisabeth Kujawa-Schmeitzner,<br />

who greets him with a list of his appointments,<br />

a folder of paperwork waiting for his<br />

signature, and other business. Occasionally<br />

the sight of the list evokes another sigh. “No<br />

appointments today,” he says hopefully, but<br />

it’s a lost cause; the calendar is already full.<br />

Administrative duties are the price that has to<br />

be paid when leading a major scientific institution.<br />

One reward is getting to develop a<br />

vision for the future and steer several hundred<br />

scientists toward it. Not that they are<br />

always easy to steer.<br />

Today’s most pressing issue is money. The<br />

deadline is drawing near for three major<br />

funding applications. Every five years the<br />

<strong>MDC</strong>’s research programs have to resubmit<br />

budgets and plans, a process involving fascinating<br />

but hard discussions about how each<br />

line of research is likely to develop over the<br />

next few years, about which topics should be<br />

pursued further, and which should be<br />

dropped. It also involves a detailed scrutiny of<br />

the progress of the research groups.<br />

One of Walter’s goals as scientific director has<br />

been to push the quality of science at the<br />

<strong>MDC</strong>. That begins in his own lab. He has just<br />

come from a meeting with the head of another<br />

group who is trying to finish a manuscript<br />

that they collaborated on. They have made an<br />

important discovery: the appearance of a protein<br />

in tumor cells may indicate whether they<br />

will turn into metastases that wander<br />

through the body, a process which often<br />

transforms a benign tumor into a deadly one.<br />

“It’s a good paper,” Walter says. “It might have<br />

been better if I’d had more time to work on it...<br />

But we have to submit it for publication<br />

quickly so that we don’t get scooped.”<br />

Other groups across the world work on similar<br />

themes, collaborating and competing. So<br />

much depends on an institute’s reputation.<br />

It’s important to attract the best staff and the<br />

24<br />

brightest students, and they are drawn to the<br />

places with the best track records.<br />

��<br />

Walter’s accent in English and German<br />

gives away his Swiss origins. “When I<br />

give talks, I don’t have to tell jokes,” he says. “I<br />

just open my mouth and people laugh. They<br />

hear my voice and they think of Emil, the<br />

famous Swiss comedian.”<br />

For a scientist, independence and a tenuretrack<br />

position usually only come after study<br />

and work in several laboratories that might lie<br />

anywhere in the world. This means a move<br />

every three or four years, repeatedly uprooting<br />

the family. Walter’s history has been typically<br />

nomadic. He worked as a school teacher in<br />

Switzerland (confronting a class of 49 unruly<br />

Elisabeth Kujawa-Schmeitzner


fifth-to-eighth graders) to put himself<br />

through the university, then received his PhD<br />

in biology from Zürich University. He joined a<br />

laboratory in New York, came back to Basel,<br />

left again for San Diego in the U.S., returned to<br />

Zürich, then moved on to Tübingen and Essen.<br />

He has been in Buch since 1993, right after the<br />

<strong>MDC</strong> was founded, becoming scientific director<br />

when Detlev Ganten was asked to become<br />

head of the Charité in 2004.<br />

So migrations are a personal theme as well as<br />

a focus of his scientific work. He has always<br />

been curious about what prompts cells to<br />

leave their homes in tissues and organs and<br />

set off on voyages through the body. This happens<br />

all the time as embryos form: cells grow<br />

in sheets that slide along each other, stretching<br />

and folding to form organs; stem cells<br />

migrate to the building sites of muscle, blood<br />

vessels, and bone. By adulthood most cells<br />

have settled into a sedentary lifestyle, embed-<br />

ded in a tissue. Blood cells are the most familiar<br />

exception, but there are many others. Skin<br />

cells are born several layers below the surface<br />

and move upwards to replace relatives that<br />

die and are shed. Wounds are patched up by<br />

cells that move in from the edges. Muscle<br />

injuries suffered on the football field may heal<br />

when the body calls up generic stem cells that<br />

travel to the site of the injury and specialize to<br />

become new muscle.<br />

Cell migrations are also of intense interest<br />

because of their role in cancer. Solid, benign<br />

tumors can often be removed before they disrupt<br />

the functions of vital organs, but all too<br />

often they spawn metastatic cells which<br />

escape and start to grow again in different<br />

parts of the body. A surgeon may not be able<br />

to remove such cells because they have<br />

already escaped the tumor, or they may look<br />

just like healthy cells. If even one is left<br />

behind, the cancer may reappear and spread.<br />

25<br />

Part one: The live of nomads<br />

A metastatic cell can’t be distinguished with<br />

the eye, but its unusual behavior means that<br />

its chemistry must be different than that of<br />

other cells. Identifying any unique molecules<br />

it contains might give a clue about why the<br />

cell sets off on a dangerous path, and might<br />

even reveal how it could be stopped.<br />

But that’s a long-term prospect. The field of<br />

molecular medicine straddles a frustrating


gap between knowledge and cures, Walter<br />

says. Knowing that something is broken doesn’t<br />

mean you can fix it. But it’s where you have<br />

to start.<br />

��<br />

In his book The Songlines, world traveler<br />

Bruce Chatwin recounts his experiences<br />

among the nomadic natives of Australia. The<br />

indigenous people of Australia believed that<br />

the world was born in the Dreamtime, a period<br />

in which formless spirit-beings assumed<br />

the shapes of plants, animals and human<br />

beings. Each feature of today’s landscape has<br />

a mythical and spiritual meaning which is<br />

recorded in stories and songs. Children may<br />

be born anywhere during the wanderings of<br />

the family, and the birthplace takes a special<br />

significance in their lives. Each child learns the<br />

traditions associated with that place.<br />

Part one: The live of nomads<br />

26<br />

Everyone’s collection is unique because new<br />

stories are added and old ones are reshaped<br />

by time and the telling.<br />

A cell also arises from a vaguer form, and its<br />

identity is defined by the landscape in which<br />

it is born. By tasting a particular neighborhood’s<br />

molecules, it learns what to become.<br />

The tasting is done by proteins called receptors<br />

which float on the surface of the cell.<br />

Each type of cell has a unique combination of


eceptors which sensitize it to different flavors.<br />

And each environment tastes different<br />

because it contains a unique set of stimuli,<br />

called ligands. Combinations of receptors and<br />

ligands give the cell its identity.<br />

Tasting is the first step in a process that tells<br />

a cell what to become and how to behave.<br />

Receptors send information into the cell via<br />

signaling pathways. They work a bit like the<br />

telephone list that my children’s teachers use<br />

to get information to parents. The teacher<br />

calls one parent, who calls me, and then I call<br />

the next parent on the list. Each bit of news<br />

reaches the same parents in the same order.<br />

In the cell information is passed from a receptor<br />

(teacher) to proteins (parents) using<br />

chemistry rather than telephone calls. When a<br />

signal triggers the same proteins in the same<br />

order it is called a pathway. Just as there are<br />

many receptors, there are many pathways.<br />

27 Part one: The live of nomads<br />

Signaling pathways cope with the same types<br />

of confusing situations that sometimes happen<br />

with telephone lists. One teacher may<br />

need to send out information to several classes.<br />

(Some receptors can trigger multiple pathways.)<br />

I have three children, so I am on three<br />

telephone lists. (One protein may belong to<br />

several pathways.) Whom I call depends on<br />

who called me. (Depending on where a signal<br />

comes from, a protein may react in different


ways.) If I can’t get in touch with the next parent,<br />

the chain may be interrupted (a broken or<br />

missing protein may block the signal). And as<br />

a child may use the list to make prank phone<br />

calls, viruses often use receptors to talk their<br />

way into cells.<br />

Eventually the information makes its way to<br />

the genes in the cell nucleus, like customer<br />

orders arriving at the kitchen of a restaurant.<br />

DNA contains the recipes for other molecules<br />

needed by the cell (RNAs and proteins).<br />

Information from signaling pathways switches<br />

on some genes and switches off others.<br />

This changes the set of molecules present in a<br />

cell, the way the arrival of new customers<br />

changes the food on restaurant tables.<br />

Tracking those orders – identifying what<br />

receptor sends a certain signal, what proteins<br />

pass it along, and what genes are affected by<br />

the information – is the subject of a great deal<br />

of today’s biological research.<br />

Signaling pathways are involved in nearly<br />

everything that happens in the cell – good<br />

and bad. They give positional information<br />

that helps stem cells develop, and they set<br />

the cell’s clock that tells it when to divide. But<br />

pathways can also break down. A protein may<br />

go missing, or it may get stuck in a “transmitting”<br />

mode where it broadcasts signals all the<br />

time, even when it hasn’t received the call<br />

from the receptor. If it tells the cell to divide<br />

all the time, rather than taking breaks, the<br />

result may be cancer.<br />

��<br />

In adult animals, sedentary cells sometimes<br />

get a signal to run sprints, such as when a<br />

wound needs to be healed. “The skin is the<br />

body’s first line of defense and its integrity is<br />

crucial,” Walter says. “Cuts need to be sealed<br />

off fast to prevent bleeding and infections.<br />

Cells move into the wound area and build<br />

new skin, after which things have to return to<br />

28<br />

normal. So we thought this might be a good<br />

system to use to look for the factors that<br />

switch migrations on and off.”<br />

In 2007 Walter’s group made some important<br />

discoveries about the signaling pathway that<br />

controls these effects. The story involved<br />

some old friends (molecules that Walter has<br />

been studying for over 15 years) and an even<br />

closer friend – his wife. Carmen Birchmeier is<br />

a coordinator of the <strong>MDC</strong>’s neuroscience program<br />

and runs a large lab with wide-ranging<br />

research interests. Sometimes the work of the<br />

two Birchmeier labs overlaps.<br />

Both Carmen and Walter have been studying<br />

a signaling pathway that begins with a protein<br />

called Hepatocyte growth factor/scatter<br />

factor, or HGF/SF for short. The name is cumbersome<br />

but revealing. A growth factor is<br />

something that stimulates cells to divide (in<br />

this case, liver cells called hepatocytes), and a<br />

scatter factor prompts them to move. The scientist<br />

who discovers a new molecule gets to<br />

name it, usually picking something that<br />

reflects its function in an organism. (Often<br />

having a bit of fun in the process – there are<br />

genes called sonic hedgehog, lunatic fringe,<br />

teashirt, dreadlocks, rolling stones, and Swiss<br />

cheese.) A lot of molecules have double or<br />

hyphenated names because different scientists<br />

start working on the same gene in different<br />

organisms, and by the time people realize<br />

they are talking about the same thing, the<br />

two names have become entrenched in the<br />

literature.<br />

Carmen started working with HGF/SF to learn<br />

how it influences the development of tissues<br />

in the embryo, particularly the liver and muscles.<br />

But when she developed strains of mice<br />

to study its activity, she found several other<br />

systems of the body in which it plays a role.<br />

Walter has mainly been interested in aspects<br />

relevant to the behavior of cells. Carmen,<br />

studying large-scale processes at work in<br />

whole animals, was mostly thought of by her<br />

colleagues as a developmental biologist and<br />

Walter was more focused on cell biology. In


the meantime it has become much easier to<br />

slide from one scale of things to another, so<br />

the fields (and a few of the people working in<br />

them) are undergoing a sort of marriage.<br />

It has been known for about 20 years that<br />

HGF/SF prompts cells to divide and move.<br />

When the protein is fed to laboratory cultures<br />

of epithelial cells (which make up the skin and<br />

the surfaces of many organs), they start to<br />

replicate. Tumor cells raised in a dish in the lab<br />

release each other and crawl away.<br />

In the body HGF/SF is tasted by a receptor<br />

called Met. Years of work by the labs of<br />

Carmen and Walter have revealed some of the<br />

molecules that transmit the information<br />

from Met on to genes. Carmen has linked this<br />

signaling pathway to several important<br />

events in the body. “The correct development<br />

of the liver and muscle depends on Met signals,”<br />

she says. “They are needed to build<br />

these tissues in the first place and to rebuild<br />

them when they have become damaged.”<br />

Usually HGF/SF and Met are made by cells in<br />

different places, which is one way to keep the<br />

pathway under control – the signal and its<br />

receiver only meet up when conditions are<br />

right. But cells in the heart, kidneys, lungs, and<br />

skin begin producing high levels of both proteins<br />

when there has been an injury. As if,<br />

Walter says, you work to improve your<br />

chances of transmitting a radio signal by<br />

increasing the power of the transmission and<br />

installing a larger antenna. The times and<br />

places where the pathway was active made<br />

him think that the proteins might be a master<br />

control switch for healing. They might be con-<br />

29 Part one: The live of nomads<br />

trolling when cells divided and when they<br />

moved into the wound area. Walter knew of a<br />

way to find out, but first he needed a new<br />

type of mouse.<br />

��<br />

Carmen’s group makes extensive use of<br />

the <strong>MDC</strong>’s animal house, a state-of-theart<br />

center for the care and study of mice, rats,<br />

rabbits, fish, and a few more exotic organisms.<br />

Animals have been key figures in biological<br />

and medical progress and their role is increasing.<br />

With the discovery that genes caused some<br />

diseases and contributed to others, scientists<br />

needed a way to study their effects over an<br />

organism’s entire lifetime. A great deal had<br />

already been learned by studying mutations


that occurred naturally in laboratory organisms<br />

such as the fly. But this meant waiting<br />

for random events to change random genes.<br />

Experiments could only be conducted on<br />

mutations that had already been discovered;<br />

there was no way to choose a gene that you<br />

were interested in and change it.<br />

In the 1930s the situation began to change<br />

thanks in part to events on the Berlin-Buch<br />

campus. The American Hermann Muller had<br />

discovered that X-rays and chemical agents<br />

caused mutations by damaging genes. The<br />

techniques still didn’t permit scientists to<br />

interfere with a particular molecule, but they<br />

provided plenty of new animals to study.<br />

Muller was immersed in his X-ray studies<br />

when he arrived in Berlin-Buch in 1932 to work<br />

with Nikolay Timofeeff-Ressovsky. At that time<br />

the campus was home to the Brain Research<br />

Institute of the Kaiser Wilhelm Society, headed<br />

by Oskar Vogt. He was half of another scientific<br />

couple: his wife Cécile was an excellent<br />

researcher and wrote or coauthored many of<br />

his scientific papers.<br />

It was a period of strong ties between<br />

research in Berlin and the Soviet Union, which<br />

was emerging from a phase of cultural and<br />

scientific isolationism. Vogt had spent two<br />

years in Moscow helping the Russians set up a<br />

Brain Institute – whose main occupation was<br />

to study Lenin’s brain. Vogt trained scientists<br />

in specialized techniques of brain dissection,<br />

worked with the brain in Moscow, and evaluated<br />

many of his findings in Buch.<br />

In Moscow Vogt had met Timofeeff-Ressovsky<br />

and invited him to Buch to head the<br />

Department of Genetics. When political difficulties<br />

forced Vogt to leave Buch in the mid-<br />

1930s, this department was transformed into<br />

a separate institute devoted to Genetics and<br />

Biophysics and Timofeeff-Ressovsky was<br />

named its director.<br />

Hermann Muller’s visit to Buch coincided with<br />

another encounter that Timofeeff-Ressovsky<br />

described many years later:<br />

Part one: The live of nomads<br />

30<br />

In the beginning of the 1930s I became<br />

a friend of Max Delbrück and, so to<br />

say, involved him in our work. He was<br />

a pure theoretical physicist, the pupil of<br />

Max Born and Niels Bohr. Actually I<br />

won him over to theoretical biology.<br />

Now he is a major virologist and theoretical<br />

biologist in America, a Nobel<br />

Prize winner, and on the whole a very<br />

remarkable person. At that time he was<br />

a young man, and as are all great theorists,<br />

was somewhat insolent, but this is<br />

excusable. We also treated him<br />

insolently, and he acquired our manners<br />

very quickly and became quite an<br />

acceptable young man. Henceforth he<br />

was invited to our Buch group.<br />

Timofeeff-Ressovsky and Delbrück collaborated<br />

on one of the most important papers<br />

about genes written in the first half of the<br />

20th century. Called the “Green Pamphlet”<br />

because it was reprinted in a green cover, the<br />

document assembled what was known about<br />

genes and made some suggestions about<br />

how to find them. Twenty years before James<br />

Watson and Francis Crick solved the riddle of<br />

what genes were made of, the scientists of<br />

Berlin gave an extremely accurate description<br />

of how genes would have to be organized to<br />

behave the way they do. The Green Pamphlet<br />

drew a lot of physicists to the question; eventually<br />

their study of DNA using X-rays gave<br />

Watson and Crick information that they had<br />

to have to understand the molecule.<br />

Within a few years of the discovery of the<br />

genetic code, scientists found proteins in bacteria<br />

and other types of cells that could cut<br />

DNA from chromosomes and paste them<br />

back in again. The molecules were adapted so<br />

that they worked the same way in other<br />

species. Now instead of breeding huge numbers<br />

of animals and waiting for mutations to<br />

happen by chance, scientists could pick a gene


and remove it (knocking it out), transplant it<br />

between species, or change it in other ways.<br />

This was a huge step in the study of genetic<br />

diseases. If a human disease was the result of<br />

a mutation in a gene, researchers could<br />

remove the mouse version of the gene and<br />

see whether the mouse developed a similar<br />

problem. The same approach is being taken<br />

with cancer and other health problems that<br />

are linked to defects in genes.<br />

Genetic engineering can now be carried out in<br />

a wide range of organisms, and scientists in<br />

Buch and elsewhere are continuing to the<br />

tools so that they can be used in new animals.<br />

The unique features of other species can provide<br />

valuable insights into human problems.<br />

Carmen’s work with mice has provided<br />

insights into neurodegeneration, muscle diseases,<br />

and other health problems. A tiny<br />

worm called planarium, studied in Nikolaus<br />

Rajewsky’s lab, may help us understand how<br />

wounds heal – or don’t – in humans. A blind<br />

and deaf creature from East Africa, the naked<br />

mole rat, is helping Gary Lewin understand<br />

how our nervous system detects and transmits<br />

sensations of touch and pain.<br />

But those stories come later.<br />

��<br />

When Carmen’s lab knocked out the<br />

gene for Met, the embryos experienced<br />

fatal problems. The liver failed to develop<br />

and many other tissues did not form in the<br />

normal way; the mice died before birth. Such<br />

animals couldn’t provide insights into Met’s<br />

role in later stages of development or<br />

tissue regeneration, so the lab had to<br />

use another method of shutting<br />

down genes called a conditional<br />

knockout. The technique<br />

These images show how cells respond to a skin wound. In the upper row, cells have functioning c-Met;<br />

they reproduce quickly and begin moving into the wound area (right side). After 48 hours, a large<br />

number of cells have moved into the area. In the lower row, most of the cells do not have c-Met. Cells<br />

respond much more slowly, and only those with c-Met move into the region of the wound.<br />

was developed by the laboratory of Klaus<br />

Rajewsky (more on him in the story “Playing<br />

the piano of planaria”), then at the University<br />

of Cologne. The method involves attaching a<br />

gene to a switch that allows scientists to shut<br />

it down at precise times and places in the animal.<br />

This usually requires mating two strains<br />

of mice – one outfitted with the switch, and<br />

another with a trigger that can shut it down<br />

only in muscle, the skin, or another specific<br />

tissue.<br />

Carmen’s lab developed an animal with a<br />

switchable form of Met and began using it to<br />

investigate regeneration. The same mouse<br />

could be used to investigate the skin, so<br />

Walter looked for someone to carry out the<br />

project. One of the group’s PhD students,<br />

Jolanta Chmielowiec, had been doing similar<br />

work with another strain of mouse but hadn’t<br />

gotten anywhere. She had the skills necessary<br />

to work on Met.<br />

I met Jolanta as she was packing up to leave<br />

for a new position at the Harvard Medical<br />

31<br />

Part one: The live of nomads<br />

School. She is a blond, thoughtful woman<br />

who often pauses to think before she speaks.<br />

Making the decision to leave was hard, she<br />

said. “The projects are going well now – finally.<br />

It wasn’t always that way. Now that things<br />

are moving it would be easy to stay, but the<br />

longer you do, the harder it gets to move. I’m<br />

not ready to settle down in one place yet.” She<br />

admitted she was a little nervous about<br />

Harvard. “But working with the two groups of<br />

Walter and Carmen Birchmeier has been the<br />

best preparation you could get.”<br />

Jolanta arrived from Poland several years ago<br />

after finishing her master’s degree on the<br />

topic of cancer. Her work had involved comparing<br />

the DNA of lifetime smokers with nonsmokers.<br />

The topic was interesting, she<br />

admits, but it was all carried out in the test<br />

tube and the computer; she wanted to work<br />

with animals. Buch offered that chance.<br />

Walter proposed two projects, and she decided<br />

to try to develop a strain of mouse with a<br />

conditional knockout of a gene called Gab1.


But two years later the work had stalled, and<br />

she desperately needed results for her PhD<br />

dissertation. So it was a good time to take on<br />

a new project.<br />

Her first task was get to mice with a version of<br />

Met that could be shut down only in the skin.<br />

Jolanta did this by mating Carmen’s mice with<br />

another strain, developed earlier in Walter’s<br />

lab, whose skin cells (called keratinocytes) contained<br />

a trigger that would shut down the<br />

gene. A test of the newborn mice showed that<br />

some had inherited both components. At first<br />

Jolanta hoped the hard work was now finished<br />

– but that didn’t turn out to be the case.<br />

“You always hope things will go easy,” Jolanta<br />

says. “I looked at the mice at various stages of<br />

development – before birth and at other<br />

stages of their lives – hoping to find something<br />

wrong with the skin. But there was<br />

nothing there. It’s the worst result you can<br />

have, to do months and months of work only<br />

to find out that your gene doesn’t do what<br />

you expected it to. Maybe Met didn’t have any<br />

effect on the skin at all.”<br />

Time pressure was on to get results – she didn’t<br />

want to be a PhD student forever. She still<br />

had to look at wound healing, so she turned<br />

to the scratch test, where the scientist makes<br />

a small wound through several layers of the<br />

mouse’s skin.<br />

She closely followed what happened to the<br />

injury over several days. “In both normal and<br />

mutant mice, the wounds healed,” Jolanta<br />

says. “I thought, ‘Here we go again,’ another<br />

case where we wouldn’t find anything. That<br />

was a really rough time. But in fact there was<br />

a small difference between the two types of<br />

animals. Mice with no Met in the skin healed<br />

more slowly.”<br />

The findings didn’t seem very promising, but<br />

she had a discussion with Walter in which<br />

they toyed with the idea of writing a paper<br />

based on them anyway. Instead, they decided<br />

to take a closer look at the wounds under the<br />

microscope to try to find the cause of slow<br />

Part one: The live of nomads<br />

32<br />

healing in the mutant mice. It meant going<br />

back to Carmen’s lab for help with the<br />

microscopy; that turned out to be a good<br />

decision.<br />

The scientists removed small patches of skin<br />

from the scratch site using a laser. Jolanta’s<br />

first discovery was that fewer cells were moving<br />

into the wound. Next she took samples<br />

from the cells to see if there were any obvious<br />

differences. After that experiment, thing suddenly<br />

made sense.<br />

“In a conditional knockout, you rarely are able<br />

to remove 100 percent of a molecule,” she<br />

says. “A few cells slip through and produce it.<br />

In this case, a few of the skin cells in the<br />

mouse were still making Met. Those were the<br />

ones that migrated into the region; they were<br />

the only one to participate in healing. Cells<br />

without the gene didn’t – either they were<br />

unable to hear the signal to move, or they<br />

couldn’t respond to it.”<br />

Again with the help of Carmen’s lab, she<br />

attached fluorescent markers to Met, which<br />

let her track the cells that produced it as they<br />

divided and moved. Some of the wound repair<br />

comes from cells at the very edge of the<br />

wound, but they need reinforcements. A distress<br />

signal is heard by keratinocytes farther<br />

away which divide quickly and then migrate.<br />

This process is also influenced by the loss of<br />

Met. The mouse has fewer cells that are able<br />

to participate in healing, which is why the<br />

process takes longer to get started. But once it<br />

does, the few cells with Met divide more and<br />

for a longer time. After a few days they make<br />

up for what has been lost.<br />

This shows that the body must have sophisticated<br />

systems to start healing, to check on its<br />

progress, and then shut it down once the<br />

injury has been repaired. “Losing control over<br />

any of these processes is potentially very dangerous,”<br />

Walter says. “It’s an important motivation<br />

for our work. Healing is a local event in<br />

which cell division and migrations take place<br />

in a highly controlled way. The hallmark of<br />

cancer is a disruption of those two processes.


Hook-like cadherin proteins (left) allow cells to snag onto each other. When Walter’s<br />

group interfered with cadherins using antibodies, or interfered with β-catenin signaling,<br />

cells released each other and began to migrate.<br />

In some cases that might be due to misbehavior<br />

by the HGF/SF signal or the Met receptor<br />

or other parts of the pathway. It may explain<br />

why we find high levels of HGF/SF and Met –<br />

a sign that the pathway is switched on – in<br />

some types of tumors as well as wounds.”<br />

If so, he says, the system might be especially<br />

likely to break down at sites of injuries. Walter<br />

recalls the hypothesis of the famous Berliner<br />

Rudolf Virchow, who long ago observed a connection<br />

between cancer and inflammations<br />

in his patients. Jolanta’s project provides a<br />

new link. “In a chronic situation where there is<br />

a long-term wound or injury, the system has<br />

to be switched on and off many times,” Walter<br />

says. “Cells might become confused by the<br />

signals, or they might forget how to switch<br />

the pathway off again.”<br />

��<br />

Most of Walter’s work has been devoted<br />

to another pathway, one which starts<br />

with a molecule called Wnt. Receiving a Wnt<br />

signal sets off a chain of interactions that<br />

change gene activity and the cell’s behavior. In<br />

the meantime the pathway, which involves<br />

over 100 molecules, has been recognized as<br />

one of the most ancient and important signaling<br />

systems in animals.<br />

The paths of Walter and Wnt first crossed 25<br />

years ago, by accident, when he was working<br />

at the Max Planck Institute in Tübingen. He<br />

was using cell cultures to try to understand<br />

why cells sometimes broke free of their tissues<br />

to become nomads.<br />

“We were using cultures of <strong>MDC</strong>K cells,” he<br />

says. “These come from the surface layer of a<br />

dog kidney. In cell cultures they form an<br />

epithelium, a skin-like layer, which is one of<br />

the things that makes them interesting in all<br />

Jeanne Mammen: Damengespräch (ca. 1949) oil on cardboard<br />

33 Part one: The live of nomads<br />

kinds of experiments. We were trying to identify<br />

proteins on their surfaces that tied them<br />

to their neighbors.”<br />

One way to find them might be to block some<br />

of those proteins and see what happened, he<br />

says. You might be able to do that with antibodies,<br />

which work by gluing themselves to<br />

proteins. In doing so, they might interfere<br />

with the connections between cells. To get<br />

antibodies against the proteins on <strong>MDC</strong>K<br />

cells, Walter and his collegues injected the<br />

cells into mice. The mouse immune system<br />

recognized that the cells were foreign and<br />

began producing antibodies against molecules<br />

on their surfaces. Walter and his colleagues<br />

harvested the antibodies, purified<br />

them, and began putting them into the cell<br />

cultures one at a time. If they found the right<br />

one, he hoped, it would make the <strong>MDC</strong>K cells<br />

lose their grip on each other. That’s what happened<br />

– one antibody broke the connections.


The cells changed shape and begin crawling<br />

around.<br />

The antibody was binding to a surface molecule<br />

called E-cadherin. “If you’ve ever seen a<br />

microscope picture of Velcro, that’s a bit what<br />

cadherins look like,” Walter says. “Each of these<br />

proteins is a long fiber with a loop on the end.<br />

They stretch away from the cell and become<br />

entangled with identical cadherins on a neighboring<br />

cell. There are lots of different types, Ncadherins,<br />

T-cadherins... each can only hook<br />

onto an identical kind on the neighbor. It’s one<br />

of the things that make specific types of cells<br />

stick onto each other without gluing themselves<br />

to every other tissue in the body.”<br />

The antibodies were somehow stopping cadherins<br />

on one side from snagging those on<br />

the other. It was an exciting finding. What was<br />

happening in cell cultures mimicked something<br />

that occurred all the time in embryos –<br />

and in cancer. The group had transformed<br />

sedentary cells into nomads. Maybe E-cadherin<br />

and the same mechanism were responsible<br />

for this switch in behavior in animals,<br />

too.<br />

One way to find out would be to check cells<br />

from epithelial cancers, taken from patients,<br />

so postdoc Jürgen Behrens and other members<br />

of Walter’s lab went out and bought 20<br />

to 30 cell lines. The lab grew the cells in cul-<br />

Part one: The live of nomads<br />

34<br />

tures and used their antibody to look for Ecadherin.<br />

“Some of the tumors grew in sheets,<br />

like epithelia, and others were mobile,” Walter<br />

says. “There was a direct correspondance. If<br />

there was E-cadherin on the surface, they<br />

formed an epithelium. If there wasn’t, they<br />

crawled around.” If the scientists treated the<br />

epithelial type with the antibody, the “skins”<br />

broke up and the cells began to invade surrounding<br />

regions.<br />

There were two explanations. “Either the antibodies<br />

were directly ‘gumming up’ the hooks<br />

somehow,” Walter says, “or they were sending<br />

a signal into the cell that changed its architecture<br />

and shape. Or maybe both mechanisms<br />

Manipulating Wnt signaling in cultures of <strong>MDC</strong>K cells causes them to leave their sheet-like structure (upper left) and begin crawling over each other (lower right).<br />

Zwei Figuren, graphisch (ca. 1951) oil on cardboard<br />

Krankes Kind (ca. 1946/7) oil on cardboard


Kleiner Kopf (ca. 1945)<br />

were at work. We couldn’t immediately tell.<br />

But we were also excited about the method –<br />

we had shown that you could use antibodies<br />

to study and manipulate the behavior of cancer<br />

cells in cell cultures.”<br />

Soon Walter and his lab had answered the<br />

question. Their antibody was indeed interfering<br />

with E-cadherin, but also with one of the<br />

cell’s signaling systems. The switch between<br />

adhesion and migration was really being controlled<br />

from inside the cell, just below the<br />

membrane. Each E-cadherin has a tail that<br />

passes through the membrane to the cell<br />

interior. There it is affixed to other molecules,<br />

including a protein called β-catenin. If this<br />

molecule is released there is no anchor, and<br />

cells can no longer hold on to each other.<br />

“With Jürgen and Jörg Huelsken, a new PhD<br />

student, we did all kinds of experiments<br />

mutating β-catenin and interfering with the<br />

signals it received,” Walter says. “It was acting<br />

as a switchboard for cell adhesion and migrations<br />

in cancer, and we were beginning to look<br />

at its role in migrations in the embryo.<br />

Needless to say we were ecstatic.”<br />

As it turned out, they were only holding the<br />

tail of the tiger.<br />

��<br />

Asongline, in the world of the Aboriginal<br />

Australians, is a chain of stories that are<br />

sung to retrace the routes traveled by ancestral<br />

gods as they created the world. Bruce<br />

Chatwin called them “the labyrinth of invisible<br />

pathways which meander all over<br />

Australia.” The elements of the songs are<br />

places, and by stringing stories together and<br />

singing them in a certain order, a tribe can<br />

navigate across the entire continent.<br />

Important geographical features are intersections<br />

in many of these routes, and they figure<br />

in a lot of different songs.<br />

β-catenin was a prominent feature in the<br />

songlines of cellular information pathways,<br />

but this fact wasn’t immediately obvious.<br />

“Originally we were thinking only in terms of<br />

the adhesion between cells,” Walter says. “We<br />

were thinking of it as a structural element, an<br />

anchor or a staple, holding onto things.”<br />

But more labs had become interested in the<br />

molecule. On the one hand that was good;<br />

35 Part one: The live of nomads<br />

information about β-catenin was coming in<br />

quickly. On the other hand it meant more<br />

pressure to keep from being scooped. In 1993<br />

Walter was sitting down with Jörg, trying to<br />

write a paper about the latest experiments,<br />

when they heard about an important new<br />

discovery. The labs of three scientists in the<br />

USA – Kenneth Kinzler, Bert Vogelstein, and<br />

Paul Polakis – had learned that β-catenin<br />

docked onto another protein called APC.<br />

That was interesting because APC had been<br />

identified as a tumor suppressor gene. The<br />

healthy forms of such molecules keep tumors<br />

in check, but if they undergo mutations, the<br />

cell loses control of an important process and<br />

the result is a tumor. Aberrant forms of tumor<br />

suppressors are often found in the tissues of<br />

cancer patients.<br />

“Jörg told me, ‘This is the best thing that could<br />

have happened to us,’” Walter says. “Tumor<br />

suppression was a new concept at the time.<br />

We knew of mutations in other molecules<br />

that caused cancer, but they were usually<br />

thought of as tumor promoters – the defects<br />

made them active; they were overstimulating


cells, sending them the wrong signals at the<br />

wrong time. Tumor suppressors were more<br />

like security guards, who needed to be around<br />

to keep the status quo.”<br />

Scientists knew that mutations in APC could<br />

cause cancer, but since APC’s functions in<br />

healthy cells were unknown, the information<br />

didn’t mean very much. That began to change<br />

through new findings, which Walter was<br />

keeping his eye on as he moved his group to<br />

Berlin-Buch. Jürgen had been looking for additional<br />

binding partners for β-catenin using a<br />

method called a yeast two-hybrid screen<br />

(described in “The pleasures and powers of<br />

green tea”), and came up with another protein<br />

called Lef. Hans Clevers’ group at the<br />

University of Utrecht discovered that<br />

β-catenin also docked onto Tcf, a related<br />

protein.<br />

Lef belongs to a family of molecules which are<br />

transcription factors – they control whether<br />

genes are active or not. Transcription factors<br />

are powerful proteins that affect important<br />

aspects of cells’ behavior by changing the set<br />

of genes that are active and thus the chemistry<br />

of the cell. They are the final links in signaling<br />

pathways, transmitting information to<br />

the cell’s executive decision-making machinery<br />

in the nucleus. When Jürgen examined<br />

how the two molecules fit together, he discovered<br />

that Lef alone couldn’t activate genes<br />

– to do so it needed help from a particular<br />

module of the β-catenin molecule, a “gene<br />

activation domain.”<br />

“In retrospect this shouldn’t have been that<br />

much of a surprise,” Walter says. “We knew<br />

that there was a form of β-catenin found in<br />

the fruit fly, and we knew that it also entered<br />

the nucleus. But we didn’t know what it was<br />

doing there. Suddenly we had connected it to<br />

transcription factors. Such factors often have<br />

to work in groups to activate genes, so it<br />

immediately suggested a completely new<br />

function for β-catenin. We had been focusing<br />

on the cell adhesion side of β-catenin. It was<br />

like being at a magic show and watching the<br />

wrong hand. Now we knew we had to start<br />

looking at the other hand, too.”<br />

This completely changed the status of βcatenin,<br />

Walter says. “It had gone from a sort<br />

of cellular ‘staple’ to an important gene-activating<br />

device,” he says. “It was now an important<br />

link in cell signaling pathways.”<br />

Scientists had already identified the main signal<br />

that controlled whether β-catenin headed<br />

for the nucleus. That originated with a small<br />

protein called Wnt, a signal known to control<br />

developmental processes in embryos. “Things<br />

kept getting more exciting,” Walter says. “The<br />

molecule we were working on was now implicated<br />

in three processes: cell adhesion, cancer,<br />

and development.”<br />

Studies in the fruit fly and other organisms had<br />

shown that disrupting Wnt signals could have<br />

serious effects on embryonic development. But<br />

experiments kept turning up new kinds of<br />

Wnts – the human genome alone contains 19<br />

closely related forms of the proteins. Why so<br />

many? What did they do in the body’s tissues?<br />

How were the signals related to cancer?<br />

Most Wnt signals influence β-catenin on their<br />

way to genes, so the protein gave Walter a<br />

good tool to find answers. That has been his<br />

major preoccupation over the last 25 years.<br />

��<br />

Discoveries started to come fast. Jürgen’s<br />

fishing expedition had turned up<br />

another binding partner for β-catenin, a protein<br />

called conductin. By docking onto both<br />

APC and β-catenin, it played a major role in<br />

the stability of β-catenin, which is controlled<br />

by Wnt signals.<br />

“By the end of the 1990s we had helped<br />

establish that how much β-catenin is free in<br />

the cell is central to Wnt’s effects,” Walter says.<br />

“Labs everywhere were turning up new binding<br />

partners for β-catenin, and we were all<br />

working hard to figure out what these different<br />

combinations were doing – what genes<br />

were being activated, and what that meant<br />

Part one: The live of nomads<br />

36<br />

for the embryo. Sometimes those effects were<br />

pretty startling, as you can see when the system<br />

is perturbed.”<br />

He isn’t exaggerating – too much Wnt in the<br />

wrong place, for example, causes a mouse or<br />

frog embryo to develop two heads. With too<br />

little signaling, there is no head at all. That<br />

made sense because it was becoming increasingly<br />

clear that locations and amounts of<br />

Wnt/β-catenin signaling were helping to set<br />

up the major directional signals in the early<br />

animal body. One of the earliest things that<br />

happens in the early embryo is a decision<br />

about which end is the head and which is the<br />

tail, which side is the front and which the<br />

back. Wnt signals are like surveyors, laying<br />

down the coordinates.<br />

Context was everything, Walter says. “It wasn’t<br />

as if everything sends β-catenin into the<br />

nucleus. In fact, frequently just the opposite<br />

happens. If it links up to Axin and APC and<br />

another protein called GSK, then the cell<br />

destroys it.<br />

“What we had understood by 1998 was that<br />

Wnt signals prevent it from being linked to<br />

these molecules. That means the Wnt signal<br />

stops β-catenin from being destroyed. Then it<br />

can enter the nucleus and team up with Tcf<br />

and Lef.”<br />

Tcf/Lef normally sit on DNA and blocks the<br />

use of nearby genes. Since some of those<br />

genes trigger cycles of cell division, they need<br />

to be shut off most of the time – especially in<br />

differentiated cells that shouldn’t divide any<br />

more. But if β-catenin gets to Tcf, the brake is<br />

released and the genes get switched on.<br />

“If that happens at the right time, when the<br />

cell should divide, fine,” Walter says. “If it happens<br />

at other times, you can get a tumor. So a<br />

Wnt signal at the wrong time can cause cancer<br />

by preventing the destruction of βcatenin.<br />

And this model explains some other<br />

causes of cancer. Anything that stops βcatenin<br />

from binding to APC, actin, or GSK can<br />

allow it to survive and do damage.”


Fontaine de Vaucluse (ca. 1955) oil on cardboard<br />

��<br />

Wnt and β-catenin are found in all animals,<br />

in all sorts of tissues. “What we<br />

started looking at as a simple clamp for cell<br />

adhesion has turned out to be an all-purpose<br />

tool for building bodies,” Walter says.<br />

Over the past decade the lab has shown how,<br />

along with a handful of other powerful signals,<br />

Wnt directs the fates of cells and sculpts<br />

tissues. “There are five or six really major pathways<br />

like Wnt that nature uses over and over,<br />

in different animals and tissues. The body produces<br />

a variety of cells, tissues and organs<br />

because of subtle differences in when pathways<br />

get switched on in different places, and<br />

they are activated in different sequences.”<br />

β-catenin and Wnt are so crucial to the<br />

growth and structure of the body that their<br />

roles could only be understood through the<br />

development of conditional knock-out animals.<br />

With the help of Carmen, Walter’s lab<br />

made animals with a switch attached to βcatenin.<br />

They mated the animals with other<br />

strains containing triggers that became<br />

active in specific tissues and cell types. This<br />

has given them the tools to begin a systematic<br />

study of Wnt signaling throughout the<br />

body.<br />

In each tissue the system works differently,<br />

but ten years of these studies have given<br />

Walter a feeling for the special nature of Wnt<br />

signaling. “It is usually not the pathway that<br />

gets switched on first in a major building project,”<br />

he says. “It doesn’t, for example, trigger<br />

early cells to become precursors for all types<br />

of blood or muscle; it doesn’t start the construction<br />

of a limb. Instead, it controls finer<br />

decisions. In the skin, for example, we found<br />

that it tells stem cells to develop into hair fol-<br />

37 Part one: The live of nomads<br />

licles, but not the skin itself. It controls the<br />

development of two types of cells – out of a<br />

dozen – that develop into cells of the brain<br />

and spine.”<br />

Wnt steps in when the major foundations of a<br />

tissue have been laid. The lab’s recent study of<br />

heart development provides a good example.<br />

The signals that shape the heart have been<br />

tricky to unravel. It’s the first organ to form in<br />

vertebrates, and it can only take shape if different<br />

types of cells migrate to the right<br />

places in a coordinated way. Wnt plays a role<br />

in that process, Walter says, but there have<br />

been conflicting accounts of its role. “In cell<br />

cultures from frog and chick embryos, Wnt<br />

activity has to be very low in order for cells to<br />

begin to form the heart. But in mice, Wnt<br />

needs to be active – if you block it, the heart<br />

doesn’t form. It seemed like a contradiction –<br />

how could we reconcile these two stories?”


Within the first days of an embryo’s life, cells<br />

grow and separate into three distinct layers<br />

that then go on to differentiate into specific<br />

organs. The heart arises from the middle layer,<br />

called the mesoderm. Cells in various areas of<br />

this tissue receive different signals from their<br />

neighbors that tell them how to specialize.<br />

Some of them get instructions to become the<br />

heart. Those cells divide into two regions<br />

called the first heart field and second heart<br />

field, and now they are ready to receive their<br />

marching instructions.<br />

The next stages happen relatively quickly, and<br />

Walter sketches a map of the embryo on a<br />

piece of paper. Cells from the first heart field<br />

move into the chest area. Tagged with a fluorescent<br />

marker and seen from the front, they<br />

form a crescent, like a large horseshoe. Cells<br />

from the second heart field move into the<br />

area and line the inside of the shoe. When<br />

everything is in place, the crescent starts to<br />

fold inwards, as if the horseshoe has started<br />

to melt. When the sides come together they<br />

form a tube like the long, skinny balloon that<br />

clowns mold into animals at carnivals. What<br />

happens next resembles balloon art, too –<br />

cells in different parts grow at different rates,<br />

which makes the tube bulge. Since there isn’t<br />

much space to grow it begins to bend, as if<br />

the balloon has been threaded into a small<br />

Part one: The live of nomads<br />

38<br />

bottle and then inflated. These patterns of<br />

growth and bending create the chambers of<br />

the heart, the way the clown might tie his balloon<br />

into a bow.<br />

Wnt may be involved in migrations and the<br />

shaping, Walter says, but there were those<br />

conflicting results from earlier studies... And<br />

the picture is made more confusing by contributions<br />

from other signaling molecules such<br />

as the bone morphogenetic proteins, or Bmps.<br />

“As you might guess from their name, these<br />

molecules are mostly known for their role in<br />

the development of bone and cartilege,” he<br />

says. “But they also help build the front-to-


ack structure of the body of the early embryo,<br />

and at least in the chicken they are needed for<br />

the heart. So Alexandra Klaus, one of our PhD<br />

students, tried to make sense of all this.”<br />

She discovered that Bmp signals were active<br />

in both heart fields. The pathway from Wnt<br />

through β-catenin, on the other hand, was<br />

only switched on the second. It was interesting<br />

to find a difference between the two tissues;<br />

it suggested that Bmp might be necessary<br />

for the first heart field. Alexandra developed<br />

a knockout strain of mouse in which<br />

Bmp was switched off only in future heart<br />

cells. As the embryos formed she investigated<br />

the structure of their hearts. She discovered<br />

that first heart field never formed the crescent.<br />

But a few second heart field cells moved<br />

into the chest to form a thin rim.<br />

At what stage were things going wrong? “The<br />

mice made progenitor cells that were supposed<br />

to become the heart,” Walter says, “but<br />

they didn’t stay around to finish their development.<br />

We took this to mean that that the<br />

Bmp signal was crucial to finish their ‘programming’<br />

in the first heart field.”<br />

But the second heart field could still form, at<br />

least partially, without it. Since Wnt was<br />

active in this region, Alexandra developed a<br />

second mouse strain in which β-catenin was<br />

shut down in the heart fields. This had a quite<br />

different effect: cells were able to form the<br />

simple heart tube, but the heart didn’t form a<br />

loop. It was as if the clown had forgotten how<br />

to twist and shape his balloon.<br />

In mouse number three, Alexandra took the<br />

opposite approach: she raised the volume of<br />

the Wnt/β-catenin signal, turning it on in all<br />

the future heart cells – even in the first heart<br />

field where it normally isn’t active. This made<br />

things go wrong fast; the basic heart tube<br />

never formed at all. When Alexandra looked at<br />

the two heart fields she discovered that cells<br />

were straying off course and clustering in the<br />

wrong places.<br />

Cells not only have to specialize – they also<br />

have to get to the right places. Both processes<br />

are dependent on signals. All that was needed<br />

to block the development of the heart was to<br />

activate one of these signals in the wrong<br />

place, or to silence it where it needed to be<br />

heard. By watching what happened in these<br />

different conditions, Alexandra showed that<br />

Bmp was overseeing the behavior of future<br />

heart cells at an early stage, in the first field.<br />

Wnt and ββ-catenin were active in the second<br />

field, stepping up later to sculpt the tube and<br />

form heart structures.<br />

��<br />

The formation of the heart shows how in<br />

the embryo, cycles of growth and migra-<br />

39 Part one: The live of nomads<br />

tions are a normal state of affairs. Cells continually<br />

leave the structures where they are<br />

born and move off to build new organs. As<br />

they do so they specialize. Eventually most<br />

cells settle down, leading to a long period of<br />

adulthood for humans and most other animal<br />

species. An inappropriate signal can lead to<br />

disease by changing their developmental programs.<br />

If the cells get so confused that they<br />

leave their tissues and become nomads, the<br />

result may be metastasis.<br />

That often leads to disaster – but perhaps not<br />

always. The earliest animals didn’t have<br />

sophisticated organs or any genetic plan for<br />

building them. Far back in evolution, the intricate<br />

structures that serve our bodies must<br />

have started off as deviations from the blueprint,<br />

like tumors. If a signal were delayed, or<br />

accidentally switched on a second time, the<br />

shapes of tissues would be changed and new<br />

structures would arise.<br />

How did organs arise in the first place? If a<br />

change in a gene caused cells to migrate and<br />

build an odd new tissue, and if it was even<br />

slightly useful, it could be passed on and<br />

refined by natural selection. That wouldn’t<br />

have to happen very often to have a huge<br />

impact on evolution.<br />

It’s an odd way to think of a process that normally<br />

disrupts a life and may even lead to<br />

death. But the work of Walter and his colleagues<br />

support this view by showing that<br />

the same mechanisms needed for growth and<br />

migrations in embryos are also involved in<br />

cancer. Maybe none of the tumors that have<br />

appeared in humans – so far – have done anything<br />

useful for their victims. On the other<br />

hand, despite all the controls that make sure<br />

the body gets built right, tumors still arise.<br />

The system shouldn’t be so perfect that it<br />

eliminates all freedom. It may be evolution’s<br />

way of trying out new building projects from<br />

time to time, a way of saying that the story is<br />

not over yet, that change is still possible.


The archeology<br />

of blood<br />

Iwish there were a photograph of that moment in March, 1879, when<br />

two Germans stood on the plains of western coastal Turkey, staring<br />

up at a mound of dirt and rubble. One of them, the archeologist<br />

Heinrich Schliemann, was having trouble convincing the world that this<br />

unassuming hill held the remains of Troy. How could it contain the magnificent<br />

city immortalized by Homer? Most scholars at the time had<br />

become convinced that the ancient Greek poet had never existed and<br />

the war described in the Iliad was a fantasy. Most of Schliemann’s contemporaries<br />

regarded him as a rich, obsessed, and misguided amateur.<br />

If so, he was a lucky one. His archeological digs in Mycanae and Troy had<br />

turned up magnificent golden artifacts that confirmed, he claimed, the<br />

fabulous stories he had read as a child.<br />

The second man was a true amateur when it came to archeology, but<br />

was held in high regard by his peers in his chosen profession. Rudolf<br />

Virchow was recognized as one of Berlin’s best physicians and<br />

researchers. His 20-year-old theory that “every cell arises from a preexisting<br />

cell” and his explanation for embolisms (blood clots) had worked<br />

their way into the textbooks. Yet Virchow found other conflicts to keep<br />

him busy. After witnessing an outbreak of typhus in Silesia, he had<br />

Part one: The live of nomads<br />

40<br />

Achim Leutz


ecome convinced that poverty and unclean<br />

living conditions were major causes of disease.<br />

He pushed Berlin to build a modern<br />

sewage system, then became Germany’s<br />

most forceful advocate of social medicine and<br />

a founder of the German liberal party.<br />

Eventually this made him such a thorn in the<br />

side of the government that Chancellor Otto<br />

Bismarck once challenged him to a duel.<br />

(Virchow’s response was to laugh.)<br />

Virchow’s interest in archeology had prompted<br />

him to found the German Society for<br />

Anthropology, Ethnology and Prehistory. (One<br />

of its concerns was the study of the “Aryan<br />

41 Part one: The live of nomads<br />

race” – Virchow concluded that no such thing<br />

existed, and that the rising trend toward<br />

“Nordic mysticism” was nonsense). The position<br />

put him into contact with Schliemann,<br />

with whom he had exchanged letters for<br />

years. Troy was the first time he took part in<br />

one of the archeologist’s digs. Upon returning


Heinrich Schliemann<br />

Part one: The live of nomads<br />

42<br />

Sophia<br />

Schliemann<br />

wearing<br />

“Priam’s<br />

treasure”<br />

home he engaged in Herculean efforts to get<br />

the city of Berlin to buy a large part of<br />

Schliemann’s collection and put it on display.<br />

Some of the archeologist’s greatest finds<br />

remained in the city until 1945, when the<br />

Russian Army removed them from a bunker<br />

under the Berlin Zoo. Their whereabouts were<br />

unknown until 1993, when they turned up in<br />

the Pushkin Museum in Moscow.<br />

The importance of the encounter at Troy is<br />

symbolic. Mankind “discovered” history in the<br />

19th century – in both the natural and social<br />

sciences. James Hutton and Charles Lyell, two<br />

Scottish geologists, had shown that the Earth<br />

was thousands of times older than people<br />

had believed – perhaps millions. Such a long<br />

period, Charles Darwin and Alfred Russel<br />

Wallace proposed, would be enough for natural<br />

processes to produce life from inanimate<br />

substances and spin off all of the Earth’s life<br />

forms.<br />

While evolution created a history for the<br />

species, Virchow provided one for the individual.<br />

The idea that every cell arose from another<br />

cell dramatically changed the way scientists<br />

thought of embryos. Each structure in<br />

the body also had a history, one that could be<br />

traced back to a single fertilized egg. Each


stage of development was built on the last,<br />

the way that Troy was rebuilt at least nine<br />

times on the foundations of earlier settlements.<br />

While changing historians’ view of ancient<br />

Greece, Schliemann made another inadvertent<br />

contribution to the modern concept of<br />

history. He published an autobiography which<br />

is a confusing mixture of mix of facts, halftruths,<br />

and probably downright lies. When it<br />

fell into the hands of Sigmund Freud, it influenced<br />

the young doctor’s views of the development<br />

of an individual. A person’s psyche<br />

arises as a mixture of facts and myth, much as<br />

the true history of Troy is hidden under thousands<br />

of years of art and story-telling.<br />

Resolving the conflicts in a person’s life<br />

requires a sort of archeology.<br />

��<br />

Virchow’s “archeology of the embryo” has<br />

been extended by molecular biologists<br />

to the life histories of cells, and this is the central<br />

interest of Achim Leutz, researcher at the<br />

<strong>MDC</strong>. Achim has been studying how generic<br />

stem cells in the bone marrow specialize to<br />

become many types of blood cells.<br />

“A cell’s identity is determined by its pattern<br />

of active and inactive genes,” Achim says.<br />

“That’s a pattern spread across the whole<br />

genome and it may involve hundreds of molecules.”<br />

Activating a gene involves hooking it up to a<br />

very complex machine made of many molecules.<br />

This machine reads the base pairs on<br />

one strand of the double helix and transcribes<br />

the information into a similar molecule, called<br />

RNA. Often the process begins when a key<br />

protein, a transcription factor, binds to DNA or<br />

does something else to it – for example, it<br />

may remove another protein that blocks the<br />

way. The change provides a site for the transcription<br />

machinery to dock on and do its job.<br />

Achim’s lab and many others have intensively<br />

studied particular transcription factors<br />

because they are so powerful. By switching on<br />

Activating β-catenin at the wrong time in blood cell progenitors can cause them to develop in unusual<br />

and dangerous ways. This sample, taken from the bone marrow, reveals unusually high numbers of<br />

immature myeloid precursors.<br />

one or more important genes, these proteins<br />

can activate programs that cause cells to specialize,<br />

divide, or die. Removing a key factor<br />

from an animal’s genome may prevent particular<br />

types of cells, tissues, and even organs<br />

from developing. If a gene suffers mutations<br />

or become defective, the result is often disease<br />

and sometimes cancer or leukemia.<br />

“The body maintains some hematopoietic<br />

stem cells which are able to produce all types<br />

of blood,” Achim says. “They reproduce, at a<br />

slow rate. Usually when one of these cells<br />

divides it makes one copy that is an exact<br />

duplicate – it will become another stem cell –<br />

and another that will first divide many times<br />

but eventually specialize.”<br />

43 Part one: The live of nomads<br />

This second-generation cell is a precursor cell,<br />

equipped with a new genetic program (a pattern<br />

of active and quiet genes). It reproduces<br />

much faster, generating offspring that usually<br />

specialize quickly. The last generation is the<br />

largest, and it is made up of cells that are fully<br />

committed to becoming a particular type,<br />

such as red blood cells or the T and B cells that<br />

play an important role in the immune system.<br />

Once they reach this phase, they stop reproducing.<br />

When they wear out (red blood cells<br />

have a lifespan of about 120 days, for example),<br />

they are replaced by new precursors.<br />

Transcription factors play a role in moving the<br />

cells from one phase of their lives to the next.<br />

For example, the proteins SCL and Myb are<br />

needed at early phases. Later they are joined


Christiane Calliess<br />

by combinations of factors such as C/EBP proteins<br />

to make cells take the last step.<br />

“Mutations that affect these molecules usually<br />

have very bad effects,” Achim says. “If a transcription<br />

factor doesn’t work properly, it may<br />

stimulate the cell to keep reproducing longer<br />

than it should, instead of specializing. Or it<br />

may just block the process of differentiation,<br />

which leads to the same thing, because the<br />

cell can get stuck in a mode where it keeps<br />

dividing. This is how a lot of kinds of<br />

leukemias start.”<br />

So taking a developmental step requires the<br />

coordinated activity of a network of many<br />

genes. “We’ve learned a lot about how the cell<br />

controls the production of single molecules,”<br />

Achim says. “But we’re much further behind in<br />

seeing how it regulates complex networks.”<br />

A cell’s identity is determined by the pattern<br />

of genes it uses and those it doesn’t. While<br />

each cell in the body carries a complete copy<br />

of the genome, no cell uses it all to create proteins.<br />

Some proteins are made only in the<br />

brain; others are unique to blood cells.<br />

Patterns begin to form early in life. The first<br />

few cells in an embryo are nearly identical, but<br />

they quickly take on individual characteristics.<br />

Some are programmed to become skin, others<br />

parts of the nervous system, others internal<br />

organs. Usually once a cell starts down a particular<br />

road, there is no going back.<br />

“We understand these processes quite a bit<br />

differently than we did a decade ago,” Achim<br />

says. “The pattern that cells inherit is a whole<br />

network. It’s not a case-by-case situation, a lot<br />

of individual decisions about whether single<br />

genes should be on or off.”<br />

That network, he says, consists of combinations<br />

of molecules that work together to coordinate<br />

various processes, such as shutting<br />

down cell division on the one hand while<br />

pushing cells to develop and to mature. Not<br />

only do cells need to keep producing the key<br />

Part one: The live of nomads<br />

44<br />

transcription factors that control what’s<br />

going on – they also have to manage the signaling<br />

pathways that alert the factors to<br />

what they should be doing. Often those pathways<br />

do other things as well; understanding<br />

blood development has caused Achim to<br />

cross paths with several other groups, including<br />

Walter Birchmeier and Claus Scheidereit<br />

(“A hub in the city of the cell”). Walter’s<br />

favorite pathway begins with a signal from a<br />

molecule called Wnt, which is then passed to<br />

a molecule called β-catenin. This pathway<br />

plays a role in the development of limbs and<br />

other tissues – and blood. The meaning of its<br />

signals also depends on the context.<br />

So the cell’s guidance of development is not<br />

like people checking into a hotel one-by-one,<br />

picking up their keys, and going into their<br />

rooms. It’s more like a huge hotel where there<br />

are several big conferences going on at the<br />

same time. The people attending the same<br />

conference all have the same schedule, and so


they enter and leave their rooms at about the<br />

same time. They have a different program<br />

than people attending another conference,<br />

and a lot of coordination has to go on behind<br />

the scenes to ensure that there are enough<br />

rooms for the talks, the equipment gets to the<br />

right place, and the restaurants can handle<br />

the crowds.<br />

There’s also a parallel to archeology: other<br />

archeologists called Schliemann a treasure<br />

hunter, a digger of “rabbit holes.” During his<br />

first excavations he chose a promising spot<br />

and dug straight down, rather than carefully<br />

exposing and cataloguing the artifacts of<br />

each layer before moving to the next. In<br />

archeology, everything is context. For Achim,<br />

the same thing is true of blood.<br />

“Large developmental patterns are maintained<br />

as the cells divide – they aren’t reset;<br />

they don’t start from zero when a new cell is<br />

born. That means the cell carries a sort of history<br />

and memory through the process of dif-<br />

ferentiation, and our research is devoted to<br />

discovering where it lies and how it works.”<br />

��<br />

The DNA in a cell’s nucleus doesn’t look<br />

much like the neat double helix drawn<br />

by James Watson and Francis Crick in 1953. You<br />

can’t magnify the image enough to see how<br />

they intertwine. Even if you could, there are so<br />

many proteins attached to DNA that for a<br />

long time, most of the world’s best chemists<br />

believed that the genetic code was made of<br />

proteins.<br />

Even under the most powerful electron microscope,<br />

DNA appears as such a thin thread that<br />

you almost have to imagine it is there. During<br />

most of the cell cycle it is loose and spread<br />

through the nucleus like a ball of yarn the cat<br />

has been playing with. Just before cell division<br />

it undergoes a process called condensation,<br />

which packs it into huge, tight bundles – the<br />

chromosomes.<br />

45 Part one: The live of nomads<br />

Achim shows me an electron microscope<br />

image of the cell nucleus. In one the DNA<br />

thread is loose. It looks like a thin string on<br />

which a lot of beads have been strung. They<br />

act like tiny spools, Achim says, and they are<br />

called nucleosomes. Each one is made of eight<br />

proteins called histones. This structure helps in<br />

the process of packing when it comes time for<br />

the DNA to condense. It also plays a very important<br />

role in gene activation and silencing.<br />

On the average, one nucleosome is found at<br />

intervals of about every 200 base pairs along<br />

the DNA strand. It wraps 1.65 times around<br />

each of these beads – that takes 146 base<br />

pairs. The remaining 50 or so base pairs form<br />

a linker region that separates one nucleosome<br />

from the next.<br />

Histones, the building blocks of nucleosomes,<br />

have long tails that stick out of the nucleosome.<br />

Other proteins can attach themselves<br />

here, or stick groups of atoms onto parts of<br />

the tail so that it behaves differently. One


effect is to slide the nucleosome to a new<br />

position. That may be important in activating<br />

a gene. The molecular machines that read<br />

information in genes to transcribe it into RNA<br />

often have a hard time getting to DNA that is<br />

wrapped around a nucleosome. By sliding the<br />

spool to a new position, proteins can clear the<br />

way for the transcription machinery.<br />

Two chemical tags that are placed on the tails<br />

of histone proteins often have strong effects<br />

on gene activity. Often if a tail is tagged with<br />

methyl groups, a nearby gene is silenced.<br />

Another type of tag, acetyl, often activates the<br />

gene. Sometimes tags are applied in a “blanket”<br />

way to large stretches of DNA. That<br />

would be one way of coordinating a lot of<br />

nearby genes to behave the same way. It<br />

might also work as a memory device.<br />

Activating or silencing blocks of genes is one<br />

way to control several things at the time,<br />

Achim says, but the story is more complex.<br />

The codes and patterns that tell cells what to<br />

become are subtle. A methyl tag doesn’t<br />

always turn a gene off; sometimes it has the<br />

opposite effect, and the same histone tail may<br />

even contain both types of tags. This makes it<br />

very hard, Achim says, to decipher the code<br />

and the interplay between the many factors<br />

that control each gene.<br />

By whatever means genetic programs are<br />

coordinated, transcription factors must play<br />

an essential role. So Achim’s group has been<br />

looking into the links between these proteins<br />

and histones. Recently they have received help<br />

from an unlikely source: a virus.<br />

��<br />

Part one: The live of nomads<br />

46<br />

IIn the early 20th century Peyton Rous, a scientist<br />

at the Rockefeller Institute for<br />

Medical Research in New York, made a startling<br />

discovery: he proved that a type of cancer<br />

could be transmitted from one animal to<br />

another by a virus.<br />

From the reaction of his colleagues, you would<br />

think he had claimed to discover Troy. No one<br />

had trouble with the idea that a virus might<br />

cause a disease – the microbe theory of Louis<br />

Pasteur and the Berlin researcher Robert Koch<br />

were known to every schoolchild. But cancer<br />

was different. Rudolf Virchow had said that<br />

cancer arose spontaneously from defects in<br />

cells, and that was the accepted dogma.<br />

Rous was working with chickens, which commonly<br />

suffered from a type of cancer called a<br />

sarcoma. These tumors grow in tissues such


as bone and muscle. After watching sarcomas<br />

spread like infections through rural farmyards,<br />

he began searching for the agent that was<br />

transmitting the disease. He removed a sarcoma<br />

from the leg of a chicken, prepared and filtered<br />

it until there were no more cells, and then<br />

injected it into a healthy bird. Transplantees<br />

developed the same types of tumors, and by<br />

repeating the process, Rous could pass them on<br />

to new chickens in the same way.<br />

Because he had carefully filtered the extracts<br />

before injecting them into a new animal, Rous<br />

hypothesized that the agent had to be something<br />

smaller than a cell – a virus. Even after<br />

Rous tried everything he could think of to disprove<br />

his own hypothesis and failed, his colleagues<br />

remained skeptical. He tried to find<br />

cancer-causing viruses in mice and failed,<br />

eventually becoming frustrated and abandoning<br />

the line of work.<br />

Two decades later, however, a colleague asked<br />

for his help in the study of warts that arose in<br />

mice because of a virus; the warts frequently<br />

developed into cancer. The idea of cancercausing<br />

viruses gradually became accepted<br />

and finally in 1966, a few years before his<br />

death, Rous was presented with the Nobel<br />

Prize in Physiology or Medicine.<br />

“It took molecular biology to build the bridge<br />

between viruses and cancer,” Achim says.<br />

“Viruses sometimes capture genes from cells.<br />

Eventually those molecules mutate. When<br />

they come back to infect a new person or animal,<br />

they act differently. But the cell may not<br />

be able to tell the difference; it may interpret<br />

the molecule as one of its own.”<br />

47 Part one: The live of nomads<br />

Many viruses that cause cancer do so by interfering<br />

with transcription factors or molecules<br />

that interact with them. That was the case<br />

with a bird virus called AMV. It once kidnaped<br />

the gene for the transcription factor Myb<br />

from an ancient host, added the molecule to<br />

its own viral genome, and carried it along in a<br />

mutated form as it invaded new hosts. When<br />

the infected cells start making proteins from<br />

this viral form of the gene (known as v-Myb),<br />

it interferes with the cell’s own version<br />

(c-Myb).<br />

“Remember that Myb is one of the important<br />

transcription factors that step in early in<br />

blood development,” Achim says. “Later it<br />

combines with other proteins to help them<br />

differentiate. It has to work perfectly to do<br />

both those jobs. v-Myb does fine at telling


A chart of blood cell development.<br />

The progenitor of all types of blood cells, hematopoeitic stem cells, divide to create copies of themselves<br />

so that the body will always have blood stem cells on hand. Some of the new cells begin to specialize<br />

into ever-more specific types, passing through intermediate stages of development along the way. The<br />

fully differentiated cells at the bottom of the chart no longer divide.<br />

cells to reproduce, but it fails at the differentiation<br />

step. This creates partially differentiated<br />

cells that keep dividing, and that means<br />

leukemia.”<br />

What happens in such cells, he says, is a failure<br />

to coordinate the programs that guide<br />

division and differentiation. Maybe comparing<br />

the two types of Myb would show what<br />

was going wrong, and that might reveal<br />

something about the coordination machinery.<br />

There were more reasons to think that Myb<br />

was crucial to decisions about the fates of<br />

cells. “The viral form of the molecule isn’t the<br />

only one that is linked to cancer,” Achim says.<br />

“In many kinds of human leukemia, cells make<br />

too much c-Myb. It makes you think that the<br />

problem might be overactive Myb, maybe it<br />

sends signals that tell the cell to divide too<br />

often. On the other hand, high levels of Myb<br />

alone aren’t enough to cause disease. A group<br />

in Japan developed a strain of mouse that produces<br />

too much of the molecule, and they<br />

don’t develop tumors.”<br />

But if you think of Myb as a coordinator of cell<br />

division and specialization, he says, some of<br />

these things make sense. He put his group to<br />

work comparing the effects of v-Myb and c-<br />

Myb on cells.<br />

The basic differences between the two molecules<br />

were well known, Achim says. “Myb is a<br />

complex protein with many different modules.<br />

The viral form of the protein is missing<br />

segments at both ends, the head and tail.<br />

Besides that, there are 11 spelling errors that<br />

change the chemistry of the protein.” Nearly<br />

20 years ago, when Achim had worked as a<br />

staff scientist at the European Molecular<br />

Biology Laboratory in Heidelberg, his colleagues<br />

in Thomas Graf’s group showed that<br />

three of these errors were the most important<br />

Part one: The live of nomads<br />

48<br />

in causing the problems related to cell division<br />

and differentiation. If they fixed the<br />

spelling errors in these three places – by<br />

changing the code of v-Myb – its ability to<br />

cause leukemia dropped to almost nothing.<br />

“With a transcription factor, your immediate<br />

thought is that mutations might change the<br />

way the molecule binds to DNA to activate a<br />

gene,” Achim says. “But here that wasn’t the<br />

case. Just a few years ago a Japanese laboratory<br />

showed that when v-Myb docks onto DNA,<br />

the part of the protein affected by the mutations<br />

faces away from the DNA. So something<br />

else is going on – we thought the change<br />

might stop it from interacting with other proteins.<br />

It was another hint that some sort of<br />

coordinating activity was being disturbed. The<br />

mutations might change how healthy Myb<br />

docks onto other proteins, which stop it from<br />

launching a cancer.”<br />

Experiments by Achim and his colleagues previously<br />

showed how some of the changes<br />

affected a module of Myb that normally binds<br />

to the transcription factor C/EBP. This interaction<br />

is needed to stimulate the differentiation<br />

of some types of blood cells.<br />

But the spelling errors in v-Myb seem to have<br />

another effect, Achim says. “One of Myb’s<br />

modules is built like a machine we have seen<br />

in many other proteins. In those molecules,<br />

the module helps change the nucleosomes –<br />

the spools that DNA is wrapped around. That<br />

changes the accessibility of certain genes and<br />

affects whether they can be activated.”<br />

Xianming Mo, a postdoc in the lab, decided to<br />

find out. He showed that the healthy form of<br />

Myb can bind to the tail of one of the eight<br />

histones in the spool, called H3. v-Myb can’t<br />

duplicate that feat.<br />

“Very interesting but not conclusive,” Achim<br />

says. “Histones play a vital role in activating<br />

genes. But finding out that things can bind<br />

doesn’t mean that they do bind. Xianming<br />

and Elisabeth Kowenz-Leutz had to prove that<br />

they did. They started correcting the spelling


of v-Myb and discovered that changing any of<br />

the three critical letters allowed it to bind to<br />

the histone tail. And at the same time its cancer-causing<br />

ability to block the activation of<br />

the differentiation genes dropped off.”<br />

On the other hand, he says, making v-Myb<br />

more like c-Myb doesn’t seem to affect how it<br />

binds to DNA. So the mutations don’t mean<br />

that the two molecules target different genes<br />

– it’s what happens afterwards that counts.<br />

And what was that? “We knew that Myb could<br />

bind to a protein called p300, which is an<br />

‘acetyl-loader,’” Achim says. “Proteins like<br />

p300 can tag histone tails with acetyl. That<br />

chemical change often causes the activation<br />

of a gene.”<br />

The scientists combined different forms of<br />

Myb and p300 in cells and checked whether<br />

acetyls were loaded onto the tail. That didn’t<br />

happen with the viral form of Myb. But they<br />

could start the tagging by correcting any one<br />

of the three crucial spelling errors.<br />

So its presence on a gene allows healthy Myb<br />

to coordinate a second process; that ability is<br />

lost when the molecule is replaced by the viral<br />

form of the molecule. The take-home message<br />

from the study, Achim says, is that the<br />

chemistry of the histone is a key player in<br />

coordinating Myb’s effects on the division and<br />

differentiation of blood cells.<br />

And it’s the place to look if you want to know<br />

how cancer arises.<br />

��<br />

The finding is theoretically interesting,<br />

Achim says. It’s good archeology – the<br />

difference between finding a bit of gold at<br />

Troy and offering it for sale on E-bay, versus<br />

learning where the metal in it was mined, and<br />

what it says about the trade and culture of<br />

the people who made it. Events shouldn’t be<br />

seen through rabbit holes; good interpretations<br />

require a context. Transcription factors<br />

don’t act alone. What they mean depends on<br />

the global condition of the cell’s DNA.<br />

49


50<br />

“And let’s not stop there,” Achim says. “It suggests<br />

a new approach to cancers that involve<br />

c-Myb. The problems might arise because H3<br />

isn’t getting its dose of acetyl. That blocks the<br />

activation of genes that tell the cells to stop<br />

dividing and start differentiating. Those genes<br />

aren’t active enough in leukemia cells. Well,<br />

we could test that. We took leukemia cells in<br />

laboratory cultures and fed them with a substance<br />

that promotes more tagging of H3<br />

tails with acetyl. Suddenly the key specialization<br />

genes were switched on again and the<br />

leukemia cells differentiated.”<br />

So there may be medical applications. Both v-<br />

Myb and c-Myb may behave the same way in<br />

terms of binding to genes. The changes in<br />

Myb that cause leukemia – even if there is no<br />

virus involved – may be doing the same thing.<br />

If that’s the case, Achim says, we need to<br />

change the focus of therapy. Aberrant forms<br />

of Myb may be sitting on the right genes,<br />

ready to do their jobs. But if they can’t communicate<br />

with the tail of histone 3, they can’t<br />

help it get loaded with acetyl.<br />

“Find another way to deliver that tag,” Achim<br />

says, “and you may patch the problem with<br />

Myb.”<br />

��<br />

One of the half-truths in Heinrich<br />

Schliemann’s autobiography may be his<br />

explanation for how he became interested in<br />

the myth – and history – of Troy. Like so many<br />

scientists, he claims to have been inspired by a<br />

book. In his case, it was a world history for<br />

children, illustrated with an engraving of Troy<br />

and its massive stone walls.<br />

Achim has one of those stories about a book.<br />

“My father was a locksmith and my parents<br />

didn’t much understand my interest in science,”<br />

Achim says. “But I had an uncle who<br />

was always giving me things. An old microscope,<br />

a telescope. And one day he gave me an<br />

old children’s book from 1938. Chemistry<br />

experiments that work, by Hermann Römpp.<br />

You could get everything you needed to do


the experiments from a normal pharmacy.<br />

That began my experimental career, which<br />

was mostly focused on producing bad smells<br />

and blowing things up.”<br />

From what I can tell after a little research, just<br />

about every German scientist born after 1938<br />

has at one time or another had his hands on<br />

Chemistry experiments that work and cites it<br />

as an inspiration to become a scientist. The<br />

list includes Gerhard Ertl, winner of the 2007<br />

Nobel Prize in Chemistry, and former director<br />

of the Fritz Haber Institute of the Max Planck<br />

Society here in Berlin.<br />

So Achim knows the importance of a good<br />

book and an education in science; maybe<br />

that’s one reason he has spent so much effort<br />

helping to build the <strong>MDC</strong>’s PhD program. As<br />

Dean of the graduate school he coordinated<br />

the program from its conception to the present<br />

state: a student body of about 170,<br />

increasingly oriented toward international<br />

students. The <strong>MDC</strong> receives far more applications<br />

from students than it can accept. That’s<br />

unfortunate for the ones we have to turn<br />

down, he says, but the end result it increases<br />

the quality of the students and the program.<br />

You can judge that for yourself by considering<br />

the fact that the main investigators on many<br />

of the stories in this book are PhD students.<br />

Maybe we could collaborate, I say. “How about<br />

a popular book: Experiments with histones<br />

that work? Where you can get everything you<br />

need from your back yard, and some tissue<br />

samples taken from your brothers and sisters?”<br />

He laughs.<br />

“You mean to excite the next generation of<br />

scientists?” he says. “You have to have the<br />

explosions and the smells. If you can’t blow<br />

anything up, and your mom doesn’t get mad<br />

at you for smelling up the house, it won’t be<br />

as inspiring to kids.”<br />

51


Breakouts in the brain<br />

If at any time during the first half of a Western film the bandit is<br />

caught and locked up in a jail cell, a breakout is sure to follow. There<br />

are a dozen variations on the theme. A wall may be blown up with dynamite.<br />

A guard may be lured into the cell, or at least close enough for the<br />

prisoner to snatch his keys. Or the sheriff might be paid off, becoming a<br />

co-conspirator whose only job is to look away during the escape.<br />

Researchers who study the brain are familiar with some of these scenarios<br />

– a type of brain cancer called a glioma is a master breakout artist.<br />

Part one: The live of nomads<br />

52


Surgery almost never succeeds in completely removing these tumors<br />

because cells escape all the time, and each one that leaves is like a<br />

seed, able to give rise to a new tumor.<br />

Many of the breakouts are witnessed by healthy brain cells that rush<br />

to the site of a tumor, including some that you would think would act<br />

as guards. Helmut Kettenmann and his laboratory at the <strong>MDC</strong> are trying<br />

to discover their roles in containing tumors or helping them break<br />

out. This work is part of a larger plan to uncover the functions of a<br />

huge group of cells in the brain that have been underappreciated for<br />

far too long. Most brain research has been focused on neurons. Their<br />

smaller and much more numerous companions, a wide range of glial<br />

cells, have been waiting patiently on the back burner.<br />

Scientists got their first real look into the structure and functions of<br />

brain cells in the late 19th century. The Italian researcher Camille Golgi<br />

had developed a new method of staining tissues which, when applied<br />

to the brain, revealed a intricate web of majestic, branching cells. Golgi<br />

believed that they formed a continuous, fused network, like blood vessels.<br />

One of his contemporaries, a Spaniard named Santiago Ramón y<br />

Cajal, promoted a contrary view – neurons were single cells, separated<br />

from their neighbors by tiny gaps called synapses. He made fantastically<br />

detailed drawings of what he saw under the microscope and concluded<br />

that neurons communicated by receiving impulses at their<br />

bush-like dendrites, then passing them along their axons to other<br />

cells.<br />

Helmut knows quite a bit about Cajal – he has visited the scientist’s<br />

historic lab in Zaragosa and followed some of his tracks through<br />

Berlin. Cajal was a medical doctor with an indomitable passion for cell<br />

biology, a field virtually unknown in Spain. He was a gymnast and an<br />

artist – which helped when it came time to draw what he was seeing<br />

under the microscope; his drawings are still used in textbooks to illustrate<br />

the types and structures of nerve cells. Under the pseudonym<br />

53 Part one: The live of nomads<br />

Helmut Kettenman


Santiago Ramón y Cajal<br />

and his drawings<br />

Part one: The live of nomads<br />

“Dr. Bacteria” he wrote science fiction stories but didn’t publish them<br />

until 1905, fearing that his scientific work wouldn’t be taken seriously.<br />

He also wrote novels, but later said they had been lost in Cuba, where<br />

he spent two years as an army doctor.<br />

A copy of the book of stories sits on the shelf behind Helmut’s head.<br />

One describes an astronaut on Jupiter who encounters giant creatures,<br />

enters their bloodstream, and watches battles between white<br />

blood cells and parasites. Scientists in the stories are not only heroes.<br />

“One concerns an old bacteriologist who suspects that his wife is having<br />

an affair with one of his lab assistants,” Helmut says. “To prove that<br />

they are meeting secretly in the lab, he rigs the couch with a seismograph.<br />

When the results confirm his fears, he infects the assistant with<br />

tuberculosis, who passes it along to his wife. The assistant dies, but at<br />

the last minute the scientists administers an antidote to his wife.”<br />

After nearly dying of malaria in Cuba, Cajal returned home and was<br />

given a full professorship at the University of València in 1881. Later he<br />

held professorships in Barcelona and Madrid. His vocation was also his<br />

hobby; any extra money he made went toward the purchase of microscopes.<br />

He learned of Golgi’s staining methods from a colleague in<br />

Madrid and rushed home to try them out. He was a precise and meticulous<br />

worker and his own improvements to the staining process<br />

allowed him to see details of nerve structure that had been invisible to<br />

Golgi. But he had few peers to appreciate his work in Spain, and trying<br />

to get his research published was a frustrating experience. Finally he<br />

decided to publish his own journal, the Trimonthly Review of Normal<br />

and Pathological Histology, but he could only afford the printing and<br />

postage costs of 60 copies, which he sent off to scientists around the<br />

world. It was like mailing them into the void.<br />

Cajal decided that the only solution was to head for one of the world<br />

capitals of science – Berlin – and establish direct contact with other<br />

scientists. He joined the German Anatomical Society and in 1889 managed<br />

to find enough money to make a pilgrimage to their annual<br />

meeting, held at the University of Berlin. At first his presence must<br />

have seemed simply odd. Spain wasn’t known for its scientists, and he<br />

didn’t speak a single word of German – he had to communicate in broken<br />

French. And he had brought along his own microscope, a good<br />

Zeiss model he was familiar with, not trusting the instruments that<br />

might be available the conference. He set up his stand among all the<br />

others, and then went through what every PhD student and postdoc<br />

experiences at their first big conference: few people came by for a look.<br />

Everyone seemed much more preoccupied with his own work.<br />

But then some people did stop by to peer through his Zeiss, including<br />

Albrecht von Kölliker, the “patriarch of German histology.” What<br />

Kölliker saw made him stay, and that drew a larger crowd. For Caval it<br />

was a life-changing moment. He wrote,<br />

54


As to be expected, these savants, then world celebrities,<br />

began their examination with more skepticism than curiosity.<br />

Undoubtedly they expected a fiasco. However, when there had been<br />

paraded before their eyes...a procession of irreproachable images of<br />

the utmost clearness, the supercilious frowns disappeared. Finally,<br />

the prejudice against the humble Spanish anatomist vanished and<br />

warm and sincere congratulations burst forth.<br />

Kölliker immediately took Cajal under his wing and whisked him away. “He took<br />

me in a splendid carriage to the luxurious hotel where he was staying; entertained<br />

me at dinner; presented me afterwards to the most important histologists<br />

and embryologists of Germany, and, finally, made every effort to render<br />

my sojourn in the Prussian capital agreeable.” Kölliker returned to his own lab<br />

where he reproduced Cajal’s results, then began promoting the Spaniard’s reputation<br />

in Europe. Almost overnight Cajal became a celebrity, receiving invitations<br />

to conferences in London and even the United States – despite the fact<br />

that the two countries had just been at war.<br />

The biggest recognition, however, came in the form of the 1906 Nobel Prize in<br />

Physiology or Medicine. Ironically, the prize was jointly awarded to Camillo<br />

Golgi. Although by 1906 most of the scientific community had become convinced<br />

that Cajal was right – neurons are separated from their neighbors by<br />

synapses – Golgi remained stubborn and bitter. Instead of devoting his acceptance<br />

speech to his own prize-winning research, he attempted to raise his old<br />

theory from the dead.


��<br />

For brain science, discovering neurons and their functions was like<br />

emerging from the dark ages. Ever since, neurons have held the<br />

spotlight of most research; they are given most credit as the architects<br />

of thought and the storehouses of memory. But as important as they<br />

are, they make up only ten percent of the cells in the brain. This figure<br />

led to the urban legend that people only use ten percent of their<br />

brains.<br />

Helmut and his lab have devoted themselves to the other 90 percent.<br />

Formerly regarded simply as assistants, or nuts and bolts that keep<br />

things in place, glial cells are now being recognized for playing active<br />

roles in nearly everything that happens in the brain. Including the<br />

growth and spread of brain tumors.<br />

Glial cells come in several types. There are shaggy oligodendrocytes,<br />

which help carry out electrical maintenance. The long, cable-like axons<br />

of neurons have to carry electrical charges along great distances, and<br />

without some sort of insulation, the charge would quickly leak away.<br />

Oligodendrocytes coat the axon in a sheath of fats and proteins<br />

that act as insulators. Several diseases, including multiple<br />

sclerosis, are caused by a loss of the sheath.<br />

The functions of another type of glial cell, astrocytes, are<br />

less clear. “We know they provide guidance during the<br />

development of the brain,” Helmut says. “They have extensive<br />

contacts with synapses and they respond to neuronal<br />

activity; they can be stimulated by neurotransmitters. One<br />

of our interests has been to understand their role in communication<br />

between brain cells.”<br />

Another type of cell, microglia, was described and<br />

named by one of Santiago Cajal’s students, Pio del Río-<br />

Hortegas, in 1919. A few years later he noticed that the<br />

cells are drawn to sites of brain injury. This suggested<br />

that they might have functions in the immune system.<br />

The idea couldn’t be proven, however, until the late<br />

1980s. Working at the University of Pennsylvania, William<br />

Hickey and Hiromitsu Kumura found immune system molecules<br />

called MHC proteins on the surfaces of microglia.<br />

Microglia respond to tumors as well as injuries.“They<br />

are attracted by small molecules released by the<br />

tumors,” Helmut says. “Their response is so strong that<br />

a typical glioma may consist of up to 30 percent<br />

microglia.”<br />

What hasn’t been clear is why they come –<br />

whether they are fighting the tumor or somehow<br />

helping it. Most of what is known about the cells<br />

Part one: The live of nomads 56<br />

comes from watching how microglia respond to disease. Their roles in<br />

the healthy brain are less clear. Helmut says there is increasing evidence<br />

that they constantly survey tissue and step in to repair minor<br />

damage to blood vessels and to remove diseased neurons. In many situations,<br />

such as strokes, microglia congregate at the site of damaged<br />

tissue. A 2007 study carried out by Mélanie Lalancette-Hébert and<br />

Jasna Kriz of the Université Laval in Quebec showed that strokes led to<br />

significantly more damage in mice which lacked microglia than in<br />

their normal counterparts. And when gerbils suffering from a strokelike<br />

condition were given extra microglia, the cells homed in on the<br />

damage site. These animals lost fewer brain cells than gerbils that had<br />

not been treated. In these cases the role of microglia is clearly protective.<br />

Their normal duties ensure health, but when faced with tumors or<br />

another serious problem, microglia can aggrevate the problem.<br />

Microglia help remove α-β, a protein fragment that accumulates<br />

between cells and leads to the death of neurons in Alzheimer’s disease.<br />

So their activation and protein-clearing activity has healthy<br />

effects. But if there is an inflammation, damaged tissue releases molecules<br />

that slow them down. So in Alzheimer’s disease<br />

and other situations, such as gliomas, it might be best to<br />

keep microglia away – or at least to control how they<br />

respond to problems.<br />

In 2005 Darko Markovic, a postdoc in Helmut’s group,<br />

headed a study which showed that the number of<br />

microglia in a glial tumor in mice corresponds to the<br />

severity of the disease. More cells are a sign of a larger<br />

tumor that is more likely to develop into metastases. This<br />

made it very unlikely that the microglia were there to<br />

fight the tumor. If so, they were doing a poor job. In fact,<br />

the opposite might be the case. They might be helping<br />

it along.<br />

Helmut’s lab had developed a method of using thin<br />

slices of brain tissue to watch the behavior of brain cells<br />

and the growth of tumors under the microscope. One<br />

discovery was that microglia appeared to be helping<br />

tumors to grow and escape.<br />

As this book was being written, Darko and his colleagues<br />

had just found an important mechanism by which this<br />

happened. The article they wrote is now in the mail to a<br />

scientific journal, which means that everyone is hanging<br />

on waiting to see if it will be accepted and published.<br />

Journals have a strict rule that information in an article<br />

can’t be made public before it appears in print, so<br />

Helmut won’t tell me the name of the molecules<br />

involved.


“Let’s call them X and Y,” he says. “What we discovered seemed to<br />

revolve around the behavior of protein X. It’s an enzyme involved in the<br />

escape of glioma cells, helping break down the ties between a cell and<br />

its neighbors. That’s an essential step in releasing the cell from the<br />

tumor so that it can migrate. Somehow this molecule was being activated<br />

in tumor cells by the arrival of the microglia.”<br />

Researchers at the University of Pittsburgh Cancer Institute in the US<br />

had detected very high amounts of X and also Y, a molecule that it<br />

interacts with, in tumors from glioma patients – levels far higher than<br />

those found in healthy brain cells. X plays a crucial role in the activity<br />

of its partner. Normally Y is attached to the membranes of cells in an<br />

inactive form. X cuts off a part of the molecule and releases it. This<br />

activates Y, which begins chewing at the protein fibers that link cells.<br />

That was interesting, but in itself it didn’t say anything about why the<br />

arrival of microglia stimulated tumors to grow or cells to escape.<br />

Darko and his colleagues began trying to overhear the molecular conversations<br />

that were taking place between the glioma and microglia.<br />

First the scientists marked the cells so that they could tell the difference<br />

between the two types. Their images suggested that most –<br />

maybe all – of the X was being produced by microglia. Those that come<br />

into contact with the tumor were producing much more of the protein<br />

than microglia elsewhere in the brain. To be sure that this was the<br />

case, the scientists returned to their brain slices, this time using tissue<br />

whose microglia had been removed. They added tumor cells, which<br />

grew and invaded the tissue. But this time they saw no rise in levels of X.<br />

“All of this suggested that tumors might be causing microglia to produce<br />

more of the molecule,” Helmut says, “so Darko did a number of<br />

experiments to find out. He extracted the medium that had been used<br />

to grow cancer cells and put it into cultures with microglia. This medium<br />

contains molecules that have been secreted by the tumor cells.<br />

After three to six hours, the microglia began producing much higher<br />

amounts of X.”<br />

Normally cells are told to start making X by a pathway of signaling<br />

molecules – like the telephone lists described in the previous story. The<br />

signals belong to a well-studied circuit called the MAPK pathway. Were<br />

these also the molecules used by gliomas to talk to microglia? Darko<br />

and his colleagues blocked the pathway with a drug called SB201290<br />

and found that cells didn’t produce nearly as much X. An antibiotic<br />

called minocycline, which interferes with MAPK signals, had even<br />

stronger effects.<br />

Microglia might be assisting in the breakout of tumor cells by bringing<br />

along the tool that cut free Y on the tumor. Gliomas didn’t seem to<br />

make this tool themselves; they didn’t produce much – if any – X. The<br />

scientists wondered why not, so they transplanted the molecule into<br />

tumor cells. Glioma cells which produced both molecules died. So to<br />

activate Y and escape, they had to obtain the keys from microglia.<br />

If escape was the topic of the dialogue between microglia and<br />

gliomas, then X might make a good place to jam the conversation.<br />

Once again using the brain slice model, Darko allowed cancer cells to<br />

invade tissue whose cells couldn’t produce the protein. Without X,


tumors grew to only half the size. If the tissues also contained no<br />

microglia, tumors only reached 16 percent of their normal size. “This<br />

shows,” Helmut says, “that X is an important tumor-promoting factor<br />

in the cells. But it isn’t the only one.”<br />

Tumors could be slowed down by removing the protein entirely –<br />

would blocking its activity do equally well? The scientists tried different<br />

scenarios using the SB201290 drug and the antibiotic minocycline.<br />

The antibiotic cut the growth of tumors by nearly half, specifically<br />

because it interrupted the production of X in microglia. By itself<br />

SB201290 had a smaller impact on tumors, but when used in tissues<br />

without microglia, it cut their size by two-thirds.<br />

“The results showed that minocycline might have a therapeutic<br />

impact, so we administered it to mice with brain tumors,” Helmut says.<br />

“Here we noticed a very significant impact. Over two weeks, tumors in<br />

the treated mice grew to a size of about one cubic millimeter. In control<br />

mice that had not received the antibiotic, they were more than<br />

four times as large. Even when we started the treatment later, after the<br />

tumor had already started to grow, we saw a significant reduction in<br />

tumor size.”<br />

Part one: The live of nomads<br />

Minocycline may have additional tumor-fighting properties. A 2005<br />

study by researchers at the University of British Columbia in Canada<br />

showed that the antibiotic interferes with the activity of microglia in<br />

animal models of Alzheimer’s disease and brain injury. “We don’t yet<br />

completely understand how the substance works,” Helmut says.<br />

“Besides blocking the MAPK signaling pathway, and stopping X, it has<br />

other effects. It seems to interfere with some of the molecules secreted<br />

by the tumor, which likely block signals that the microglia use to find it.”<br />

��<br />

Microglia are not the only cells that take notice of brain tumors<br />

and respond. They are joined by neural precursor cells – a population<br />

of cells that have not yet completely differentiated. For a long<br />

time it was not known that adults had such cells. Most brain-building<br />

activity occurs in the embryo and young animals and is lost in adults.<br />

But a few regions of the brain protect and keep precursor cells through<br />

adulthood. In some situations they are pulled from storage and triggered<br />

to develop to take the place of injured neurons.<br />

One of the difficulties in studying the development and functions of<br />

these cells has been that they could not be kept alive and cultured in<br />

the laboratory. Recently members of Gerd Kempermann’s lab and<br />

Helmut’s group developed a new method for doing so. It’s a crucial step<br />

in learning about the chemistry of the cells and their funcions.<br />

“The fact that neuronal precursors target gliomas has made the<br />

cells interesting from a therapeutic point of view,” Helmut says. “If<br />

58<br />

Neural precursor cells (green) migrate to the site of a tumor (dark<br />

area in the center). The red dye helped Helmut and his colleagues<br />

establish where in the brain they migrated from.


The Leiner room in the City of Konstanz is a "museum within a museum." It was established in 1870 and today has the status of a protected historical monument. The<br />

closely arranged displays and hand-written descriptions reflect the museum style of the time.<br />

their activity and development could be controlled, it might be<br />

possible to use them to repair other types of damage. And they have a<br />

second quality which has attracted interest – they are drawn to<br />

gliomas.”<br />

In 2000 this led Karen Aboody in Evan Snyder’s lab at the Harvard<br />

Medical School to suggest that precursor cells might make good vehicles<br />

to deliver gene therapies to tumors. The same year scientists at<br />

the Istituto Nazionale Neurologico Besta in Milan modified the cells to<br />

produce extra interleukin-4, a molecule that powerfully triggers<br />

immune reactions. When they injected the cells into mice with<br />

tumors, the animals survived longer. Interestingly, when unmodified<br />

precursor cells were injected, the mice survived longer, too. So far the<br />

studies had shown that cells injected into animals could move to the<br />

tumor, but what about those produced by the animal’s own body? The<br />

cells might have beneficial effects on their own. Helmut had several<br />

members of his lab – Rainer Glass, Michael Synowitz, and Golo<br />

Kronenberg – look more deeply into the relationship between gliomas<br />

and precursor cells.<br />

It was important to be able to tell the difference between cells attracted<br />

to the tumor and any new ones that might arise from the tumor<br />

itself – gliomas often arise as mutations in precursor cells. The lab of<br />

Gerd Kempermann, a colleague in the <strong>MDC</strong>’s program on Functions<br />

and Dysfunctions of the Nervous System, had developed a line of mice<br />

whose precursor cells that could be identified because they produced<br />

a protein with a green fluorescent tag. The tumor cells had been engineered<br />

to emit a red fluorescent signal.<br />

After two weeks, Rainer and his colleagues discovered that a large number<br />

of green precursor cells had been drawn to the tumor. Not only had<br />

they penetrated it – they had it surrounded, coating it in several layers<br />

of cells. “In the border zone they were intermingled with cells from the<br />

glioma, a few of which had escaped,” Helmut says. “Interestingly, most<br />

of them were also accompanied by green precursor cells. So the precursors<br />

are also strongly attracted to cancer cells outside the main mass of<br />

the tumor.” This attraction was much stronger than to other types of<br />

injuries, where migrations of precursor cells were also found, but at<br />

lower levels and in a much looser structure.<br />

59 Part one: The live of nomads


Steffi Siefert<br />

If precursor cells somehow reduced the effects of tumors, it might<br />

help explain why children rarely develop gliomas; most cases appear<br />

when people reach their mid-fifties. While the brain retains precursor<br />

cells its whole life, there aren’t as many with advancing age, and those<br />

that are still around may not function as well. The study by Rainer and<br />

his colleagues confirmed that the number of precursors attracted to<br />

tumors in mice dropped sharply with age<br />

In young mice the cells envelop the tumor, and their appearance<br />

restricts its growth. In older mice, fewer cells are attracted, and the<br />

gliomas grow more vigorously. To show that this was a direct effect of<br />

the precursor cells, Rainer injected adult mice with a mixture of precursor<br />

and tumor cells. The mice survived much longer than animals<br />

that had not received the precursors. In fact, their prospects for survival<br />

were just as good as those of much younger mice.<br />

Another round of experiments helped to show why. Tumor cells reproduce<br />

rapidly in cell culture, and few of the cells die. When grown alongside<br />

precursor cells, four times as many suffered that fate.<br />

The experiments of Helmut and his colleagues show that the body<br />

takes notice of growing brain tumors; they attract the attention of<br />

microglia, precursor cells, and many other types. Once they arrive on<br />

the scene, relationships become complex. The precursor cells clearly<br />

act as guards, whereas microglia might even be helping in their<br />

escape. There are a lot of molecular conversations going on in the border<br />

zone of the tumor. Keeping gliomas contained may not require listening<br />

in on all of them – but it will certainly require defining the role<br />

of those that are present, distinguishing those who are fighting the<br />

tumor from those who are helping it.<br />

��<br />

The city of Konstanz, situated on the great lake in southwestern<br />

Germany, is home to a unique museum within a museum. In 1870<br />

Part one: The live of nomads 60<br />

the pharmacist and city councilman Ludwig Leiner moved his enormous<br />

collection of fossils, human artifacts, minerals, and stuffed animals<br />

into an old house belonging to the butcher’s guild. Displays were<br />

mounted on the wall or placed in glass display cases, waist-high. Each<br />

object was labeled with a descriptive card, made out by hand in a perfect<br />

cursive script.<br />

Today one large room of the Rosgartenmuseum holds Leiner’s exhibits,<br />

perfectly preserved – except for smudges on the glass where parents<br />

have lifted their children to peer at dinosaur bones and trilobites. Now<br />

that the site has been declared a World Heritage landmark, nothing<br />

can be moved or loaned out. Entering the exhibit is like a stroll into the<br />

19th-century mind, with its fascination for collecting and classifying<br />

and its endless curiosity regarding the natural world. The museum<br />

was established only 11 years after the publication of On the Origin of<br />

Species, and it reveals how eagerly Darwin’s theories were embraced in<br />

Germany by natural scientists and fanatical collectors such as Leiner.<br />

Fossils and stones are enthusiastically dated as being tens or hundreds<br />

of millions of years old.<br />

I think Helmut Kettenmann would feel perfectly at home in the<br />

Rosgartenmuseum. Visiting his office is like time travel (just ignore<br />

the computer). Helmut has become one of the city’s preeminent historians<br />

of science. His shelves are lined with leather-bound volumes,<br />

many of which date to the 19th century; they give off the faint and lovely<br />

smell of old books that have been well cared for. Most of them deal<br />

with the brain, and many of them have a connection to the history of<br />

Berlin. From the bottom shelf he produces an enormous folio with<br />

reproductions of Ludwig Hoffman’s original architectural drawings for<br />

the hospitals of Buch. As he talks about one of the diagrams, he seems<br />

to be seeing it in his mind, the way it used to be.<br />

The Berlin-Buch campus doesn’t have a Leiner room, but a small historic<br />

laboratory has been recreated in the Oskar and Cecilia Vogt


house, complete with instruments that have been used at various<br />

phases in the campus’ history. It’s one stop on the tour when the Long<br />

Night of Science comes every year to campus, an event that attracts<br />

thousands of visitors. Helmut has often been deeply involved in planning<br />

the event and giving talks to the public. Sometimes when he<br />

gives scientific talks he brings along some of his old books.<br />

He also collects antique scientific instruments; several of his microscopes<br />

are on display outside the lecture rooms of the Max Delbrück<br />

Communications Center. And then there are antiques of a completely<br />

different nature – I have heard that he drives in historic automobile<br />

rallies. I ask him about it.<br />

“Oh, that’s private,” he smiles.<br />

As I pack my things to leave, I take a last look at the office. A hundred<br />

years ago, every office would have looked this way. A book was a precious<br />

thing, and the phases of your career could be read from the titles on your<br />

shelves. Your professor gave you one and you would carry it along for the<br />

rest of your life, from station to station of a scientific career. You would<br />

keep it close at hand, where it could be consulted at any time.<br />

I think of 60 copies of Santiago Ramón y Cajal’s small journal, spreading<br />

slowly to the world by carriage and ship and train. I think of the<br />

fact that Charles Darwin’s library held a copy of a scientific journal<br />

which mentioned some experiments on heredity in peas carried out by<br />

a monk named Gregor Mendel. In the 19th century, journals often<br />

arrived with their pages uncut – they had to be sliced apart with a letter<br />

opener to be read. The pages of Darwin’s copy had never been cut.<br />

Mendel never found an Albrecht von Köllinger to launch him into<br />

fame. Darwin might have played that role, but he never witnessed the<br />

birth of genetics. He exchanged letters with Karl von Nägeli, whom<br />

Mendel looked to for advice, but Nägeli never mentioned the monk’s<br />

work.<br />

In the 19th century every scientist had to know where his science came<br />

from and see it as part of a much longer story. Things are not the same<br />

today. Landmark discoveries are flying by much faster; there is so much<br />

to learn about the science of this year, and maybe last year, that few<br />

have time to go deeper.<br />

This is another surprising thing about the campus: in contrast to so<br />

many modern scientific institutes, history is omnipresent here, from<br />

the buildings (which represent diverse architectural styles from most<br />

decades of the 20th century), to the sculpted heads of people who have<br />

defined the the campus, mounted outside the buildings: the Vogts,<br />

Max Delbrück, and many others. There is a Holocaust memorial along<br />

one of the wooded paths – a grim reminder of what can happen when<br />

science becomes the slave of politics, or when it fails to impose ethical<br />

boundaries on itself. The brains of victims of the Holocaust were<br />

examined here in Buch. It is a good thing to walk by the statue from<br />

time to time, to think, and remember.<br />

I am standing in front of it when group of young PhD students walks<br />

by on their way to lunch, heads bent, deep in discussions about their<br />

work or the upcoming weekend. Today they take no notice of the<br />

sculptures. But at some time during their stay here, I know they will<br />

hear of this and the rest of the history of the campus, the story that<br />

they are becoming a part of. Helmut Kettenmann, Heinz Bielka, Jens<br />

Reich and a few others will see to it.<br />

61 Part one: The live of nomads


Stem cells in a jar<br />

Berlin, like most industrial cities, has become one of those places<br />

where you can get almost any type of food at any time. In this culture<br />

of instant grocery gratification, the art of canning food has almost<br />

disappeared. Elena Vasyutina’s mother still knows how to preserve<br />

fruits and vegetables by cooking them, sealing them in jars, then laying<br />

up reserves to be doled out slowly over the winter months, but it’s<br />

becoming a lost art. The best Elena can do, she admits, is to read the<br />

expiration dates on packages.<br />

Our bodies, on the other hand, remain experts at conserving, stockpiling,<br />

and rationing cells, and Elena and her colleagues in Carmen<br />

Birchmeier’s lab at the <strong>MDC</strong> are helping to reveal how this is accomplished.<br />

“Some of our cells are more exposed and prone to injury than<br />

others, and various types have different lifespans,” Elena says. “Shortlived<br />

cells needed to be replaced. To solve this problem our bodies keep<br />

reserves of cells which have specialized part of the way but which have<br />

not taken the last developmental steps.”<br />

Red blood cells, for example, survive only about 120 days and are then<br />

replaced by fresh cells from the blood marrow. Those stockpiles are<br />

Part one: The live of nomads<br />

62


Carmen Birchmeier


already committed to becoming blood cells,<br />

but there are several types, and the cells are<br />

still flexible about exactly what they will<br />

become. Other tissues draw on other kinds of<br />

stem cells to regenerate themselves. The system<br />

isn’t perfect; the body can’t replace all of<br />

its cells, and the ability to heal decreases with<br />

age. It may be that the reservoirs of intermediate<br />

stem cells are used up, or maybe they<br />

have forgotten the secrets of rejuvenation. It<br />

would be useful to find all the types and learn<br />

the genetic programs that cause them to specialize.<br />

The answers might teach us how to<br />

cope with injuries that the body doesn’t repair<br />

very well because the stem cells are lacking, or<br />

possibly because they have forgotten how to<br />

make repairs.<br />

��<br />

Elena was born “in a small village in Russia<br />

you’ve never heard of,” moved to the city<br />

of Novosibirsk as a teenager, then came to<br />

Germany seven years ago. She joined<br />

Carmen’s lab at the <strong>MDC</strong> to do her doctoral<br />

degree. When she arrived the group was<br />

working on several projects – including molecules<br />

involved in the development of the liver,<br />

Alzheimer’s Disease, and the origins of<br />

muscle.<br />

Why such diverse projects? It turns out they<br />

are related; the molecules that Carmen’s<br />

group set out to investigate several years ago<br />

have turned out to have diverse functions. “It’s<br />

always been hard to do a gene knockout in an<br />

animal,” Carmen says. “Even with all the<br />

improvements that have been made, it usually<br />

takes at least a year between your decision<br />

to look at a particular gene and getting animals<br />

that permit you to study its functions.<br />

Then there can be a disaster – you may not<br />

see anything odd about the animal at all.<br />

After so much work I was always so glad to<br />

find something that we pursued the leads<br />

wherever they took us.<br />

Elena says she wanted something “active”, so<br />

she chose muscle. The muscles of our arms<br />

and legs are born in ball-shaped clusters of<br />

cells called somites which form in the early<br />

embryo. Some of the cells in somites receive<br />

signals that tell them to develop into<br />

nomadic muscle. They travel a very long way,<br />

along established routes, as if embarking on a<br />

pilgrimage. Eventually they arrive at bumplike<br />

structures that are the building sites for<br />

future limbs. There they begin fusing to each<br />

other to make long fibers.<br />

Carmen knew that signals which passed<br />

through Met (the wound-healing pathway of<br />

the last chapter) were needed to send muscle<br />

cells on their way. “Without these signals,<br />

there was no migration,” she says. She shows<br />

me a slide of an early mouse embryo, with<br />

Met protein stained in purple. In the normal<br />

mouse, the cells start off in a cluster and then<br />

disperse. In the knockout mouse, they stay<br />

clumped together. That made sense because<br />

Met was known to make travelers out of<br />

many different types of cells, but it raised<br />

other questions. The cells have to travel an<br />

amazing distance and find very specific tar-<br />

Part one: The live of nomads<br />

64<br />

gets. How did the cells know what route to<br />

take – what signs showed them the way? At<br />

about the time Elena arrived in the lab,<br />

Carmen had started following a promising<br />

lead. Cells that later form muscle in the legs<br />

produce a molecule called Lbx1. It is a transcription<br />

factor: a protein that helps switch<br />

genes on and off, and it is produced by these<br />

particular cells as they specialize to become<br />

muscle.<br />

Carmen’s lab made that discovery when they<br />

removed the gene from the mouse in hopes<br />

of discovering its functions. “There was something<br />

wrong with the mice, but we weren’t<br />

sure what it was,” she says. One of her postdocs<br />

carefully listed the tissues which needed<br />

Met to develop properly. Some types of muscle<br />

were on the list. That rang a bell; Carmen<br />

remembered having heard of similar results<br />

from work on another molecule. When the lab<br />

began looking at muscles in different parts of<br />

the embryo, she had an “Aha” moment.


“There were no muscles in the limbs,” she<br />

says. “No cells were migrating there and so<br />

the muscles never formed.” But cells were still<br />

able to travel to other locations to build muscle.<br />

Losing Lbx1 was sending only this particular<br />

type of muscle astray.<br />

“To know where to go, cells have to read cues<br />

along the way, like following signs,” Elena<br />

says. “Because cells take different routes, they<br />

read different signs and respond to them in<br />

individual ways. Without Lbx1 that wasn’t<br />

happening for these particular cells, but we<br />

didn’t know why. We didn’t know what signal<br />

activated the molecule, or what pathway was<br />

involved. That’s what I decided to try to find<br />

out during my PhD work.”<br />

A few years ago it would have been much<br />

harder to find other molecules involved in the<br />

Lbx1 pathway, but Elena could draw on some<br />

new technology to start her search. Decoding<br />

the genomes of mice, humans, and many<br />

other organisms has given scientists a recipe<br />

book for these species’ genes. Those recipes<br />

have been turned into probes that can compare<br />

gene activity in cells. The technology is<br />

called DNA chips, or microarrays. It plays an<br />

important role in the research of Norbert<br />

Hübner’s laboratory at the <strong>MDC</strong>, and his<br />

group offers a service that helps other labs<br />

use it.<br />

Norbert shows me one of the “chips”. It looks<br />

like a normal glass microscope slide; there’s<br />

nothing unusual about it until he holds it up<br />

to the light. Then row after row of very faint<br />

spots become visible – DNA printed onto the<br />

slide. Each of the spots contains billions of<br />

molecules that act as a probe for one gene.<br />

The spots work as detectors because the<br />

chemistry of a DNA molecule makes it able to<br />

bind to a matching RNA – an RNA molecule<br />

that has been made from that gene. Scientists<br />

extract all the RNAs from cells, wash them<br />

over the slide, and then check to see what has<br />

been trapped. A single DNA chip like this one,<br />

containing probes for all of a mouse’s genes,<br />

can trap all the RNAs made by the cell. This<br />

65


Diana Lenhard<br />

66<br />

reveals which genes are active and which are<br />

silent.<br />

The genes that helped muscle migrate and read<br />

signs ought to be active in the cells that produced<br />

Lbx1. Elena used DNA chips to compare<br />

them to other types of cells and obtained a large<br />

amount of data. A microarray experiment often<br />

has a bewildering outcome, showing that dozens<br />

or hundreds of genes behave differently even in<br />

very similar cells. That was the case this time, and<br />

the only way to find answers was to sift through<br />

the data. Here, too, Norbert could help; his group<br />

has become experts at distinguishing important<br />

genes from false leads.<br />

Many of the genes that had turned up in Elena’s<br />

experiment had already been studied by other<br />

labs for other reasons. One molecule on the list<br />

was particularly intriguing, a protein called<br />

CXCR4. Several laboratories, including that of<br />

Martin Lipp at the <strong>MDC</strong>, had been working on the<br />

protein because of its role in the immune system<br />

– and in AIDS.<br />

��<br />

Viruses usually need the help of a conspirator<br />

to invade cells – like an insider who provides<br />

codes to the alarm system. That’s the case with<br />

HIV, the virus that causes AIDS. It can can only<br />

enter human cells that are outfitted with a receptor<br />

protein called CD4. Since this molecule<br />

appears mainly on cells of the immune system,<br />

those are the first victims of the virus, and that<br />

helps explain why AIDS causes a breakdown of<br />

the body’s defenses against other diseases.<br />

But when scientists mounted the receptor on<br />

animal cells, the virus couldn’t enter. This probably<br />

meant that it needed help from a second<br />

human protein. In 1996 Edward Berger’s lab at<br />

the National Institutes of Health in Maryland,<br />

USA, discovered the identity of this partner –<br />

CXCR4.<br />

CXCR4 didn’t evolve in order to make things easy<br />

for HIV. It’s an accident that the virus takes<br />

advantage of the molecule, which has an important<br />

function in healthy cells. In 1998 Reinhard<br />

Förster, a PhD student in Martin Lipp’s lab,


showed that the CXCR4 receptor is produced,<br />

used, and recycled by many types of immune<br />

cells. These are the body’s hunters; they<br />

migrate through tissues, recognize invaders,<br />

and destroy them. CXCR4 is part of the guidance<br />

and tracking system. It is attracted to<br />

SDF1, a protein produced by other cells.<br />

Elena had to go back and read up on the two<br />

molecules. She discovered that SDF1 acts like a<br />

powerful signaling beacon for cells with<br />

CXCR4. “It’s used by the body in several different<br />

ways. A lot of things go wrong in animals<br />

that have a mutant form of either molecule.<br />

Cells don’t migrate properly to form parts of<br />

the brain, and the parents of egg and sperm<br />

cells – which begin in one part of the body<br />

and have to migrate a long way to the reproductive<br />

organs – also get lost.”<br />

The two proteins also seem to play a role in<br />

cancer. CXCR4 has been spotted on many<br />

types of tumor cells that undergo metastasis.<br />

When these cells migrate, they behave a lot<br />

like white blood cells, and they even seem to<br />

use the same tracking system. Interactions<br />

with SDF1 may explain where these cells go<br />

and why they settle in particular places. Cells<br />

that begin in tumors in the breast, for example,<br />

typically settle in the lymph nodes, the<br />

bone marrow, the lungs, and the liver. When<br />

Albert Zlotnik’s lab at the DNAX Research<br />

Institute in Palo Alto, California, blocked contact<br />

between SDF1 and its receptor in mice,<br />

breast cancer cells had a much harder time<br />

taking up residence in the lymph nodes and<br />

lungs.<br />

Were muscle cells using the same system to<br />

move to the limbs? “SDF1 is produced in the<br />

area of the legs right as developing muscle<br />

cells begin to migrate,” Elena says. “That’s<br />

what you would expect to see in a beacon. A<br />

good test would be to see what happened if<br />

we moved the beacon somewhere else – so<br />

we added SDF1 to locations that don’t normally<br />

produce it.”<br />

The decoy worked; migrating cells were pulled<br />

off their normal route and attracted to the<br />

artificial signal. So the two molecules were<br />

acting as a directional signal for muscle, too.<br />

The discovery gave Elena enough material for<br />

her doctoral dissertation – and a paper in the<br />

journal Genes and Development – but she still<br />

had the feeling she was only touching the tip<br />

of the iceberg when it came to understanding<br />

the migration and development of muscle<br />

cells. A lot of questions remained to be<br />

answered. For example, how did cells interpret<br />

multiple signals? Met was also needed to<br />

prompt migrations. Did the two pathways<br />

work together?<br />

Carmen had been working on a protein called<br />

Gab1, a molecule discovered in Walter’s lab in<br />

1996. Gab1 receives the signal from Met and<br />

passes it on to other molecules. Removing<br />

Gab1 from mice also disturbed muscle migration<br />

and development. When Elena studied<br />

mice that lacked both CXCR4 and Gab1, she<br />

found that cells no longer migrated to form<br />

some types of muscle in the tongue. But<br />

when only one of these molecules was missing,<br />

the cells behaved properly. This showed<br />

that there was some kind of connection<br />

between the two pathways.<br />

Once the cells arrive at their destinations,<br />

they reproduce to create enough muscle. They<br />

also develop further, changing their structure


Muscle tissue from normal mice (left) and animals lacking the molecule RBP-J (right).<br />

Normal mice have muscle stem cells called satellite cells (red) but mutants do not.<br />

so that they can fuse and make muscle fibers.<br />

These processes are also controlled by signals,<br />

and Carmen had been collecting evidence<br />

that the Gab1 and Met pathways were<br />

involved.<br />

��<br />

The switch-like behavior of stem cells<br />

became the topic of Elena’s next project,<br />

which she worked on with Diana Lenhard, a<br />

PhD student who had recently joined<br />

Carmen’s group. This time the theme wasn’t<br />

migration but repair: how muscles recover<br />

from injuries. And it involved a signaling system<br />

that is used again and again, throughout<br />

the body, over an entire lifetime.<br />

Notch is a receptor protein like Met. It floats<br />

like a fishing cork on the cell surface. Part of it<br />

rises above the surface, where it can receive<br />

signals; another part passes through the<br />

membrane, and there is a tail which hangs<br />

inside the cell. When the outer part comes in<br />

contact with a signal on another cell, the protein<br />

is cut. This frees the tail, which travels to<br />

the cell nucleus and works with other mole-<br />

cules to activate genes. Which genes it affects<br />

– and what happens to cells – depends on its<br />

partners.<br />

A new Notch partner, a protein called RBP-J,<br />

was discovered a few years ago by a Japanese<br />

group. Usually RBP-J is bound to other molecules<br />

that help it keep genes inactive. But<br />

when Notch’s tail is freed, it bumps off these<br />

other proteins. RBP-J can then link up with<br />

other partners, and the new combinations<br />

become gene activators.<br />

“Several things got us interested in the relationship<br />

between Notch and muscle,” Carmen<br />

says. “First, Notch is very important in the<br />

development of somites – remember, those<br />

are the ball-shaped clusters of cells where<br />

some types of muscle cells are born. Secondly,<br />

as we age, our muscles heal poorly. That’s<br />

partly due to changes in the protein which<br />

activates Notch. This molecule is called Deltalike-1,<br />

and it appears on the surface of neighboring<br />

cells. When there has been an injury,<br />

muscle cells produce Delta-like-1 and this triggers<br />

Notch in their neighbors. But as we get<br />

older, injured cells produce less and less of the<br />

Part one: The live of nomads<br />

68<br />

protein. The signal is much weaker and so<br />

there is less healing.”<br />

Elena and Diana developed a new strain of<br />

mouse that shut down RBP-J only in migrating<br />

muscle cells. They did this by attaching<br />

switches to the RBP-J gene in one type of<br />

mouse and then adding a trigger to the Lbx1<br />

gene, the gene that is only active in migrating<br />

muscle cells, in another type. Mating the two<br />

strains would give them mice in which Notch<br />

could no longer use RBP-J to get signals to the<br />

genes of migrating muscle cells.<br />

But getting such offspring sometimes takes<br />

months. “That’s a long time to wait, especially<br />

if there is no guarantee that the animals will<br />

shed light on the process you are interested<br />

in,” Diana says. “It’s why Carmen likes us to be<br />

working on two or three projects at a time; if<br />

one of them doesn’t work out, you have something<br />

to fall back on.”<br />

Finally the litter was born, and some of the<br />

mice pups had both elements needed to control<br />

the gene. Elena and Diana looked carefully<br />

at the mice for changes in muscle. Losing


RBP-J didn’t have any effect on cell migrations;<br />

half-way through embryonic development,<br />

the normal number of muscle stem cells had<br />

traveled from the somites to the limbs. But<br />

once they were in place, something unusual<br />

began to happen. After one day there was a<br />

sharp drop in the number of stem cells. At the<br />

same time, the scientists found a higher number<br />

of developed muscle cells.<br />

“This seemed to mean that the stem cells<br />

were developing earlier than usual,” Diana<br />

says. “At first that might not seem like such a<br />

big thing, but it soon started to cause problems<br />

for the mice. A short time later the mice<br />

without RBP-J had considerably less muscle in<br />

the limbs, back, and diaphragm than normal<br />

embryos.”<br />

Just as canned foods have to be rationed to<br />

last through the winter, stem cells need to be<br />

rationed to build muscles and repair them<br />

later in life. “Muscle keeps a stock of stem<br />

cells, called satellite cells, in reserve,” Diana<br />

says. “When we investigated the muscle in<br />

our mutant mice, we found that these cells<br />

were gone. The animals could no longer maintain<br />

the cells; they specialized wildly, too<br />

quickly, so they were all used up.”<br />

RBP-J was acting as a “gene brake,” blocking<br />

the final specialization of the cells so that the<br />

animal keeps some reserves and form satellite<br />

cells. Without the protein, the cells aren’t<br />

around long enough to build proper muscles<br />

in the embryo. This shows that the signal<br />

passed along by the Notch protein is crucial in<br />

setting aside a pool of cells that can be used<br />

to build muscle in embryos and repair it<br />

throughout the animal’s lifetime.<br />

Work by other labs supports the idea that<br />

RBP-J may act widely, in many types of tissue,<br />

as a brake that sets aside supplies of stem<br />

cells for future use. That’s probably the case in<br />

the brain, where nerve cells produce Deltalike-1<br />

as they mature. This activates Notch,<br />

which triggers RBP-J. It docks onto genes and<br />

prevents neurons from taking their final<br />

developmental steps. Things slow down and<br />

the nervous system preserves a population of<br />

stem cells.<br />

This system, too, wears down with age. After a<br />

child has reached the age of two, the body has<br />

a hard time making new neurons. The reason<br />

may be similar to what happens in muscle:<br />

there may be defects in the braking system. If<br />

stem cells develop too quickly, the reserves are<br />

used up. Then if the brain is damaged in a<br />

stroke, Parkinson’s disease, or many other diseases<br />

in which neurons die, they can’t be<br />

replaced.<br />

Could RBP-J be used to change that situation?<br />

Knowing that brakes exist doesn’t mean you<br />

can build them or install them in a car, Elena<br />

says. But it’s crucial to know how they work.<br />

Researchers hope to develop therapies in<br />

which they transplant stem cells into damaged<br />

brains, spinal cords, or muscle. The new<br />

immigrants will need to understand the signals<br />

that tell them how to behave, but they<br />

may face the same problems as native cells.<br />

The body may no longer produce the signals<br />

needed to store cells. An alternative – if scientists<br />

can learn how – might be to fit the wanderers<br />

with their own brakes.


Capturing a firefly<br />

IIt is an unusual place – a cool, empty room of muted colors. The walls<br />

are covered with pale green tiles. In the middle of the floor is a hole,<br />

a small crater, as if a tiny meteorite has struck. More likely a drain has<br />

been dug open for repairs. Above the hole hangs an odd lamp with large<br />

bulbs, mounted to the ceiling on a mechanical arm.<br />

To me it is just a strange, abandoned space. But when Peter Schlag<br />

stands in the middle of the room he sees the ghosts of the past. This is<br />

the theater where, as Chief of Surgical Oncology of the Charité, he performed<br />

thousands of operations. A few months ago the Robert Rössle<br />

Cancer Clinic bustled with doctors and patients; now it is almost completely<br />

deserted. Everything has been moved across the street into the<br />

new HELIOS Clinic, a modern hospital complex that will soon host over<br />

a thousand beds.<br />

Peter snaps off the lights and leads the way through long, quiet hallways,<br />

down a flight of stairs, into another corridor. Our footsteps echo<br />

in the silence. He knocks on a white door and enters. A colleague sits at<br />

a computer beside a blank television screen and a row of video cassettes.<br />

“Films of operations,” Peter explains. He gestures at the other<br />

equipment in the room. “This was a communications center. From here<br />

70


Peter Schlag<br />

we remotely directed operations all over the world, on oil rigs and other<br />

hard-to-reach places.”<br />

Down the hallway he tries another door; it’s locked and none of his keys<br />

work. The locks have just been changed. A colleague produces a huge<br />

ring of keys and tries one after another until the door opens. Inside we<br />

find a table mounted with lasers and lenses, an elaborate microscope.<br />

“We’re right below the operating room,” Peter says. “This installation is<br />

why I decided to come here.” The idea, he explains, was to give the surgeon<br />

an extra pair of eyes. The microscope would be able to illuminate<br />

fluorescent molecules that had been attached to markers for deadly<br />

cancer cells. Usually those remain invisible, indistinguishable from<br />

healthy cells. Left in a patient, they may spawn tumors and metastases,<br />

sometimes years later.<br />

Peter dreams of rendering those cells visible. “At the moment, for many<br />

types of tumors, it is extremely difficult to give a patient a prognosis<br />

about whether his cancer will return,” he says. “A prerequisite to completely<br />

removing a cancer is to see all the tumor cells, particularly<br />

those that will metastasize. That would be a huge step forward.”<br />

��<br />

Asurgeon necessarily has a different perspective on “molecular medicine”<br />

than other researchers on campus. Peter gets to see things<br />

from both sides. His position itself is a bridge between the worlds of<br />

clinical and basic research; he is professor at the Charité, the hospital of<br />

Berlin’s university, and a researcher at the <strong>MDC</strong>. Several other scientists<br />

have joint appointments at these two campus partners. Others split<br />

positions at the <strong>MDC</strong> and the third research institute on the Buch campus,<br />

the Leibniz Institute for Molecular Pharmacology (FMP).<br />

Back in his office, Peter orders tea and begins to initiate me into a new<br />

world view. “Molecular biology breaks bodies and bodily systems down<br />

to their basics, that’s what it’s for,” he says. “It’s a set of techniques used<br />

to understand fundamental questions. We need the techniques.<br />

Knowledge is essential, but you can’t treat a patient only with knowledge.<br />

A patient is a whole. You don’t do only oncology – you do diagnosis,<br />

you do surgery, you do psychooncology.”<br />

71 Part one: The live of nomads


The molecular sciences have a great potential to improve diagnostics,<br />

he says. “Surgeons will always be needed, but as Theodor Billroth said,<br />

‘Surgery is the worst of all kinds of treatment.’ Have you heard of<br />

Billroth? He was born close to Berlin, in Bergen, lived and worked in<br />

Vienna, where he developed a great number of new techniques – he’s<br />

known as the father of modern general surgery. He was also an excellent<br />

amateur pianist and violinist by the way, and one of the closest<br />

friends of Johannes Brahms.”<br />

We talk about history and music for awhile. Billroth studied in<br />

Göttingen and Trieste, where he<br />

learned dissection by studying the<br />

nerves of an odd creature called the<br />

torpedo fish. (One of the <strong>MDC</strong>’s scientists<br />

has worked on the animal<br />

and it reappears later in this book,<br />

in the chapter called “the electrician’s<br />

toolbox”.) Billroth received his<br />

doctoral degree in Berlin and<br />

worked in the university’s surgical<br />

clinic for several years before moving<br />

to Vienna. There he developed<br />

successful new operative techniques<br />

to cope with a wide range of<br />

diseases – becoming the first to<br />

remove a colorectal cancer.<br />

“Billroth was a great surgeon, but<br />

he very clearly recognized that surgery<br />

is the front line of an undesirable war,” Peter says. “It’s a last resort,<br />

it’s what you need if there’s been a car crash. It’s not really a solution<br />

for cancer, which is harder than most other diseases. Cancer will never<br />

be like tuberculosis, which is caused by a bacteria. Fighting tuberculosis<br />

is hard, but there’s an identifiable enemy. Cancer is more heterogeneous;<br />

we need to understand the pathways, the mechanisms, the preconditions.<br />

We’re finding new drugs that influence tumor growth, and<br />

in some cases it’s possible to block metastases for months. But as<br />

promising and helpful as these discoveries are, they are not a real solution.<br />

The only way we will learn to block metastases for years is to<br />

intervene in the mechanisms that underlie them.”<br />

Peter came to the <strong>MDC</strong> from Heidelberg 15 years ago and alongside his<br />

duties in the clinic, has run a research lab the whole time. “Even during<br />

surgery I have tried to see the ‘science’ aspect of cancer,” he says. “The<br />

operating room is our research lab. Having this facility on campus has<br />

given us unique opportunities. One of the major questions in cancer<br />

research is what changes cells undergo when they migrate to new<br />

locations in the body and establish new tumors there. Well, we could<br />

collect tissue from tumors in different places in the same patient at<br />

Part one: The live of nomads<br />

different stages of the disease. We’ve had direct access to the samples<br />

you need to answer those types of questions.”<br />

The publications he has worked on reveal both aspects of his career.<br />

Alternating with articles on surgical techniques (“Sentinel lymph node<br />

biopsy for gastrointestinal cancers,” and “Image-guided surgery of liver<br />

metastases by three-dimensional ultrasound-based optoelectronic<br />

navigation”), he has worked on the basic processes that underlie<br />

metastases, and reviews that combine the perspectives (“Clinical, biological,<br />

and molecular aspects of metastasis in colorectal cancer”). The<br />

more basic projects are carried out<br />

by three groups in his department,<br />

headed by Ulrike Stein, Wolfgang<br />

Walther and Wolfgang Kemmner.<br />

Conducting such a wide range of<br />

projects requires interdisciplinary<br />

expertise, so the groups include<br />

doctors, basic researchers, computer<br />

specialists, mathematicians, and<br />

laser specialists.<br />

One project has been the development<br />

of new technology to help<br />

surgeons recognize and cope with<br />

dangerous cells. The group invented<br />

a new “jet injection” device, a<br />

method of dispersing DNA directly<br />

into tissues. The idea for the technology<br />

began during a trip that Ulrike and Wolfgang Walther made to<br />

the United States over a dozen years ago. They spent two years at the<br />

National Cancer Institute in Maryland, collaborating with Priscilla<br />

Furth at the University of Maryland in Baltimore.<br />

“Priscilla was very interested in gene transfer applications – ways of<br />

delivering foreign DNA to cells,” Ulrike says. “The purpose is to get the<br />

cells to absorb a new gene and use it to make proteins. There are several<br />

reasons you might want to do this. If a person had a mutation that<br />

caused a genetic disease, you could give them a healthy copy of the<br />

gene. Another use would be to mark cancer cells so that they could be<br />

seen and removed during surgery. Or you could feed tumor cells a ‘suicide<br />

gene’ that would make them destroy themselves.”<br />

Upon their return, Ulrike and Wolfgang began working on the apparatus,<br />

a needle-less jet-injection device built by the company EMS<br />

Medical in Switzerland. It recently passed phase one trials in humans,<br />

in which a harmless marker gene was jet-injected into tumor tissues.<br />

“We sat back and held our breaths,” Ulrike says. “There were a lot of reasons<br />

why it might not have worked. But not only did the cells take up<br />

the gene and use it to produce proteins – those molecules were func-<br />

72


tional, and there were no side effects.” Soon the method will undergo<br />

its next test at the Charité downtown, in an attempt to deliver a therapeutic<br />

gene to cancer cells.<br />

Meanwhile, the groups have been digging into the biology of metastases.<br />

For most types of cancer, it is still impossible to tell the difference<br />

between these dangerous nomads and benign tumor cells, and sometimes<br />

even healthy cells. But their unique behavior must be guided by<br />

a unique set of molecules. Identifying them, Peter says, is the first step<br />

toward seeing them, and maybe finding ways to destroy them.<br />

I think of summer evenings as children, when we spent hours catching<br />

fireflies and putting them in jars, then carrying them around like<br />

lanterns in the night. Even if a firefly was sitting still on a branch, you<br />

couldn’t see one, or catch it, until it glowed. The same thing will probably<br />

be true of metastases – they will have to be marked to be trapped,<br />

studied, and conquered.<br />

��<br />

Ulrike attended the university in Halle when it still belonged to the<br />

GDR. Only two universities offered degrees in biochemistry, and<br />

there were places for only 40 students a year; excellent grades won her<br />

one of those places.<br />

Ulrike’s son is now studying at the Technical University of Berlin. She’s<br />

a professor and she can’t help comparing her own university experiences<br />

to those of her son. “Students have so many more opportunities<br />

now,” she says. “They’re free to take any courses they want. Our time<br />

was very strictly managed; there was no time to do anything extra. On<br />

the other hand, there was much more in the way of social structure. We<br />

were a small, tight-knit group; we did everything together – we helped<br />

each other both academically and psychologically. Today things are<br />

very competitive and individualistic and it isn’t always very healthy.<br />

Students have to manage their own studies – they even have to take<br />

care of a lot of organizational problems that the university used to<br />

arrange for us.”<br />

She is also in an ideal position to study the campus’ marriage of medicine<br />

and fundamental research. She teaches biology courses for med-<br />

73 Part one: The live of nomads


Ulrike Stein


ical students and always has some in her research group. They have<br />

been particularly motivated and she has rarely had problems integrating<br />

them. Ulrike admits that she has been a bit spoiled, because as a<br />

rule, the physicians of the Robert Rössle Clinic have been very interested<br />

in research.<br />

But she acknowledges that there have been a few culture clashes –<br />

sometimes amusing ones. “Writing a dissertation for a PhD in molecular<br />

biology often takes years,” she says. “MDs sometimes finish their<br />

theses in a semester. Once an MD student applied to the group and I<br />

asked him how much time he<br />

would be able to devote to his dissertation.<br />

He thought about it a<br />

moment and said, ‘I can work on it<br />

every Thursday afternoon, for a<br />

year.’” She laughs. “It’s a running<br />

joke in our lab now. If a really hard<br />

problem comes up, somebody says,<br />

‘Sure, I’ll take care of it on Thursday<br />

afternoon.’”<br />

Working in a medical clinic with<br />

Peter Schlag has been a tremendous<br />

experience, she says. “It’s<br />

exceptional to find a surgeon who<br />

takes such a passionate interest in<br />

the biology underlying what he’s<br />

doing,” she says. “I’m very thankful<br />

I could work in this environment.<br />

As a biologist you dream that what<br />

you are doing can be turned into<br />

therapies, but working with physicians, you start to learn what makes<br />

sense and what doesn’t, what can be applied to patients and what<br />

can’t. In biology we often don’t think things through to the end in clinical<br />

terms; you can spend years pursuing something that may not be at<br />

all applicable to patients in the way you imagined.”<br />

Significant differences in the education systems of physicians and<br />

researchers can lead to misunderstandings, she says. “But these things<br />

are usually easy to solve if there is direct contact, and if both sides<br />

make an effort to understand each other.”<br />

��<br />

Recently she and her colleagues have made a breakthrough with<br />

the discovery of a molecule that helps transform tumor cells into<br />

metastases. Ulrike was under pressure to finish the paper. She looked<br />

tired when she came by to talk about the manuscript.<br />

“Until recently it was difficult to define the difference between a nonmetastasizing<br />

cell in the primary tumor and one that would metasta-<br />

size,” she says. “The old view was that a cell’s DNA suffered damage;<br />

there was a mutation that caused a tumor. Then when the tumor had<br />

grown for awhile, there was a second event, probably another mutation,<br />

that caused one of the cells to metastasize. If that is really how<br />

things work, you can analyze patient tissues all you want and you may<br />

never find the one cell in a tumor that will turn into a metastasis. It<br />

may be only one among billions of cells, and it might change just<br />

before it casts off and wanders to a new tissue. The chances of seeing<br />

that happen would be very small.”<br />

But Ulrike and many scientists are<br />

no longer sure that this scenario is<br />

correct. “We’re starting to think<br />

that there are different types of<br />

cells in the tumor from the very<br />

beginning. They may begin as a<br />

sort of stem cell that moves into<br />

an area, replicates quickly, and<br />

then undergoes a process of development.<br />

Some of those cells<br />

become migrating cells that move<br />

away. That happens over and over<br />

during normal embryonic development,<br />

and in that situation it isn’t<br />

mutations that make cells migrate.<br />

Rather, it’s a change in signaling<br />

and the activation or repression of<br />

genes. If you look at the question<br />

that way, it may not be only a mutation that causes a metastasis.<br />

Instead, we’re looking for signals and a pathway that become active in<br />

some of the tumor cells.”<br />

In other words, metastases may behave like muscle stem cells which<br />

arise in a distant part of the body and then migrate to the limbs, the<br />

process described in the previous chapter. The difference is that cancer<br />

cells are defective, and upon arrival they don’t develop into a useful<br />

structure. Instead they form a mass of tissue that eventually interferes<br />

with the functions of healthy organs. And the process may start all<br />

over again at the new site. Cells may detach themselves and travel<br />

elsewhere to form further tumors.<br />

The focus of the lab’s work has been colorectal cancers, a disease which<br />

is on the rise in Western countries. “About fifty percent of these tumors<br />

can be treated successfully through surgery, but it is notoriously difficult<br />

to tell whether the cancer will return or metastasize,” Peter says.<br />

“Diagnostic tools that help in other forms of cancer don’t work well<br />

with this type. Since metastases are the major cause of death, what<br />

75 Part one: The live of nomads


we’d like is a way to scan tumors for signs that they will develop this<br />

way. If we find such markers, we’ll know which patients need careful<br />

monitoring and further treatment and which don’t.”<br />

Peter’s surgical clinic gave Ulrike and her colleagues unique access to<br />

samples and data. “He has strongly supported us from the clinical side<br />

the whole time,” she says. “He provided us with tumor tissues taken<br />

from patients at various stages of the disease. We had access to even<br />

more samples in the ‘tumor banks’ that were started by Peter, the<br />

Robert-Rössle Clinic and the <strong>MDC</strong>. This unique repository of tissues<br />

from cancer patients allows the validation of experimental hypothesis<br />

in physiological situations and creates an important link to perform<br />

translational research. We could examine colorectal tumors that<br />

seemed benign, those that were not metastasizing, and cancerous tissue<br />

that had already spread to other parts of the body. It was really<br />

valuable to be able to look at several samples from the same patient.”<br />

Peter also provided a biostatistician, Markus Niederstrasse, who<br />

worked hard on the analysis. After comparing the complete gene activity<br />

of different types of cells, one gene stood out; metastatic cells were<br />

producing much higher amounts of a particular protein. The molecule<br />

had been discovered when comparing the RNA molecules produced by<br />

non-metastasizing and metastasizing tumors, but it hadn’t been stud-<br />

Part one: The live of nomads<br />

ied and didn’t yet have a name. Ulrike and her colleagues could now<br />

give it one: they called it Metastasis-Associated in Colon Cancer 1, or<br />

MACC1. Ulrike calls it “Maxie.”<br />

The protein has turned out to be a very reliable sign that a colorectal<br />

tumor will metastasize, or that it has already done so. Patients whose<br />

tumors did not produce the protein usually had good prospects for<br />

a lasting recovery. But when their tissue showed high levels of the<br />

molecule, the prognosis was bad; metastases were almost sure to<br />

develop.<br />

Was MACC1 a key to tumor cell migrations, or just a side effect of some<br />

other process? Ulrike and her colleagues fed the gene to cells in laboratory<br />

cultures; when the cells began to produce the protein, they cast<br />

off ties to their neighbors and began to crawl away from each other.<br />

When the scientists added the human gene to tumors in animals,<br />

metastases formed in the liver. So the protein was directly involved in<br />

the cell migrations.<br />

In another experiment the scientists removed parts of MACC1 that<br />

allowed it to communicate with other proteins. “A signaling protein is<br />

like a machine with a transmitter and receiver,” Ulrike says. “If you take<br />

away either part, it can no longer hear messages or pass them along.<br />

When we removed these components of Maxie, it no longer stimulat-<br />

76


ed crawling or caused metastases. We had successfully blocked whatever<br />

signal was making the cells migrate. That convinced us that the<br />

protein was something like a master regulator of metastasis.”<br />

The way the cells migrated in culture dishes resembled what Walter<br />

Birchmeier had seen in his work with another signaling protein – Met.<br />

Maybe there was a connection – maybe Met was passing information<br />

to MACC1, or vice versa. “We had indirect evidence for this because cells<br />

that produce MACC1 also make Met,” Ulrike says. “So the next step was<br />

to prevent the cells from making any MACC1. Immediately they reduce<br />

their production of Met as well.”<br />

The recipe for the Met protein is contained in one of our genes, but few<br />

cells use it. Like other genes it is usually silent until another protein<br />

called a transcription factor binds to it and activates it. Experiments<br />

showed that the amount of Met produced by the cell directly depended<br />

on how much MACC1 was being made. A careful study of MACC1<br />

showed that it could dock onto the Met gene – it was acting as a transcription<br />

factor.<br />

��<br />

The next step will be to try to find something to block MACC1’s signaling<br />

activity in tumors in mice, and eventually human beings.<br />

This is the point at which basic research might become transformed<br />

into real help for patients. With a great deal of luck, an existing drug or<br />

substance might disrupt the protein’s activity. But the chances of an<br />

easy success are very slim. An effective drug has to be precise – it<br />

should target only the disruptive molecule, only in cells that are misbehaving.<br />

The first problem is hard enough, although Ulrike says that<br />

MACC1 has features that might make it susceptible to drugs. A second<br />

hurdle is delivery: getting the substance to just the right cells in sufficient<br />

quantities. Many believe that some sort of new conceptual breakthrough<br />

will be necessary to solve it.<br />

That may be on the verge of happening. In the past few years scientists<br />

have discovered that you don’t necessarily have to rewrite the recipe in<br />

a gene to keep a cell from producing a particular protein. There’s an<br />

intermediate step; genetic information is first rewritten into a form<br />

called RNA before proteins are made. The idea is explained in more<br />

detail in the next chapter, but it basically involves building an artificial<br />

RNA molecule that interferes with the one the cell has made. Ulrike<br />

and her colleagues have tried this with MACC1 in laboratory cell cultures.<br />

With a single treatment they have managed to block cell migrations<br />

for months.<br />

The group is now using genetically engineered mice to see if MACC1<br />

RNA can be blocked in animals, and if so, whether that prevents the<br />

development of metastases. Even if the strategy works, Ulrike admits,<br />

it will be a long road to doing something similar in humans. Once<br />

again, the hurdle will be to shut down the molecule only in tumor cells,<br />

because blocking it everywhere in the body would surely have unwanted<br />

side effects. “We don’t know what the healthy functions of Maxie<br />

are,” Ulrike says. “But it is produced in a wide range of healthy tissues,<br />

where it probably has important jobs.”<br />

Pure knowledge isn’t a cure, as Peter Schlag says. Yet it may still have a<br />

powerful impact on health. At the very least, MACC1 may help accomplish<br />

something that has been impossible so far. It will probably make<br />

an excellent tool to predict whether a patient’s colorectal tumor is likely<br />

to spread and take on a much more deadly form. Metastases can’t<br />

yet be caught and trapped in a jar, like fireflies. But maybe soon surgeons<br />

will be able to see where they alight.<br />

77 Part one: The live of nomads


A very quiet cure<br />

In 1952 composer John Cage amused, mystified, and enraged the public<br />

with the first performance of his work Four minutes and thirtythree<br />

seconds of silence. David Tudor came onto a stage in Woodstock,<br />

New York, sat down at a piano, and closed the lid of the keyboard. Then<br />

he sat there, staring at his music and a stopwatch. He briefly opened the<br />

lid and closed it again between each movement of the piece. Not that<br />

there was complete silence during the performance; one commentator<br />

said, “Mostly what you could hear was people getting up and walking<br />

out.”<br />

Two events inspired the composition. Cage’s friend Robert Rauschen -<br />

berg had just painted a series of pure white canvases that never looked<br />

completely white because of changing light conditions, the shadows of<br />

viewers, and other factors. (The series may also have inspired the last<br />

painting of Buch artist Jeanne Mammen, which is also white – as it<br />

aged, she said, the surface would gradually turn gold, and underlying<br />

colors would come through.) The second influence was Cage’s visit to a<br />

unique room that had been built at Harvard University, a cork-lined<br />

chamber that absorbed all sound. Upon entering, he hoped to<br />

Part one: The live of nomads<br />

78


experience absolute silence. Instead, he<br />

wrote, “I heard two sounds, one high and one<br />

low... the high one was my nervous system in<br />

operation, the low one my blood in circulation.”<br />

There was no such thing as complete<br />

silence, he realized, and one aim of the new<br />

composition was to demonstrate the impossibility<br />

of achieving it.<br />

In January 2004 the work was performed by<br />

the BBC Symphony Orchestra and broadcast<br />

live on Britain’s Radio 3. (Beforehand the station<br />

had to switch off an emergency system<br />

which breaks in automatically if a long silence<br />

occurs during a broadcast.) Although the<br />

piece has been performed many other times,<br />

it never became wildly popular. That might<br />

have happened if Cage had succeeded in his<br />

plan to sell it to the Muzak company – people<br />

in elevators and department stores probably<br />

would have preferred it to the alternatives.<br />

The popularity of silence has been growing<br />

among biologists, who hope to use it as a<br />

method of treating genetic diseases, cancer,<br />

and a range of other illnesses. “Obviously we<br />

don’t mean silence in music,” says Thomas<br />

Christély. He is Chief Executive Officer at the<br />

79 Part one: The live of nomads<br />

Thomas Christély<br />

company Silence Therapeutics AG, one of the<br />

many private companies on the Berlin-Buch<br />

campus. “Silencing genes means preventing<br />

them from being used to make proteins.<br />

Proteins are usually what disrupt cells and<br />

ultimately cause the symptoms of disease, so<br />

learning to block their production would give<br />

us a handle on cancer, genetic diseases, and a<br />

wide range of other health problems.” As the<br />

last chapter showed, if tumor cells didn’t<br />

make the protein MACC1, they might not set<br />

off on dangerous journeys to other tissues. In<br />

other cases, things break down because<br />

mutations give a protein the wrong shape


and chemistry and it behaves badly. Switching<br />

it off might solve the problem.<br />

One of the challenges in the approach has<br />

been to develop treatments that work as long<br />

as possible. Four and a half minutes of silence<br />

might seem almost unbearably long during a<br />

radio broadcast, but you wouldn’t want to<br />

have to treat a patient thirteen times every<br />

hour. Most therapeutic molecules survive<br />

longer than that, but not long enough, so biologists<br />

have been trying to find ways to<br />

increase their lifespans.<br />

Silence Therapeutics AG is working on a solution.<br />

They are building new kinds of molecules<br />

and packing them in new ways. Their products<br />

don’t have to achieve the length of John<br />

Cage’s longest work, called As slow as possible.<br />

(The piece lasts 639 years and is currently<br />

being performed in St. Burchardi, a church in<br />

the German town of Halberstadt.) But the<br />

company hopes to make therapeutic molecules<br />

that last a very long time.<br />

��<br />

Important discoveries in biology often arise<br />

in unlikely places. In the late 1980s, scientists<br />

in a biotech company in California were<br />

using genetic engineering to try to give pale,<br />

purple petunias a more intense color. Rich<br />

Jorgensen’s lab inserted an extra copy of the<br />

gene responsible for purple pigment, expecting<br />

that more genes would produce more pigment<br />

protein. Instead, the result was a completely<br />

white flower without any pigment at<br />

all.<br />

A look into the plant’s cells showed that they<br />

were actually using both genes to produce<br />

RNAs – the molecules which bridge the gap<br />

between genetic recipes and their final form<br />

of proteins. Somehow RNAs from the two<br />

genes were interfering with each other and<br />

blocking the synthesis of purple pigment.<br />

Labs across the world were discovering the<br />

same phenomenon in other plants, and soon<br />

Part one: The live of nomads<br />

80<br />

it would be observed in animals as well.<br />

Finding out why took several more years.<br />

A DNA molecule is a string of building blocks<br />

called nucleotides, or bases. The cell builds two<br />

strands of DNA with complementary bases –<br />

that means the units in one strand chemically<br />

attract those in the other. The two strands<br />

wind around each other and form a shape<br />

called a double helix. RNAs are made of the<br />

same types of building blocks, but they are<br />

usually found as single strands in the cell.<br />

They could form double strands if the cell produced<br />

two RNAs with complementary<br />

sequences.<br />

That’s what was happening in Jorgensen’s<br />

petunias. The RNAs produced by the two purple<br />

pigment genes were complementary.<br />

When they doubled up, they came to the<br />

attention of a cellular defense system. Many<br />

viruses contain double-stranded RNAs, and<br />

the cell perceives them as alien and dangerous.<br />

When one is spotted, it is attacked by pro-


teins that chop up and destroy RNA molecules.<br />

The first examples of this process turned up<br />

in genetic engineering projects like<br />

Jorgensen’s, but in the meantime this type of<br />

RNA interference was found to happen naturally<br />

in cells. New analyses of the human<br />

genome show that it contains the recipes for<br />

a huge number of RNAs, usually very small<br />

ones, whose sole purpose seems to be to lock<br />

onto other RNAs and prevent them from<br />

being used to make proteins. This gives the<br />

cell yet another system of control over how<br />

the information in genes is used – like giving<br />

a car an extra set of brakes. The discovery was<br />

recognized as so important that Andrew Fire<br />

and Craig Mello, who discovered the process<br />

by which cells dismantle interfering RNAs,<br />

were awarded a Nobel Prize in 2006.<br />

Scientists quickly saw that this could be<br />

turned into a new method of controlling<br />

genes in the lab, and possibly in patients. It is<br />

now routinely used in experiments with ani-<br />

mals. Now Silence Therapeutics and others<br />

are beginning to use the method in the treatment<br />

of human diseases.<br />

��<br />

Surgeons have a different perspective<br />

on molecular medicine than basic<br />

researchers; biotech companies have yet<br />

another. To survive they have to create and sell<br />

products, provide useful services, and convince<br />

investors that they will continue to do<br />

so. Thomas Christély admits that it is a<br />

tremendously competitive field. Success<br />

requires finding an original approach to a<br />

problem that can be realistically addressed<br />

and which fills a niche in the market.<br />

“Our company is doing that by focusing on<br />

two crucial aspects of RNA interference therapies,”<br />

he says. “The first is to make long-lasting<br />

small interfering RNA (siRNA) molecules<br />

that shut down genes involved in important<br />

diseases. The second is to develop a new strat-<br />

81 Part one: The live of nomads<br />

egy for delivering those molecules to target<br />

cells.”<br />

Thomas says the company’s approach is the<br />

result of a lot of hard work by several people.<br />

“The Chief Scientific Officer of Silence<br />

Therapeutics AG is Klaus Giese, who founded<br />

the company in 1998 under its previous name,<br />

‘Atugen AG.’ Other people who have been really<br />

important are his colleagues Jörg<br />

Kaufmann and Anke Klippel, who all studied<br />

for many years on the West Coast in the US,<br />

then worked for Chiron Corporation in<br />

California before they joined the company.<br />

With Oliver Keil, who joined later in 2002, they<br />

jointly invented our proprietary siRNA and<br />

delivery technologies.”<br />

Several factors make RNA interference particularly<br />

interesting, he says. One reason that<br />

drugs are so expensive is that the development<br />

process typically takes an incredibly<br />

long time – ten years or more. “That process<br />

usually starts by identifying a protein that is


Britta Dieckhoff<br />

crucial to the development of an illness – one<br />

that would stop the disease if you could block<br />

it,” Thomas says. “That in itself usually takes<br />

years and for many diseases it hasn’t happened<br />

yet. If you get to that stage, many more<br />

years are usually required to find and refine a<br />

compound that influences the disease<br />

process – if one exists. If you’re successful it<br />

has to be studied for toxic side-effects, and<br />

then undergo a number of other tests before<br />

you can try it in humans. RNA interference is<br />

based on a natural mechanism. It works with<br />

just about every gene, and once you have<br />

identified the one you’re interested in, you can<br />

produce the siRNA in two or three months.”<br />

One big advantage of siRNA molecules is that<br />

the total development time, compared to<br />

conventional drug technologies, can be shortened<br />

by a few years – which is extremely<br />

important from the commercial point of view<br />

of a pharmaceutical company. “If the revenues<br />

from the sale of siRNA drugs are generated<br />

a few years earlier, this will significantly<br />

Part one: The live of nomads<br />

82<br />

improve the cash flow for such drugs” he says.<br />

“In addition, development costs will be significantly<br />

lower due to the shorter development<br />

time. Our siRNA technology will make it easier<br />

to earn back the huge investment involved<br />

in creating a new therapy.”<br />

Companies with conventional therapeutic<br />

technologies usually focus almost exclusively<br />

on cures for diseases that affect huge numbers<br />

of people. Bringing a drug to the market<br />

faster would reduce the price and also entice<br />

companies to work on serious diseases that<br />

strike fewer people.<br />

��<br />

Asmall interfering RNA molecule looks like<br />

a bit like a comb, or a ladder that has<br />

been sawed in half from top to bottom, leaving<br />

one side bar and half-rungs. When it binds<br />

to its partner in the cell, it forms a complete<br />

ladder. The RNAs designed by the chemists of<br />

Silence Therapeutics have a different structure<br />

than most of those made elsewhere.


Jens Endruschat<br />

“A conventional siRNA doesn’t completely<br />

match up to the RNA made by the cell,”<br />

Thomas says. “It’s like a ladder with a longer<br />

bar on one side and a few extra rungs that<br />

don’t connect to the other side, which are<br />

called overhangs. My colleagues Jörg<br />

Kaufmann, Klaus Giese and Anke Klippel have<br />

developed a special kind of molecule called<br />

AtuRNAi which is a small interfering RNA molecule<br />

with ‘blunt ends’ – in other words, without<br />

any overhangs. When it docks onto its<br />

partner, the ends match.” He lists some other<br />

differences: AtuRNAi contains only naturally<br />

occurring RNA building blocks, whereas most<br />

other companies build molecules out of a<br />

mixture of elements from RNA and DNA. The<br />

AtuRNAi molecule can also be produced faster<br />

and less expensively than other types of<br />

siRNAs.<br />

The most significant difference, though, may<br />

be that AtuRNAi is more durable and still<br />

highly potent. It works longer before the cell<br />

breaks it down and a new treatment is<br />

required. The short lifetimes of most RNA<br />

molecules limits their use in therapies. The<br />

lifetime of company’s AtuRNAi molecule has<br />

been extended from a few minutes to a week<br />

or so, which means that less of the molecule<br />

is needed for therapeutic applications.<br />

Another of the company’s innovations confronts<br />

the problem of delivering therapeutic<br />

molecules to diseased cells. “An siRNA has to<br />

survive while it travels through the bloodstream<br />

and it also has to be taken up by cells,”<br />

Thomas says. “This means attaching it to a<br />

delivery vehicle that will both protect it and<br />

then be absorbed once it encounters the target.”<br />

Most solutions that have been tried involve<br />

attaching the molecules to liposomes, small<br />

bubble-like compartments. They are built of<br />

lipids, the fat molecules which make up cell<br />

membranes. In the body, molecules are often<br />

packed into liposomes when they have to be<br />

shuttled between cells, so the approach takes<br />

advantage of the fact that cells know how to<br />

83<br />

Part one: The live of nomads<br />

absorb liposomes. Silence Therapeutics started<br />

with the same basic structure but then<br />

modified it to create something new.<br />

The company’s calls its delivery vehicle the<br />

AtuPLEX. “It’s a combination of AtuRNAi<br />

attached to a specific lipid sphere,” Thomas<br />

says. “The size of our sphere is about one-anda-half<br />

times larger than the liposomes that<br />

are typically used, but since different tissues<br />

usually cope with liposomes of different sizes,<br />

we can alter its dimensions to fit the situation.<br />

One advantage of a large vessel is that<br />

it can be used to deliver several different<br />

small interfering RNAs to the same cell.<br />

That’s important in the many cases where<br />

you would like to shut down several<br />

proteins. For example, it would allow you<br />

to block multiple components of a signaling<br />

pathway.”<br />

Another difference has to do with the liposome’s<br />

electrical charge. Lipid molecules carry<br />

a charge that play an important role in<br />

whether they are absorbed. Cells often let


Britta Dieckhoff<br />

them enter – or block the process – because<br />

they need to balance their own charge with<br />

that of the environment. The AtuPLEX has a<br />

lower charge than most liposomes, which<br />

makes delivery more likely in most situations.<br />

The company has several specific projects<br />

under development, usually with major<br />

industrial partners like Pfizer/Quark and<br />

AstraZeneca. An AtuRNAi built to interfere<br />

with a protein called RTP801 is currently<br />

undergoing Phase I clinical studies in human<br />

patients; the goal is to combat a common disease<br />

called macular degeneration. The macula<br />

is a small yellow spot in the retina that plays<br />

an important role in the sharpness of our eyesight<br />

and how well we perceive colors. In<br />

many people it degenerates with age, a<br />

process that can lead to blindness. Many of<br />

the earliest therapies involving interfering<br />

Part one: The live of nomads<br />

84<br />

RNAs have focused on this disease as a sort of<br />

proof of principle; the protein culprit is clear;<br />

the eye is a closed system.<br />

By the end of 2008, clinical trials of five other<br />

AtuRNAi molecules are scheduled to begin;<br />

two more should follow soon after that. The<br />

new therapies will target cancers of the gastrointestinal<br />

tract, lungs, prostate, and liver, as<br />

well as other very serious diseases.


Some of the preliminary results have been<br />

very promising, Thomas says. The laboratory<br />

of Prof. Bertram Wiedenmann of the Charité<br />

recently used one of the company’s small<br />

interfering RNAs, called Atu027, in a study of<br />

pancreatic tumors in mice. The RNA blocks the<br />

production of a protein called PKN3, whose<br />

role in cancer was discovered in 2004. A company<br />

research group headed by Anke Klippel<br />

discovered that the protein passes along<br />

information from a signaling pathway that<br />

causes metastases in laboratory cell cultures.<br />

Samples taken from patients with prostrate<br />

tumors also showed high levels of PKN3. All of<br />

these factors suggested it might make a good<br />

target for an anti-cancer therapy.<br />

In Wiedenmann’s study, mice with tumors<br />

received ten injections of the interfering RNA<br />

molecule over a three-week period. The<br />

tumors of untreated animals grew at their<br />

normal, devastating pace, whereas their<br />

growth rate in treated animals was significantly<br />

reduced, as was the metastatic spread<br />

of cells. “This is strong evidence that our system<br />

very effectively blocks the production of<br />

cancer-causing proteins and has a huge<br />

impact on the further course of the disease.<br />

It’s also proof that the delivery system works<br />

in live animals. We expect it to be equally<br />

effective in humans,” Thomas says.<br />

��<br />

Silence Therapeutics has labs and offices on<br />

the first and second floor of an elegant<br />

white building on the west side of campus.<br />

Aside from the security downstairs and the<br />

firm’s logo printed on the glass doors, there is<br />

little to indicate this is the domain of a company<br />

rather than a research lab. But the world<br />

of business has its own customs and<br />

etiquette which surface from time to time as<br />

Maj Britt Hansen tries to take photos –<br />

some rooms are off limits – and during<br />

my meeting with Thomas. When we sit down<br />

in a quiet conference room, he is prepared<br />

to talk me through a standard company presentation<br />

that he gave last week at a<br />

BioEurope meeting – a mixture of science,<br />

performance indicators, logos of impressive<br />

partners, and of course money matters.<br />

All outlined in 16 impeccable Powerpoint<br />

slides.<br />

When I tell Thomas I am trying to get a feeling<br />

for the personality of the company and understand<br />

how it fits into the puzzle of a very complex<br />

campus, he shifts gears. He talks a bit<br />

about his own background. You would never<br />

85 Part one: The live of nomads<br />

guess it from talking to him, but he never<br />

studied science – although his majors in high<br />

school were chemistry and biology. He<br />

received degrees in business and law from the<br />

University of Hamburg, worked in mergers<br />

and acquisitions of a Swedish investment<br />

bank in London, then became managing partner<br />

of an investment firm in Moscow. In 1996<br />

he began working as Senior Vice President<br />

and Chief Financial Officer at the Swiss pharmaceutical<br />

company OXO Chemie AG and<br />

founded a subsidiary called OXO Chemie Inc.<br />

in San Francisco. He has been with Silence<br />

Therapeutics since 2001. His career has taken<br />

him from Belgium to England, Switzerland,<br />

Russia, and the US. It’s not only scientists who<br />

are nomads, but the many people associated<br />

with it.<br />

Thomas sees long-term perspectives for<br />

Silence Therapeutics in Buch. “We’re not in the<br />

city center,” he admits. “But being on the campus<br />

has forged relationships with groups who<br />

might discover new targets for therapies, who<br />

can look at the behavior of those molecules in<br />

diseases, and who can carry out clinical trials.<br />

There are opportunities for partnerships at<br />

each stage of the process of turning a target<br />

into a marketable product. It’s rare to find that<br />

range of expertise clustered so that they are<br />

in continual, direct contact with each other.”


Teleportation for beginners<br />

Part one: The live of nomads<br />

It’s not only scientists and cells that undertake epic journeys –<br />

sometimes artifacts from institutes become nomads as well. There<br />

are the leather-bound books on Helmut Kettenmann’s shelves, and the<br />

postal scale in Martin Lipp’s office (which once belonged to Nobel<br />

Prize-winner Feodor Lynen). The farthest-roving travelers are surely the<br />

cult objects of Thomas Jentsch, which sailed the Pacific before settling<br />

in Berlin. But the strangest case is that of the cornerstone of the FMP,<br />

which traveled hundreds of kilometers in a feat that any magician<br />

would admire.<br />

July 13, 1998, saw the groundbreaking ceremony for the new FMP<br />

building on the Berlin-Buch campus. Staff of the institute and the<br />

architects prepared a time capsule, a metal cylinder to be placed inside<br />

the cornerstone. Among the contents were:<br />

· plans of the new building<br />

· a collection of coins (German marks)<br />

· the FMP’s 1996-97 annual reports<br />

· recent Berlin newspapers<br />

· samples of proteins and DNA related to FMP research<br />

The cylinder was lowered into the foundations – all of this observed by<br />

over a hundred witnesses, including reputable representatives of the<br />

local and national government – and later encased in concrete.<br />

Jump to almost exactly five years later and 300 km away. On July 17,<br />

2003, a team from the regional environmental office of Niedersachsen<br />

86


was carrying out a routine investigation of an abandoned gravel pit.<br />

They discovered a mysterious metal container – a bomb? Radioactive<br />

waste? An artifact from a forgotten (but technically advanced) civilization?<br />

They immediately called the police and the fire department, who<br />

came with equipment to investigate toxic waste. An initial investigation<br />

revealed no traces of explosives or drugs. The cannister was<br />

shipped to various laboratories where it was X-rayed and submitted to<br />

a gamut of other tests. Experts attempted to penetrate the container<br />

with an endoscope but were unsuccessful.<br />

After a month someone came up with the bright idea of simply prying<br />

the thing open. Inside were the contents of the FMP time capsule, in a<br />

deteriorated condition. The institute received a call from a press officer<br />

in Niedersachsen, saying, “We think we have something that belongs<br />

to you.”<br />

The time capsule and its contents were returned to Berlin-Buch, and<br />

until recently they could be seen in a display on the ground floor of the<br />

FMP building. If you want to see them now, you’ll have to talk to Björn<br />

Maul.<br />

There are various hypotheses about how the cannister disappeared<br />

from the campus and rematerialized somewhere else. The most likely<br />

explanation – although not the most interesting one – is that the cornerstone<br />

was accidentally removed when the FMP built a new<br />

entrance to give access to the handicaped. (My personal favorite is<br />

that one day the FMP switched on all their NMR machines at the same<br />

time, briefly disturbing the Earth’s magnetic field, and accidentally<br />

invented teleportation.)<br />

You can’t have an institute without a cornerstone, so a new one was<br />

made and outfitted with a time capsule which was filled with, well,<br />

updated versions of the old stuff. (Euros, for example, instead of<br />

German marks.) On June 1, 2004, it was packed into a stone and<br />

embedded in the ground at the FMP. There it remains to this day.<br />

At least, that’s where we think it is.<br />

(freely adapted from an article by Björn Maul)<br />

87 Part one: The live of nomads


Ludwig Thierfelder


Interlude:<br />

A glass heart<br />

Sometimes the history of medicine reads like a long line of unlikely<br />

stories, and this is one of them. It goes like this: a group working on<br />

one side of the campus removes a gene from a strain of mouse and discovers<br />

that the animal is now unable to build a proper heart. The fate of<br />

this small creature has large repercussions: it sheds light on work being<br />

done with human patients on the other side of the campus, leading to<br />

a new way of diagnosing a rare type of heart disease and a simple surgical<br />

procedure that is saving people’s lives as you read this book. It’s<br />

exactly what people mean when they talk about molecular medicine; it<br />

is the way science is supposed to work. Yet it rarely happens so directly,<br />

and when it does, it seems like a minor miracle.<br />

The fragility of the heart has been the topic of many curious stories in<br />

science. One of them dates back 80 years and involves a famous aviator<br />

who decided to dabble in medical research. In May 1927 Charles<br />

Lindbergh became an instant celebrity when he completed the first<br />

non-stop flight across the Atlantic, taking off in New York and landing<br />

in Paris. Four years later he published a small paper in the journal<br />

Science, making his first contribution to the field of cardiology.<br />

89 Interlude: A glass heart


The Carrel-Lindbergh pump<br />

Charles Lindbergh (second from right)<br />

Lindbergh Picture Collection. Manuscripts and Archives, Yale University Library<br />

Interlude: A glass heart<br />

90<br />

Lindbergh’s quiet work on heart research only became public in<br />

July 1935, when his second Science paper caught the attention<br />

of the New York Times. By that time he had been collaborating<br />

for several years with co-author Alexis Carrel, a Nobel prize-winning<br />

surgeon and pioneer in the science of tissue cultures.<br />

Carrel, who was working at the Rockefeller Institute for Medical<br />

Research in New York, had a vision of a day when organs could<br />

be removed from the body and kept alive, possibly even transplanted<br />

from one person to another. This could only be accomplished<br />

if tissues removed from the body could be supplied<br />

with blood, which would require some kind of machine and an<br />

engineer to build it, and there Carrel was stumped. His colleagues<br />

at Rockefeller built device after device. Every machine<br />

capable of feeding blood to organs also pumped them full of<br />

bacteria, leading to massive infections and death.<br />

Charles Lindbergh’s interest in the heart began with a personal<br />

tragedy. In 1929, the year of his marriage, his wife’s sister was<br />

diagnosed with rheumatic heart disease. The condition soon<br />

turned fatal because doctors could not yet operate on an open<br />

heart. What was lacking was a machine that could temporarily<br />

take over the heart’s job and pump arterial blood to tissues. It<br />

was, Lindbergh reflected, essentially an engineering problem.<br />

While he knew nothing about medicine, and had never finished<br />

his academic degree in mechanical engineering from the<br />

University of Wisconsin, he could repair his own airplane; he<br />

knew about pumps and valves and vacuum seals. As he began<br />

taking notes and drawing plans of devices, he needed to talk to<br />

physicians. This soon brought him into contact with Carrel, and<br />

one of the most unlikely collaborations in scientific history was<br />

born.<br />

Originally Lindbergh planned to build some sort of heart-lung<br />

bypass machine. Carrel quickly convinced him that the idea was<br />

too ambitious. Instead Lindbergh should focus on a way to keep<br />

tissues alive outside the body, which would permit researchers<br />

to carry out a more thorough basic study of the biology of tissues<br />

and organs. His first invention was a culture flask in which<br />

liquid medium could circulate continuously. He moved on to<br />

methods that could separate blood serum from plasma, and his<br />

final invention was a system of glass vessels and pumps that<br />

could supply whole organs with blood. The device permitted<br />

the handling, examination, and transfer of organs without danger<br />

of contamination. In one experiment Carrel kept epithelial<br />

tissue alive for over 100 days. When the two men finally gave a<br />

demonstration in a Copenhagen theater in 1936, their work<br />

received enthusiastic reviews from scientists and a wild<br />

response from crowds outside. The public misunderstood the<br />

findings and began placing orders for “Lindbergh’s artificial


heart,” as a replacement when their own biological<br />

one wore out.<br />

The tissue culture system became part of the<br />

standard repertoire of the Rockefeller<br />

Institute and other labs, but the Carrel-<br />

Lindbergh pump was difficult to operate and<br />

fell out of use within a few years. It was no<br />

longer necessary because most of the work<br />

biologists wanted to do didn’t require whole<br />

organs. They could work equally well with<br />

cells or thin tissue slices which could be kept<br />

alive through much simpler means. With<br />

Carrel’s death and the beginning of World War<br />

II, the project came to an end. But scientists of<br />

Lindbergh’s time regarded his contributions<br />

as important. His work on the heart, some<br />

said, would be remembered far longer than<br />

33-hour ride in an airplane. Crossing the<br />

Atlantic, after all, was little more than a symbolic<br />

gesture.<br />

��<br />

As long as we are healthy, it is easy to<br />

think of the heart as a solid, engine-like<br />

pump that will keep beating as long as it is<br />

exercised and provided with fuel. But those<br />

who are born with a heart defect know that<br />

this is an illusion, and the rest of us start to<br />

learn the lesson with middle age. The heart is<br />

more delicate than any of us would like to<br />

believe.<br />

91 Interlude: A glass heart<br />

One person who knows this very well is<br />

Ludwig Thierfelder, Acting Chief of Cardiology<br />

at the Franz Volhard Clinic of Charité/HELIOS<br />

and head of a research lab at the <strong>MDC</strong>.<br />

Ludwig is tall and youthful and to talk to him<br />

you wouldn’t think he has a stressful job. To<br />

get to his office you follow the road past the<br />

Robert Rössle Clinic, walk past the oldest laboratory<br />

building on campus, where the busts<br />

of Oskar and Cécile Vogt keep vigil. (Cécile is<br />

the taller one.) Across the Lindenberger Weg<br />

is what seems like a different world: the new<br />

HELIOS complex. The buildings are spacious<br />

and quiet. Color-coded signs act as guides to<br />

the Cardiology section. As you pass, patients<br />

are being transported through the wide halls<br />

on beds and wheeled into elevators. There are<br />

open waiting rooms with large round reception<br />

desks that would look equally at home in<br />

one of Berlin’s ultramodern hotels.<br />

Ludwig’s relationship with the heart began<br />

during his medical studies in 1988 at the<br />

University Clinic in Freiburg, where he treated<br />

cardiology patients while carrying out clinical<br />

research. “I had already done some work on<br />

blood pressure in rats during a six-month fellowship<br />

in the United States,” he says. “But it<br />

was in Freiburg that I really started to get<br />

interested in the genetics of heart muscle diseases.<br />

The first project I participated in was a<br />

collaboration with Hans-Peter Vosberg, of the<br />

Max Planck Institute for Medical Research in<br />

Bad Nauheim. Our aim was to study a large<br />

family with inherited hypertrophic cardiomyopathy.<br />

Cardiomyopathy is another way of<br />

saying ‘heart muscle disease,’ and hypertrophy<br />

means that the heart becomes dangerously<br />

enlarged. In general, these diseases are<br />

often connected to heart failure and a risk of<br />

sudden death. This particular problem was a<br />

rare condition caused by a dominant gene<br />

that nobody had found yet.”<br />

I have been wondering about the attraction of<br />

such rare diseases – wouldn’t it be better to<br />

spend your time working on problems that


Brenda Gerull<br />

affect huge numbers of patients? Especially<br />

in heart disease, which claims so many<br />

victims?<br />

“In this case it’s a bit deceptive to think of the<br />

problem as rare,” Ludwig says thoughtfully.<br />

“Congestive heart failure caused by various<br />

types of hypertrophy is a major cause of<br />

death. What we’re finding is that there are<br />

many genetic ‘roads’ that lead to these conditions.<br />

In most cases, each family has its own<br />

unique mutation. Why so many different<br />

defects lead to similar problems is an interesting<br />

question in itself.<br />

“Common cardiovascular diseases probably<br />

arise because of the influence of several<br />

genes and environmental factors. Most of the<br />

time they probably don’t involve what we<br />

would consider mutations. In the population<br />

you can find many versions of each gene<br />

involved in building the heart and operating it<br />

throughout a person’s lifetime. Some of these<br />

variants may make it more likely for you to<br />

develop a disease, if you happen to inherit it<br />

with a certain combination of other genes.<br />

These cases are extremely complex, and figuring<br />

out the contributions of each molecule<br />

and the environment is very difficult. That<br />

may not be easy to do, either, in a disease<br />

caused by a single mutation. But it’s certainly<br />

much easier.<br />

“Most such single-gene diseases are rare,<br />

that’s just the way evolution has worked. But<br />

the genes that cause them are often molecules<br />

whose ‘healthy’ forms may contribute in<br />

a weaker way to these more common diseases.<br />

So by looking at rare conditions we discover<br />

genes that will help us understand common<br />

ones. And it also tells us a lot about basic<br />

processes at work in the heart – which helps<br />

us understand the difference between<br />

healthy functions and unhealthy ones.”<br />

At the time of the Freiburg study, Ludwig says,<br />

only one gene had been linked to cardiomyopathies,<br />

a gene that encoded a muscle protein<br />

called the β-myosin heavy chain. Ludwig<br />

worked intensively with local physicians to<br />

find families with the condition in Freiburg. By<br />

the end he had recruited a family with over a<br />

hundred members and completed their pedigrees<br />

– filling out a chart showing the family<br />

relationships of affected and non-affected<br />

people, over several generations.<br />

Interlude: A glass heart<br />

92<br />

“I still have those somewhere,” he says, and<br />

goes to his bookshelf. He pulls down a binder<br />

and spreads one of the pedigrees out on the<br />

desk. Today pedigrees can be done on the laptop,<br />

with programs that instantly turn everything<br />

into neat colorful charts, but this one has<br />

been painstakingly filled out by hand. He walks<br />

me through one of the families. “Since the disease<br />

involves a dominant gene, if one parent<br />

has one copy of the gene, then half their children<br />

will inherit it, statistically speaking.”<br />

Comparing the DNA of affected and nonaffected<br />

family members might reveal the<br />

location of the gene. “On the basis of this data<br />

and our samples, we proved that the disease<br />

wasn’t caused by β-myosin,” he says. “That<br />

was exciting because it meant another gene<br />

was involved.”<br />

He received a postdoctoral fellowship from<br />

the German Research Council (DFG) to follow<br />

up on the case – in the United States. He<br />

packed up his family and moved to Boston,<br />

joining the group of Christine Seidman, a<br />

geneticist at the Harvard Medical School.<br />

“We had a great time in the US, both professionally<br />

and personally,” he says. “The other


people in the group were very nice; we’ve<br />

maintained contact and are still friends, still<br />

see each other every few years. One of our<br />

children was born there, so we have an<br />

American citizen in the family.<br />

“In Boston the project that had started in<br />

Freiburg moved along very successfully,” he<br />

says. “First I pinned the families’ problem to<br />

chromosome 15, and then specifically to a<br />

mutation in the α-tropomyosin gene. This<br />

molecule plays a key role in muscle contraction.”<br />

(How scientists track down such genes<br />

is described in “The case of the short-fingered<br />

muskateer.”)<br />

The results linked hypertrophic cardiomyopathy<br />

to a problem with small structures called<br />

sarcomeres, the key movers in muscle contraction.<br />

“That meant we ought to be looking for<br />

other defects in sarcomeres as potential causes<br />

for heart muscle diseases.”<br />

��<br />

Most proteins carry out their jobs in cells<br />

as parts of “molecular machines.” It’s a<br />

metaphor based on the fact that the shapes<br />

and chemistry of proteins let them snap onto<br />

each other to carry out jobs like moving other<br />

molecules, cutting them, tying them up, or<br />

transmitting signals. Sometimes these<br />

assemblies exist for just a split-second before<br />

being dismantled; then the pieces can be<br />

used for other things.<br />

Sarcomeres more closely resemble what we<br />

normally think of machines. They are built of<br />

many parts that last for a long time. They<br />

carry out a mechanical job: they are shaped<br />

like miniature pistons and that’s how they<br />

work as muscles expand and contract. The<br />

whole structure is shaped like a tube. At the<br />

center is a thick fiber made of the myosin protein,<br />

surrounded by sheath-like actin fibers<br />

that slide back and forth to lengthen and<br />

shorten the structure. The pieces are held<br />

together by the long, spring-like molecule<br />

titin. Thousands of sarcomeres, lined up endto-end,<br />

give muscle its striped appearance.<br />

Over the past 20 years scientists have worked<br />

out many of the chemical and physical details<br />

that allow these machines to work. Sticking<br />

out of the myosin fiber are small knob-like<br />

“heads”. When they dock onto the nearby<br />

actin sheath, the shape of the head changes.<br />

That pushes the actin filament a short distance<br />

and it frees up part of the head to grab<br />

small energy molecules called ATP. When that<br />

happens, the head lets go of the filament and<br />

snaps back to its original shape and position.<br />

It docks onto a new spot further down the<br />

actin filament and the process starts over<br />

again.<br />

For this system to work, all the molecules<br />

have to have the right shapes and chemistry.<br />

If one of the proteins has a mutation, it may<br />

function slightly differently, which can make<br />

the sarcomere work less efficiently. Magnified<br />

billions of times, over the length of a muscle<br />

fiber, the effect is amplified. That easily causes<br />

problems for organs such as the heart.<br />

α-tropomyosin, the gene in which Ludwig had<br />

found mutations, lies alongside the myosin<br />

fiber and fine-tunes the contact between<br />

myosin heads and actin. Mutations affect its<br />

performance and eventually result in heart<br />

disease.<br />

If one defect in sarcomeres could cause disease,<br />

so could others. Ludwig and his Harvard<br />

colleagues soon discovered a new defect in<br />

sarcomeres that caused hypertrophic cardiomyopathy.<br />

A project headed by Hugh<br />

Watkins, another postdoc in Christine<br />

Seidman’s lab, turned up mutations in a protein<br />

called MyBP-C. The abbreviation stands<br />

for myosin binding protein C, and that’s what<br />

the molecule does: it forms a bridge between<br />

the myosin filament and other parts of the<br />

sarcomere.<br />

��<br />

After Harvard, what next? Ludwig wanted<br />

to return to Germany, where he could<br />

practice medicine. From Boston he began<br />

looking for jobs, occasionally flying back for<br />

93 Interlude: A glass heart<br />

interviews. One stop was Berlin-Buch, where<br />

he visited the clinics of the Charité and the<br />

new laboratories of the <strong>MDC</strong>. Its scientific<br />

director, Detlev Ganten, was encouraging. The<br />

institute needed young experts at home in<br />

worlds of clinical and basic science.<br />

“There were very few places in the world<br />

where I could pursue both parts of my career,”<br />

Ludwig says. “In Buch, that would be strongly<br />

supported.” He toured the Franz-Volhard<br />

Clinic tucked into a stand of trees on the spacious<br />

grounds of Area 1, not suspecting that in<br />

eight years he would be walk through the<br />

doors as Chief of Cardiology<br />

When his family moved to Berlin in 1994, he<br />

had few illusions about what faced him. At<br />

least two years of intensive medical practice<br />

lay ahead before he completed his specialty<br />

training. On top of that, he had to set up a<br />

new research group, normally a full-time job<br />

in itself. At Harvard he had begun working<br />

with animal models of heart disease, particularly<br />

rats. He wanted to continue that, which<br />

meant establishing new lines of animals at<br />

the institute. He needed to hire clinicians who<br />

could learn molecular genetics. He would<br />

need to establish relationships with other<br />

physicians to recruit families with various<br />

types of heart muscle diseases. And all of this<br />

needed to lead to scientific results within just<br />

a few years, if he hoped to turn his initial fiveyear<br />

contract as a junior group leader at the<br />

<strong>MDC</strong> into a long-term career.<br />

Over the next few years the group studied a<br />

variety of inherited heart muscle diseases,


making pedigrees and taking samples from<br />

large families in Germany, Canada, Australia<br />

and elsewhere. Their investigation of a heart<br />

muscle disease called dilated cardiomyopathy<br />

revealed yet another mutation in a sarcomere<br />

component.<br />

Titin, the largest protein made by human cells,<br />

is a long spring-like molecule. It is lodged in<br />

the ends of the sarcomere and stretches<br />

toward the middle, linking all the pieces. As<br />

the myosin and actin fibers crawl along each<br />

other, the entire sarcomere is stretched. Titin<br />

helps keep things anchored. But it doesn’t<br />

only do structural jobs. The healthy form of<br />

the protein consists of hundreds of modules<br />

and many different types of domains. Some of<br />

them help pass signals; others are important<br />

in muscle development, probably because the<br />

protein acts as a scaffold to attach things to<br />

as sarcomeres are assembled from individual<br />

proteins.<br />

Brenda Gerull, an MD who has been working<br />

on cardiomyopathies in Ludwig’s group, found<br />

the mutation while studying a large<br />

Australian family with an inherited form of<br />

the disease. “They have just two additional<br />

letters in the DNA code for this huge molecule,”<br />

she says. “That has big effects, because<br />

it scrambles some of the information needed<br />

to make titin protein. The cell ends up making<br />

titin proteins that are too short. You wouldn’t<br />

expect a machine to work right if you shortened<br />

its springs, so it’s no surprise that<br />

changes in this molecule disturb the sarcomere<br />

and the heart.”<br />

Ludwig is frank about the meaning of these<br />

discoveries. “Our original focus was to identify<br />

genes,” he says. “Knowledge may be very helpful<br />

– particularly in diagnosis. If you have a<br />

gene that causes a problem in adulthood and<br />

you find it in a young person, you may be able<br />

to take precautions that will prevent the prob-<br />

Interlude: A glass heart<br />

94<br />

lem from becoming really serious. But we’d<br />

like to do more, and so increasingly the lab is<br />

turning to identifying and studying the<br />

processes by which defective genes cause<br />

problems for the heart and other organs.”<br />

Yet the power of knowledge and diagnosis<br />

shouldn’t be underestimated, as the lab<br />

was about to find out from an unexpected<br />

source.<br />

��<br />

The first chapter of this book describes<br />

Walter Birchmeier’s work on proteins<br />

that govern cell migration. Cells in most tissues<br />

are firmly tied to each other, but when<br />

they migrate the links have to be dissolved.<br />

This happens all the time as embryos develop,<br />

and it happens as cancer cells free themselves<br />

from a tumor. To work out how cells control<br />

these processes, Walter and his colleagues<br />

have been systematically removing genes in


mice and observing their effects on development<br />

and other processes.<br />

One of the molecules they are interested in is<br />

plakophilin 2, which appears in tissues such as<br />

growing skin, tumors, and heart cells. It is part<br />

of the desmosome, a sophisticated set of set<br />

of proteins which hook neighboring cells to<br />

each other. Fibers that tie cells together have<br />

to be anchored somewhere, and that’s what<br />

plakophilin 2 does. Just under the membrane<br />

at the fringes of the cell, it acts as a peg.<br />

“Mutations in many desmosome proteins<br />

have been linked to disease,” Walter says. “But<br />

until until we knocked out the gene in the<br />

mouse, no diseases had been linked to<br />

plakophilin 2, and we didn’t have much of an<br />

idea of its functions in development.”<br />

Katja Grossmann, a PhD student in the group,<br />

discovered that mice with no plakophilin 2<br />

died mid-way through embryonic develop-<br />

ment. The problem lay with the heart. “Under<br />

the microscope you could see that the fibers<br />

that were supposed to join cells were lying<br />

loose and scattered around,” Walter says. “As if<br />

someone had forgotten to secure the lines<br />

when they tied two boats up to each other.”<br />

Connections between heart muscle cells are<br />

crucial, he says, to ensure that they can communicate<br />

with each other and pass signals<br />

telling them when to contract.<br />

Walter says that many of the key proteins that<br />

establish these connections were originally<br />

uncovered by Werner Franke, a researcher in<br />

Heidelberg at the German Cancer Research<br />

Center (DKFZ) of the Helmholtz Association.<br />

“Over the past 30 years, Werner has discovered<br />

many of the proteins that play a crucial<br />

role in connections between cells,” Walter<br />

says. “And he has carried out a number of elegant<br />

studies that show how the pieces fit<br />

together.”<br />

95 Interlude: A glass heart<br />

One of those studies was published in 2002.<br />

As Walter’s group made knockouts of some of<br />

the pieces of desmosomes, one-by-one in the<br />

mouse, Franke and his colleagues were putting<br />

the pieces together. They were interested<br />

in the fact that cells in different tissues used<br />

many of the same components to attach<br />

themselves to their neighbors, but there were<br />

subtle differences. One way to understand the<br />

details would be to extract whole desmosomes<br />

and other types of junction complexes<br />

from cells, but that approach had failed<br />

because of the complexity of the structures<br />

and the fact that they were bound to membranes.<br />

Franke’s lab decided to investigate the question<br />

by adding proteins to cells piece-by-piece<br />

and watching what happened. They began<br />

with a laboratory cell line that produces one<br />

protein of the desmosome, called desmoglein<br />

2, but doesn’t produce most of the other com-


ponents or use them to build proper links to<br />

neighboring cells. The group began inserting<br />

various combinations of other desmosome<br />

molecules. One finding was that plakophilin 2<br />

plays a key role in organizing many of the<br />

other proteins, helping them get to their<br />

proper places. But properly building the structures<br />

required three proteins: desmoplakin,<br />

plakoglobin, and plakophilin 2.<br />

��<br />

Walter knew only too well that mutations<br />

in heart muscle proteins could<br />

cause disease. A few years earlier his group<br />

had knocked out plakoglobin, one of the<br />

essential pieces of the desmosome.<br />

“Removing it also caused fatal heart ruptures<br />

in the mouse, but we didn’t immediately follow<br />

up to see if there was a connection to dis-<br />

Interlude: A glass heart<br />

96<br />

ease,” he says ruefully. “The result was that we<br />

got scooped.”<br />

The lapse meant that William McKenna’s<br />

group from St. George’s Hospital Medical<br />

School in London was faster in linking mutations<br />

in human plakoglobin to Naxos disease.<br />

This is one of several diseases called arrhythmogenic<br />

right ventricular cardiomyopathies, or<br />

ARVCs. (The translation for the non-cardiologist<br />

is, “diseases of heart muscle in which<br />

there are defects in the right ventricle and the<br />

heart beats irregularly.”)<br />

Naxos takes its name from the Greek island,<br />

where a recessive copy of the gene is very<br />

common. People who inherit two copies of<br />

the mutation suffer from sudden, wild fluctuations<br />

of cardiac rhythm that often lead to<br />

heart failure and sudden death. ARVC is much


less common in other parts of the world, but<br />

still it is the leading cause of sudden death for<br />

cardiac reasons among young people.<br />

Vowing not to make the same mistake twice,<br />

Walter immediately took the plakophilin 2<br />

findings to the <strong>MDC</strong>’s experts on heart muscle<br />

diseases. Ludwig, Brenda Gerull and their<br />

colleagues identified ARVC patients, enlisted<br />

them in the study, and began checking for<br />

mutations. In 120 unrelated patients, they<br />

found 32 people with mutations in the<br />

plakophilin 2 gene. In all, Ludwig says, about<br />

30 percent of patients with the condition<br />

have shown mutations in plakophilin 2. Many<br />

of them may have a high risk for sudden cardiac<br />

death.<br />

“To give you an idea of how bad this condition<br />

can be,” Ludwig says, “in a study of ARVC<br />

which we completed in 2005, 50 percent of<br />

the men in the high-risk group were dead by<br />

the age of 39. Women at high risk survive<br />

much longer – the average age of death was<br />

71 years. We don’t yet know why there is a difference<br />

between the sexes, and we also don’t<br />

know what gives some people with the mutation<br />

better chances of survival than others.”<br />

In a collaboration that has been going on for<br />

many years with a group at Memorial<br />

University in Newfoundland, Canada, the scientists<br />

investigated an extended Canadian<br />

family with an inheritable form of ARVC. Here<br />

defects in another gene were responsible for<br />

ARVC – this time, on the third human chromosome.<br />

In early 2008 the scientists finally<br />

traced the problem to a mutation in a gene<br />

called TMEM43. They do not yet know what<br />

the molecule does, but it may have something<br />

to do with deposits of fat that build up in<br />

muscle fibers.<br />

“We knew we couldn’t cure the disease, but<br />

we could try to do something about the irregularities<br />

in heartbeat rhythm that caused the<br />

fatalities,” Ludwig says. “ARVC patients do not<br />

reliably respond to drugs that stabilize the<br />

heart, but they can be surgically implanted<br />

with a small defibrillator. This device detects<br />

An early sketch of Lindbergh’s pump<br />

when the heart begins to beat erratically and<br />

then administers an electric stimulation to<br />

bring it back to a normal pattern. It also<br />

records the events so we can later check to see<br />

how the heart – and the devices – have functioned.”<br />

The team carried out a clinical study in which<br />

the defibrillators – called ICDs – were implanted<br />

in 30 young, high-risk men and 18 women<br />

who had mutations in the region of DNA that<br />

held TMEM43. The results of the intervention<br />

were very clear. “In five years after the operation,<br />

70 percent of the devices had sprung<br />

into action at least once,” Ludwig says. “But<br />

there were no deaths among the patients.<br />

Compared to men who could not receive the<br />

defibrillator (because they died long before<br />

the connection between the heritable heart<br />

disease and TMEM43 was made), 28 percent<br />

died within five years. Within ten years they<br />

were all dead. So by identifying people with<br />

AVRC with a genetic test, and using the defib-<br />

97 Interlude: A glass heart<br />

rillator as therapy, we have been able to save<br />

several lives.”<br />

One of those cases was local, here in Berlin.<br />

The Franz Volhard Clinic also performs the<br />

operation, implanting the devices in high-risk<br />

patients. One person to receive the implant<br />

was a young local woman who carried the<br />

plakophilin 2 mutation. As she was outfitted<br />

with the defibrillator, her husband underwent<br />

first-aid training to learn what to do in the<br />

event that his wife suddenly experienced<br />

problems. That happened more quickly than<br />

anyone expected. Shortly afterward, she experienced<br />

a heart attack during the night.<br />

Thanks to the implant and her husband’s<br />

preparation, her life was saved.<br />

Ludwig expects to find more families with<br />

other mutations that cause AVRC. Linkage<br />

studies should help scientists find the genes –<br />

or at least regions of DNA – that are responsible.<br />

Then new types of tests can be made to


identify those who have the highest risk of<br />

sudden cardiac death. And then he knows<br />

what to do.<br />

You can tell it’s a very good feeling.<br />

��<br />

One Christmas when I was eight or nine<br />

years old, my parents gave me a<br />

Transparent Man, a human body that was 50<br />

centimeters tall, with see-through plastic<br />

skin, a skeleton, and organs that could be<br />

removed. I gave it a name, which I can’t<br />

remember. I remember being unhappy that<br />

the intricate network of veins and arteries<br />

were stamped into the plastic and painted<br />

blue and red rather than threading their way<br />

from the lungs to the heart and into the<br />

organs. I remember removing the cap of the<br />

skull and taking out the brain. And lifting the<br />

ribcage to expose the heart. Later my little sister<br />

hid one of the lungs and we never found it<br />

again.<br />

In high school I got the chance to watch surgery<br />

for real, and remember that the inside of<br />

the body was nothing like the tidy inner cavity<br />

of my toy. Little wonder that doctors and<br />

surgeons had to study so long – how could<br />

you learn to see structure in the chaos of<br />

blood and tissue under the skin?<br />

Ludwig was never given a Transparent Man,<br />

and the first time he saw surgery was during<br />

his medical studies. And that was an opportunity<br />

he almost never had.<br />

“I was a bad student in school,” he laughs. “I<br />

just didn’t get it; I had bad grades. I never<br />

would have been admitted to medical school<br />

in Germany.”<br />

Interlude: A glass heart<br />

98<br />

He was born and grew up in this country, but<br />

his father was Austrian and that gave him<br />

Austrian citizenship. It also entitled him to<br />

study there. So at the age of 20, he applied to<br />

the University of Vienna and was accepted.<br />

He loved the city and the medical school and<br />

its great teachers. Suddenly, he says, “I got it.”<br />

Medicine appealed to him very much. “I liked<br />

the mix of practical skills and intelligence. I<br />

seem to have been good at it,” he smiles.<br />

It’s a mighty understatement for someone<br />

who went from the back row of biology class<br />

to Harvard, and then to one of the most important<br />

medical positions anywhere: Head of<br />

Cardiology at one of Germany’s best hospitals.<br />

“There are a lot of kids in Germany today who<br />

won’t have the chance I had,” he says. “I’ve<br />

seen it with my own children. The current


educational system puts too much emphasis<br />

on great performance at too early an age. At<br />

the age of ten or eleven, children get pushed<br />

into tracks that are hard to escape from. Most<br />

of them won’t be able to go abroad, like I did.<br />

It’s a shame, because Germany is losing a lot<br />

of young people with great potential.” It’s a<br />

subject we could talk about for a long time,<br />

but there are patients waiting.<br />

As we shake hands I think that for our generation,<br />

this is the real molecular medicine. The<br />

miracle cures – the molecules that will slip<br />

into cells and replace defective genes, the<br />

immune system cells that will be taught to<br />

kill tumors – those are still somewhere on the<br />

horizon, or just beyond. For now what can certainly<br />

be done is to take the best of molecular<br />

biology and use genetics and other methods<br />

to probe the mechanisms that cause disease,<br />

then turn that knowledge into diagnostic<br />

tools. If the problem can be clearly identified,<br />

it can be attacked with the best of today’s<br />

medical techniques and a long line of new<br />

drugs. That’s not an answer that will satisfy<br />

everyone; it won’t let us live forever. The<br />

crowds are still outside, still placing orders for<br />

the artificial heart that Lindbergh dreamed of<br />

but never managed to build.<br />

From across the street I look back and realize<br />

that Ludwig and his colleagues are making<br />

the new transparent man, a better one. It will<br />

be different than the model I used to have.<br />

There will be far more parts. You’ll be able to<br />

zoom in as far as you like, zoom through the<br />

transparent skin and into transparent organs,<br />

transparent cells, transparent genes. You’ll be<br />

able to wind it up and watch the smallest<br />

parts dance and whirl.<br />

99 Interlude: A glass heart<br />

A lot of us are too old for such a toy. But that’s<br />

all right. We’ll give it to our children to play<br />

with. They’ll grow up with it, learn where all<br />

the parts go, and give it a funny name. And<br />

that will be their vision of a human being as<br />

they invent the medicine that Lindbergh<br />

dreamed of, that we dream of.


Part Two: Identity crisis


Martin Lipp<br />

In the year 1248, the city of Cairo constructed what<br />

was probably the largest hospital that has ever<br />

existed – a palace-like complex with high walls, open<br />

courtyards, and graceful towers. Al-Mansouri hospital<br />

held more than 8,000 beds, in specialized wards<br />

for surgery, fractures, fever, eye diseases, mental illness<br />

and other medical problems. Musicians and<br />

storytellers were hired to entertain the ill. The hospital<br />

accepted patients of all races, religions, and social<br />

classes and kept them until they either died or were<br />

cured. Patients had to prove they were ready to be<br />

discharged by eating a whole chicken. Those who<br />

passed were sent on their way with a new set of<br />

clothes and the equivalent of about ten Euros in<br />

pocket money. As well as a full belly.<br />

The first Chief of Physicians at Al-Mansouri was the<br />

great healer Ibn al-Nafis, author of an 80-volume<br />

medical encyclopedia that became the definitive<br />

103 Part Two: Identity crisis<br />

Where slow<br />

rivers meet


text for generations of physicians. Its ornately-scripted<br />

texts and gilded drawings contain<br />

history’s first correct description of the functions<br />

and structure of the human circulatory<br />

system, one of the accomplishments that<br />

secured al-Nafi’s place in history. (A copy of<br />

some of his manuscripts turned up in Berlin in<br />

1924.) But his teachings contradicted those of<br />

the great Roman physician and pundit Galen,<br />

who did not believe that blood recirculated, or<br />

that the heart pumped it. Because Western<br />

physicians still preferred the opinions of the<br />

ancient scholars over experiments and direct<br />

experience, it took more than 350 years for<br />

William Harvey and a group of Italians to rediscover<br />

circulation and introduce it in the West.<br />

It’s odd to imagine a culture ignorant of the<br />

role of blood vessels, which are as near as the<br />

next cut or scratch – and what could be more<br />

obvious than the pumping activity of the<br />

heart? But today’s society has its own blind<br />

spots when it comes to the body. People of the<br />

future might find it equally strange that we<br />

were almost totally ignorant of a second riverlike<br />

system that sends tributaries through all<br />

of our tissues and organs. It is rarely noticed<br />

until something goes wrong; then the lymphatic<br />

system manifests itself as swollen<br />

nodes, or tonsils, or much worse problems.<br />

William Harvey would have done well to<br />

know more about it; he and several other<br />

physicians were barely escaped execution<br />

when they failed to prevent the death of King<br />

James I, who had been under their care. One<br />

of the king’s ailments was arthritis. Many<br />

autoimmune diseases, including some forms<br />

of arthritis, are now being traced back to<br />

defects in the lymphatic system.<br />

Through this network of vessels and nodes<br />

flows a slow river, carried along by the contractions<br />

of muscles that line its tubes rather<br />

than a central pump. But it is by no means a<br />

uiet place. It carries white blood cells and<br />

other police of the immune system throughout<br />

the body, and serves as a training ground<br />

where cells are taught to fight pathogens –<br />

Part Two: Identity crisis<br />

104<br />

viruses, bacteria and other invaders. Pirate-like<br />

cancer cells that have detached from a tumor<br />

also use this trafficway. They take advantage<br />

of the route to spread to new tissues, causing<br />

deadly metastases and often taking up residence<br />

in the lymph system itself.<br />

��<br />

Martin Lipp of the <strong>MDC</strong> is a quiet-spoken<br />

man with blond hair that is becoming<br />

silver ("It's all blond!" he says) and enough<br />

patience to give me a beginner’s tour of the<br />

lymph system, which he has studied intensively<br />

for over two decades. We sit down at his computer,<br />

where he runs through a set of slides<br />

with neat, clear graphics. I ask who made them.<br />

“Me,” he says. “I like to do illustrations; it’s<br />

relaxing. A good change.”<br />

One slide shows a huge network of lymph<br />

vessels spread throughout the body. At regular<br />

intervals are small round nodes, like rest<br />

stops along a highway. “More like meeting<br />

points and training centers,” he says. “Those<br />

are two of the nodes’ functions.”<br />

Martin and his colleagues have found new<br />

kinds of cells that travel through lymph vessels,<br />

identified signals that guide their migrations,<br />

made important discoveries about how<br />

the cells work together, and helped untangle<br />

the intricate structure of lymph nodes. In the<br />

process he has learned a great deal about<br />

how the system arises in embryos and how it<br />

contributes to disease.<br />

Lymph consists of blood and other liquids that<br />

collect between cells. The fluids drain into tiny<br />

lymph capillaries that flow into larger and<br />

larger vessels and finally into a vein near the<br />

heart. This connects the lymphatic and circulatory<br />

systems. The routes are traveled by<br />

white blood cells and other immune system<br />

cells that help fight off foreign invaders, are<br />

responsible for rejecting transplanted organs,<br />

and sometimes turn against the body’s own<br />

tissues. It has taken decades to work out what<br />

these cells are, how they work with each<br />

other, and how they distinguish self from not


self. In other words, how the immune system<br />

is able to recognize what belongs to your<br />

body and what doesn’t.<br />

“That process might not be so complicated if<br />

all the cells in your body carried something<br />

like an identical I.D. card, but they don’t,”<br />

Martin says. “Each kind of human cell is<br />

unique and looks different to the immune<br />

system. Each foreign virus or bacteria is also<br />

unique. Yet the body usually still manages to<br />

tell the difference.”<br />

That task belongs to white blood cells called B<br />

cells and T cells. Both types are born in the<br />

bone marrow, but T cells leave at an immature<br />

stage. They move to the thymus, a lymph<br />

gland near the heart, to finish their development.<br />

B cells stay behind to mature. That<br />

involves producing antibodies – Y-shaped proteins<br />

that decorate their surfaces. Antibodies<br />

are created in an unusual way, by cutting and<br />

pasting genes together in random arrangements.<br />

The result is that the body can probably<br />

produce about ten billion different types<br />

of antibodies – as if a locksmith were to take<br />

billions of blank keys and cut them in different<br />

shapes. Chances are, one of them would<br />

fit and open the door of a particular house. In<br />

this case, an antibody needs to fit an antigen<br />

– a protein on the surface of a bacteria, virus,<br />

or another invader. If it succeeds, the cell that<br />

carries it can trigger a reaction from the<br />

immune system.<br />

This usually happens as a cooperative effort<br />

involving B and T cells and another type called<br />

105 Part Two: Identity crisis<br />

dendritic cells. This latter kind of cell is stationed<br />

in the skin and other tissues that constantly<br />

come in contact with the environment<br />

– the lungs, stomach, and intestines – where it<br />

is usually first to come in contact with foreign<br />

organisms and substances. It chews them up<br />

and glues fragments of their molecules to its<br />

surface, bound to molecules called MHCs,<br />

which stands for the major histocompatibility<br />

complex. Slight differences between the structures<br />

of these molecules in different people<br />

are important in helping the body distinguish<br />

between native and foreign cells. They are<br />

largely responsible for the immune system’s<br />

rejection of transplanted cells or organs.<br />

As they develop, T cells are outfitted with<br />

receptors that can recognize combinations of


MHCs and foreign molecules. As with antibodies,<br />

these receptors are made in a random way.<br />

If a T cell encounters a dendrite that one of its<br />

receptors can recognize, and it then meets a B<br />

cell with a matching antibody, the B cell<br />

begins to divide very quickly. This creates a<br />

huge number of identical daughter cells that<br />

produce and secrete the antibody. They glue<br />

themselves onto the surface of the invader,<br />

acting as a sort of alarm beacon. This summons<br />

macrophages and other types of white<br />

blood cells that digest the foreign object and<br />

break it down. (Anyone who has seen the<br />

movie Fantastic Voyage will remember<br />

macrophages as the huge cells that swallow a<br />

miniature submarine.)<br />

“Why don’t B and T cells attack the body?” I<br />

ask Martin. “What keeps B cells from making<br />

‘keys’ that fit proteins on the surfaces of their<br />

neighbors?”<br />

That happens all the time, he says; billions of<br />

the receptors and antibodies are made that<br />

would behave that way, if they were released<br />

into the body. But B cells that bear self-recognizing<br />

antibodies are usually destroyed before<br />

they leave the bone marrow.<br />

“The process for T cells is quite different,”<br />

Martin says. “During their development in the<br />

thymus, they have to pass two types of tests.<br />

The first is to make sure that they can bind to<br />

MHCs. Cells that can’t do this wouldn’t be<br />

much help to the immune system, so they are<br />

destroyed. Those that do survive undergo a<br />

second test, to see if they bind to combinations<br />

of MHCs and human proteins. If a T cell<br />

that did this got loose, it would probably start<br />

an autoimmune reaction. So in the second<br />

round, the cells that ‘succeed’ are trapped and<br />

destroyed. In the end only about two percent<br />

of T cells pass both tests.”<br />

The B and T cells that survive these rites of<br />

passage are now mature. They circulate<br />

through the bloodstream and travel to lymph<br />

nodes, meeting points that bring them into<br />

contact with dendritic cells. There the T cell<br />

may meet a foreign protein it recognizes,<br />

which makes it reproduce and develop. Some<br />

of the T cells leave the node to find and digest<br />

the invader. Others become the helper cells<br />

that activate B cells. B and T cells that aren’t<br />

activated exit the node and pass back into the<br />

lymph system. They will continue to cycle as<br />

long as they live, until they meet a cell they<br />

recognize. New dendrites continue to arrive in<br />

the lymph nodes. Like scouts sent ahead<br />

down slow rivers, they continually look for<br />

threats, then bring the news back home.<br />

��<br />

While working in Uganda in the 1950s,<br />

the Irish physician Denis Burkitt<br />

noticed a disturbing pattern of disease<br />

among children. One day a mother brought in<br />

a very ill child with swollen lymph nodes<br />

under the jaw. Burkitt’s diagnosis was cancer<br />

of the lymph system. It was a tragedy for the<br />

child, and the physician had an additional reason<br />

to be worried – he had seen the same<br />

thing before. He began digging through medical<br />

records and made the alarming discovery<br />

that there seemed to be an epidemic of such<br />

cancers. The victims were almost always children,<br />

and the tumors nearly always appeared<br />

in the same place, under the jaw. In 1958<br />

Burkitt published his findings in a medical<br />

journal, and the world became aware of the<br />

disease now known as Burkitt’s lymphoma.<br />

Three years later the British Medical Council<br />

gave him a grant – 250 British pounds, which<br />

today seems ridiculously small – to study the<br />

extent of the disease. Burkitt and two other<br />

physicians set off on an epic journey in which<br />

they traveled over 15,000 kilometers. At the<br />

end the picture was clear: the tumors<br />

appeared almost exclusively along the equator,<br />

in regions that were also threatened by<br />

malaria and yellow fever. Over time Burkitt<br />

figured out the connection. Other infections<br />

had weakened the children’s immune systems,<br />

making them easy targets for a very<br />

common type of herpes called the Epstein-<br />

Barr virus.<br />

Part Two: Identity crisis<br />

106<br />

Most people across the world are infected<br />

with this virus, which invades B cells. Usually<br />

it lives inside the cells without causing any<br />

harm. If it threatens to break out, other<br />

immune cells can handle it. But that can<br />

change quickly – especially if the immune system<br />

has been challenged by another disease.<br />

So this type of cancer does appear in other<br />

parts of the world that are constantly beset<br />

with epidemics. Fortunately, Burkitt and his colleagues<br />

were able to develop a cure, a type of<br />

chemotherapy using a drug called cyclophosphamide.<br />

It is so effective that a single treatment<br />

sometimes causes the tumors to vanish.<br />

In the 1970s Friedrich Luft (see the next chapter<br />

and “the Case of the Short-fingered<br />

muskateer”) heard Burkitt give a talk about<br />

one of his other interests: the connection<br />

between diet and cancer. Burkitt began a<br />

campaign to get people to eat more fiber. “He<br />

observed that colon cancer was inversely<br />

related to stool bulk,” Friedrich says. “This<br />

observation formed the ‘roughage’ hypothesis


to combat colon cancer. Needless to say, he<br />

showed some slides of things you normally<br />

don’t see in a scientific talk.”<br />

B cells are especially susceptible to cancer<br />

because of their ability to divide rapidly –<br />

always a risky business because the process<br />

may get out of control. In the early 1990s,<br />

when Martin was working at the Ludwig-<br />

Maximilian University in Munich, he decided<br />

to try to find out what made healthy B cells<br />

develop cancer. With his colleagues he compared<br />

the genes activated in healthy and diseased<br />

cells and discovered three candidate<br />

genes that were being used to create proteins<br />

at much higher levels in the tumors. “We had<br />

to pick one,” he said, “so we picked the most<br />

promising one.”<br />

The choice was a good one: later the team discovered<br />

that the protein’s structure closely<br />

resembled a common type of receptor called a<br />

G protein-coupled receptor (GPCR). These proteins<br />

are usually involved in passing crucial<br />

signals into cells. In the eyes, GPCRs transform<br />

light into signals that can be passed to the<br />

brain.<br />

“At the time only a handful of GPCRs had been<br />

cloned molecularly,” Martin says.“ Most were<br />

known as hormone and neuropeptide receptors,<br />

which transmit signals that trigger cardiac<br />

and metabolic functions. Other GPCRs<br />

were thought to regulate the movement of<br />

white blood cells from the blood into tissues<br />

under inflammation, but they had not yet<br />

been identified.” The protein that Martin had<br />

picked and provisionally named BLR1 (for<br />

Burkitt Lymphoma receptor 1) was the first<br />

GPCR identified in lymphocytes and might<br />

have some of these functions in B cells. To find<br />

out, Martin and his colleagues developed a<br />

strain of mice without the blr1 gene. He<br />

brought the animals along when he moved to<br />

the <strong>MDC</strong> in 1994.<br />

“We had knocked it out and discovered that<br />

animals without the molecule had several<br />

unusual features,” he says. “They were lacking<br />

some of their lymph nodes entirely. Those<br />

107 Part Two: Identity crisis<br />

they had were not organized correctly – cells<br />

didn’t migrate properly within them. This<br />

made it clear that the protein was playing a<br />

role in at least two processes: directing cells to<br />

their proper locations and helping the lymphoid<br />

tissues form.”<br />

The group wrote up a paper and sent it off to<br />

the prestigious journal Cell. The editors were<br />

hesitant to take the paper. “Everything was<br />

fine – they were convinced that we had found<br />

an important signaling molecule, guiding B<br />

cells into and through lymphoid organs,”<br />

Martin says. “But we didn’t know what external<br />

signal triggered it. Their problem was they<br />

had never published a paper where the receptor<br />

was known, but not its ligand. We wrote<br />

back, ‘The changes we see in the bodies of<br />

these mice won’t be any different once the<br />

ligand once is identified.’ And then they sent<br />

the paper out for review.”<br />

Later the protein would be called CXCR5, for<br />

chemokine receptor 5. “But it only got this<br />

name in 1998, when we knew what it was


inding to,” he says. Independently, two<br />

groups were able to identify its partner, a<br />

chemokine called CXCL13, which triggers<br />

migration of B cells through CXCR5.<br />

But understanding exactly what the<br />

chemokine receptors did would take several<br />

more years of work and the discovery of a second<br />

receptor protein called CCR7. “We wanted<br />

to see if T cells contained similar receptors, so<br />

we looked for a gene similar to CXCR5,” Martin<br />

says. “CCR7 was what we found.”<br />

When the scientists developed mice without<br />

this second receptor, they found further<br />

abnormalities in lymph nodes and the lymph<br />

system. Putting the two scenarios together<br />

gave them a new view of how cells move<br />

through the lymph nodes and how this<br />

affects the immune system. To explain, Martin<br />

needs to give me an overview of what normally<br />

goes on inside lymph nodes, and to do<br />

that he reaches to the top of one of his filing<br />

cabinets and pulls down a diagram that looks<br />

forbiddingly complicated. He promises to<br />

keep things as simple as possible.<br />

He reminds me that lymph nodes are<br />

the meeting points for dendritic cells that are<br />

usually the first to encounter foreign substances<br />

and organisms, and the T and B cells<br />

that actually mount the immune response. “To<br />

understand how these cells interact<br />

and their importance, we have to look at the<br />

structure of a node,” he says. “First, there is<br />

the border of the node itself. To cross over, all<br />

three kinds of cells have to have the<br />

CCR7 receptor. It’s attracted to a powerful<br />

signal, a chemokine called CCL21, which is<br />

produced by cells in the node. Without<br />

the receptor there’s no attraction and no entry.”<br />

Inside the perimeter is a region called the T<br />

zone. While it is mostly populated by T cells, all<br />

Part Two: Identity crisis<br />

108<br />

three types enter here. There is a second compartment<br />

inside, predictably called the B cell<br />

follicle because it is a breeding ground for B<br />

cells. A few T cells and dendrites also make it<br />

inside. The entry key for this compartment is<br />

the CXCR5 receptor.<br />

Most immune responses depend on all three<br />

kinds of cells, he reminds me, and they usually<br />

meet up in the T zone. “When B and T cells<br />

enter, they are usually ‘naive’ – they haven’t<br />

yet encountered a substance that matches<br />

their antibodies or receptors,” he says. “But<br />

some of the dendrites that come in bear fragments<br />

of foreign molecules. In that case, the<br />

first thing that may happen is for the dendrite<br />

to be recognized by a T cell. That triggers the<br />

cell to start making copies of itself and also to<br />

develop in various ways. Some of the T cells<br />

leave the lymph node and start hunting whatever<br />

it is that bears the antigen. If they find it,


they have various ways of destroying it. Other<br />

T cells, called T helpers, move to the edge of<br />

the B cell follicle. If they find a matching B cell<br />

– one with an antibody that fits what their<br />

own receptors have found on the dendrite –<br />

the B cell is activated.”<br />

It moves into the follicle, where it begins to<br />

reproduce. Here it gets assistance from another<br />

type of T cell – one which Martin discovered<br />

and christened as a B follicular helper T cell.<br />

“These cells are special because once they<br />

have been activated, they lower the amount<br />

of CCR7 they produce and build CXCR5 receptors,”<br />

he says. “This allows them to cross the<br />

compartment boundary. They have a quality<br />

control function. Sometimes the antibody on<br />

a B cell isn’t a very good match for the antigen.<br />

The T cell sorts those out and they are<br />

destroyed before they begin making defective<br />

antibodies.”<br />

Cells that pass the test reproduce very quickly.<br />

Now, instead of making unique, individualized<br />

antibodies, the cells turn into copying<br />

machines for molecules that can bind to the<br />

invader. They leave the follicle and the lymph<br />

node and secrete the antibodies. These attach<br />

themselves to antigens on the intruders,<br />

which are now attacked and destroyed by T<br />

cells and other white blood cells.<br />

The array of different types of cells is bewildering.<br />

Are there more to be found? I ask.<br />

“I’m sure there will be,” he says. “This is our<br />

current view of how things work. There are<br />

still questions. But our lab has filled in several<br />

of the key steps in what we know, and those<br />

findings have become widely accepted. It’s<br />

gratifying to see that some of the things you<br />

discover become incorporated into the textbooks.”<br />

��<br />

In the corner of Martin’s office is a palm tree<br />

that seems to have designs on the whole<br />

room – it has stretched leafy branches over his<br />

printer and half his desk. Next to his computer<br />

monitor is a small Buddha, the size of two<br />

clasped hands, sitting placidly in equilibrium<br />

atop an old mechanical postal scale. The statue,<br />

he explains, was a gift from two former lab<br />

members from India who married and left.<br />

The scale once belonged to Feodor Lynen, a<br />

Nobel prize-winner who headed the Institute<br />

of Biochemistry at the University of Munich. It<br />

was still standing there years after his departure,<br />

when Martin joined the institute in 1982,<br />

and it was about to be replaced by a digital<br />

scale. So it has accompanied Martin ever<br />

since.<br />

One urgent reason to study the development<br />

of the immune system and how its cells<br />

behave is to understand how – in spite of the<br />

body’s safeguards – they participate in<br />

autoimmune diseases. An increasing number<br />

of diseases are now recognized to have an<br />

autoimmune component. Martin also hopes<br />

that the work will help explain cancers like<br />

Burkitt’s lymphoma which affect the lymph<br />

system. Here, too, there is mounting evidence<br />

for a connection. More than a century ago<br />

Rudolf Virchow postulated a connection<br />

between chronic infections and cancer. The<br />

link may well be the immune system.<br />

Martin’s studies underline the connection<br />

between the behavior of cells, the development<br />

of structures in the body, and disease.<br />

Without key receptors such as CXCR5 and<br />

CCR7, mice did not develop proper lymph<br />

nodes. Without the complex architecture of<br />

the nodes, B and T cells might not develop and<br />

interact properly. If that didn’t happen, some<br />

of the cell types might be lacking, and crucial<br />

parts of the immune system might fail.<br />

These themes have come together, for example,<br />

in a study headed by Uta Höpken, a senior<br />

member of Martin’s group. She is now a<br />

Helmholtz fellow with her own PhD students.<br />

One of her interests is the behavior of T cells<br />

outside the lymph system.<br />

“We have learned quite a bit about the guidance<br />

of immune cells in the nodes and secondary<br />

lymph organs,” she says. “But these<br />

cells spend part of their lives outside, in the<br />

109 Part Two: Identity crisis<br />

bloodstream and surrounding tissues, before<br />

reentering the lymph system. That is an<br />

important part of their surveillance functions<br />

– it’s where they usually find foreign<br />

microbes. And this type of circulation is also<br />

known to help the T cells of newborns learn to<br />

distinguish self from non-self. Yet we don’t<br />

know much about what guides their migration<br />

and behavior outside the lymph organs.”<br />

Thinking that chemokine receptors might be<br />

involved, Uta and her colleagues looked carefully<br />

at the mice without CCR7. They discovered<br />

massive accumulations of T cells outside<br />

lymph vessels. In the stomach, intestinal tract<br />

and lungs the cells collected in follicle-like<br />

bulbs. “These are tissues that come in contact<br />

with substances from the environment – food<br />

and air – and so they play a special role in the<br />

immune system,” Uta says. “Normally the<br />

stomach and gut have a few follicle-like<br />

lymph structures, but not as many, and not as<br />

large. The animals still have some immune<br />

defenses, but the cells don’t migrate properly.”<br />

As the mice aged, they developed symptoms<br />

that looked very much like a human illness<br />

called Ménétrier’s disease. In this condition,<br />

masses of wrinkled tissue form on the walls<br />

of the stomach. Eventually the stomach<br />

grows much larger and becomes covered with<br />

large folds – until it almost looks like a brain.<br />

The disease is rare and usually affects middleaged<br />

men; it often leads to ulcers, and people<br />

who suffer from it have a higher risk of stomach<br />

cancer.<br />

“We think this means there is a strong connection<br />

between the formation of these<br />

unusual follicles and the way the lymph system<br />

normally develops,” Martin says. “Our<br />

main approach to investigate this has been to<br />

remove the receptors and block the signals.<br />

But we’ve also learned from studies of other<br />

groups that take the opposite approach –<br />

increasing the strength of signals.” For example,<br />

two laboratories in the US forced cells in<br />

unusual places to produce high amounts of<br />

CXCL13 or CCL19, the molecules that attract


A CD 3<br />

CXCR5 ond B cells and CCR7 on T cells, repectively.<br />

As a result, masses of the cells accumulated<br />

and formed follicle-like structures similar<br />

to those seen by Martin’s lab. But they<br />

were not active. The cells organized themselves,<br />

but the stimulus from the immune<br />

system was missing.<br />

These ectopic (“out-of-place”) follicles also<br />

resemble structures that are seen in several<br />

human autoimmune diseases, including<br />

rheumatoid arthritis. “About half the people<br />

who suffer from arthritis develop follicle-like<br />

lymphoid structures in unusual places,” Martin<br />

says. “They grow ectopically in the joints, and<br />

most likely trigger aberrant immune responses<br />

causing pain, bone destruction and finally<br />

immobility. If they really are similar to the follicles<br />

we have seen in mice, they might arise for<br />

some of the same reasons. The mice could give<br />

us a good system to investigate the mechanisms<br />

that cause the disease.”<br />

Previous mouse models of arthritis have not<br />

accurately reflected the chronic course of the<br />

disease, he says. A substance called mBSA<br />

causes chronic inflammations of the animals’<br />

joints, but it is artificial and the mechanisms<br />

by which it operates probably do not reflect<br />

what happens in the real disease. Another<br />

strategy has been to inject a molecule called<br />

collagen, derived from cattle, which provokes a<br />

systemic arthritis lasting for only three weeks.<br />

B B220 C PNAd<br />

D CCL21 E CXCL13 F BrdU<br />

But this method only works in a strain of<br />

mouse which is hard to investigate because of<br />

a lack of genetic tools.<br />

PhD student Antje Wengner started working<br />

on the theme using the mBSA method of creating<br />

arthritis in animals that lacked CXCR5<br />

and CCR7. She was hoping to peel apart the<br />

links between inflammations, immune<br />

responses, and arthritis. Antje followed the<br />

animals’ health for a very long time – over<br />

nine months. All the genetically normal mice<br />

treated with the substance develop ectopic<br />

follicles that are typical of arthritis, and it<br />

takes a long time – three to six months – for<br />

them to arise.<br />

“In rheumatoid arthritis patients, large numbers<br />

of T cells, B cells, and other immune system<br />

cells build up in the regions around the<br />

joints,” Martin says. “At different times each of<br />

these cell types has been accused of being<br />

responsible for the joint destruction seen in<br />

arthritis. B cells are particularly good suspects<br />

– therapies that eliminate them have had<br />

impressive successes.”<br />

Another sign of their involvement is that<br />

inflammed tissues in many patients produce<br />

high levels of the protein CXCL13. This is the<br />

target molecule for CXCR5 receptors on B cells,<br />

drawing them to B follicles and other places. If<br />

B cells collect around the joints, they can stim-<br />

Part Two: Identity crisis<br />

110<br />

These images show eptopic follicles (circular<br />

structure in the middle) obtained from a mouse<br />

with artificially induced arthritis. In the image<br />

on the top left, T cells are tagged in red. They<br />

gather in a different region of the follicle than<br />

B cells (tagged red in the image at the top, in<br />

the middle).<br />

ulate the formation of local lymph follicles<br />

with clear zones for T and B cells, which is a<br />

hallmark of a functioning lymph node-like<br />

structure. “Building a lymph structure on-site<br />

might allow it to cope with a prolonged, local<br />

inflammation. Instead, it develops into a continuous<br />

autoimmune problem.” Martin says.<br />

“We wondered what would happen if that<br />

structure didn’t form. Maybe it was a necessary<br />

step on the way to the prolonged inflammation<br />

and tissue damage that happens during<br />

rheumatoid arthritis.”<br />

Since the loss of receptors disrupts the structure<br />

of follicles, Antje wondered what would<br />

happen if she induced arthritis in mice that<br />

didn’t have CXCR5 and/or CCR7. Using both<br />

methods of stimulating arthritis in the mouse<br />

allowed her to watch the immune system’s<br />

response to inflammations in the early phases<br />

as well as later, chronic stages of the disease.<br />

The first phases were similar in mice with and<br />

without the receptors; the researchers discovered<br />

local inflammations and an influx of<br />

immune system cells. But mice lacking the<br />

receptors had fewer T cells that recognized<br />

the foreign substances. And over the long<br />

term, there was much less bone damage in<br />

animals lacking the CXCR5 receptor.<br />

Additionally, the lack of the receptors prevented<br />

the formation of functioning lymph follicles<br />

around the joints.


Martin’s lab is continuing the project by challenging<br />

the mice with multiple methods of<br />

inducing arthritis. “In the novel model, it only<br />

takes four to six weeks for the mice to develop<br />

the full-blown symptoms of the chronic<br />

disease, and 95 percent of them develop<br />

active ectopic follicles,” Martin says. “This will<br />

make a much more efficient model to study<br />

the progress of arthritis and to grasp the contributions<br />

of different receptors and cell<br />

types.”<br />

But the recent results are already quite promising.<br />

“This suggests that in many cases,<br />

CXCR5 may play a central role in some of the<br />

most chronic destructive effects of rheumatoid<br />

arthritis,” Martin says. “Finding a way to<br />

shut it down locally might be a very effective<br />

strategy to treat patients who suffer from this<br />

and other autoimmune diseases.”<br />

��<br />

There is a diagram of the lymph system in<br />

my daughter’s eighth-grade biology<br />

book. When we look at it together, she says it<br />

reminds her of the streams and rivers whose<br />

names she has been learning in geography<br />

class. Soon there will be a test and she will<br />

have to draw and label the Rhine, the Neckar,<br />

and dozens of other waterways on a blank<br />

map.<br />

We normally think of our bodies as well-built,<br />

finely tuned machines that have to be maintained<br />

through diet and exercise. We watch<br />

ourselves grow and age, but just as we seldom<br />

see rivers change their course, we aren’t<br />

usually very aware of the dynamic nature of<br />

our inner world. Martin and his colleagues see<br />

the lymphatic system as a evolving structure<br />

that grows and changes through its activity<br />

and challenges from the outside. Maybe it<br />

shouldn’t be any surprise that new lymph follicles<br />

spring up at the site of a chronic inflammation.<br />

After all, the same thing happens on<br />

a map, over long periods of time. Villages also<br />

arise at the places where rivers or roads run<br />

together, where the train stops, where the<br />

111 Part Two: Identity crisis<br />

mail is delivered. Battles are fought there, and<br />

the future is often decided at such places.


Lost anchors<br />

and wrecked vessels<br />

Iheard that Sibylle von Vietinghoff, a house-staff physician at the<br />

<strong>MDC</strong> and the Charité, was investigating a disease involving B cells<br />

and other white blood cells called neutrophils, and an autoimmune<br />

reaction in which the body builds antibodies against one of its own<br />

molecules. In this case, the target is a protein found in and on the neutrophils.<br />

She worked in the groups of Ralph Kettritz and Friederich Luft,<br />

clinicians who treat patients while running laboratories concerned with<br />

the mechanisms that underlie disease. It seemed like an ideal position<br />

for a resident in internal medicine to get an exposure to both worlds,<br />

good preparation for a day when medicine and basic research work<br />

much more closely together. But it’s a challenge to serve both masters.<br />

Budding clinical scientists like Sibylle have the double burden of spending<br />

time in the lab as well as working long, intensive shifts in hospital<br />

wards, so it was hard to find time to meet. In March Sibylle was finishing<br />

up a round of experiments while getting ready to start a six-month<br />

rotation in the intensive care unit.<br />

In spite of the stress, she found a spare hour on a beautiful day in early<br />

spring. I took a ten-minute bus ride through Berlin-Buch to reach her<br />

lab in the Franz Volhard Clinic, which specialized in hypertension and


Ralf Kettritz and Sibylle von Vietinghoff<br />

systemic vascular diseases. The clinic has now<br />

moved into the Helios Krankenhaus Berlin-<br />

Buch complex, directly across the road from<br />

the Berlin-Buch campus. But in the spring it<br />

still resided in House 134 of “Area 1,” one of five<br />

large hospital grounds scattered throughout<br />

the village. Most of the other clinics in Area 1<br />

had already moved, leaving behind an array of<br />

elegant, aging buildings, like a ghost town of<br />

the history of medicine.<br />

The buildings are separated by large, park-like<br />

spaces with gazebos and ornate fountains.<br />

Children of some of the staff were running<br />

around, climbing on things, peeking into<br />

buildings. Some of the houses had evidently<br />

been empty for a long time. Sunlight cast odd<br />

shadows through the broken windows, into<br />

old hospital wards, nurses’ stations, cafeterias.<br />

Other structures have been carefully maintained.<br />

An elderly couple was having lunch on<br />

a sun-washed balcony of one house. They<br />

reminded me of an anecdote from one of<br />

Heinz Bielka’s books on the history of Berlin-<br />

Buch: the town was home to the first senior<br />

citizens’ home in Germany that permitted<br />

husbands and wives to move in together.<br />

I shouldn’t have been surprised – but was – to<br />

come through the front door of the Franz<br />

Volhard Clinic and find myself in the waiting<br />

room for patients. A nurse at the front desk<br />

provided directions to the labs, which were in<br />

the basement. Another surprise was to find


Sibylle in the white shirt and slacks of a physician.<br />

There’s no dress code for molecular biologists,<br />

unless you count the white lab coat<br />

they pull on to do some types of experiments.<br />

Sibylle’s outfit – and the stethoscope around<br />

her neck – were more reminders of a culture<br />

gap. From time to time she looked at her<br />

watch. She apologized. After we talked she<br />

had to make rounds on a ward upstairs.<br />

One of the projects of Ralph Kettritz’s lab<br />

involves the way that white blood cells defeat<br />

pathogens, and how this process sometimes<br />

leads to an autoimmune disease. White blood<br />

cells, or leukocytes, squeeze through the lining<br />

of blood vessels to seek out viruses, bacteria,<br />

and infected cells that have been marked by<br />

antibodies. When a cell finds one, it destroys<br />

the pathogen by releasing a burst of oxygen<br />

atoms and toxic enzymes.<br />

Sibylle compares the effect to a molecular<br />

hand grenade. “The oxygen atoms are highly<br />

charged with energy and that overwhelms<br />

the invader. Its molecules become overloaded<br />

and start to carry out too many reactions;<br />

they spin out of control. This tears the<br />

pathogen apart.”<br />

The reactions have to happen at the right<br />

time and place to avoid injuring the body.<br />

There are nearly always healthy cells around<br />

when the bursts happen, so they have safety<br />

devices: they possess molecules that can<br />

break down the overloaded oxygen, or they<br />

enlist the help of antioxidants such as vitamin<br />

C. But in a rare disease called systemic vasculitis,<br />

the bursts are triggered too early, when<br />

the neutrophils are still in the vessel or in the<br />

vessel wall, and have not reached yet the site<br />

of inflammation. This premature reaction<br />

injures the blood vessels and eventually<br />

destroys the organs they supply with blood.<br />

“This is one thing that happens in autoimmune<br />

diseases,” Sibylle says. “If a leukocyte<br />

Part Two: Identity crisis<br />

114<br />

has been taught to target a healthy human<br />

cell, it will treat it as a pathogen and try to<br />

destroy it. That will obviously lead to inappropriate<br />

bursts and collateral damage.”<br />

One form of vasculitis is an auto-immune disease<br />

that occurs when the body builds antibodies<br />

(called ANCA) against proteins which<br />

appear on the leukocytes themselves. Sibylle<br />

and her colleagues in Ralph Kettritz’s lab<br />

traced the problem back to bad behavior on<br />

the part of two proteins.<br />

Ralph knew that ANCA-vasculitis patients<br />

develop antibodies against a protein called<br />

PR3, which is produced only in neutrophils.<br />

Most PR3 is kept out of reach, glued to internal<br />

compartments of these cells, but occasionally<br />

it appears on the outer surface. In<br />

ANCA-vasculitis, much more of it moves from<br />

the interior to the outer membrane. If the<br />

body builds antibodies against it, the result is<br />

a cycle of autoimmune reactions that eventu-


ally become deadly because of severe organ<br />

damage. Neutrophil activation by ANCA is<br />

central to this tissue injury.<br />

“The therapies we currently use to treat<br />

ANCA-vasculitis block the body’s ability to<br />

manufacture antibodies against PR3, but they<br />

also shut down the production of a wide<br />

range of other antibodies,” Ralph says. “That’s<br />

a very aggressive tactic, because it weakens a<br />

person’s defenses against other diseases. The<br />

thing to do was to try to figure out why PR3<br />

was migrating to the surface of the cells.”<br />

Samples of white blood cells were obtained<br />

from healthy people and then compared to<br />

samples from patients with ANCA-vasculitis.<br />

As expected, the patients had more PR3 on the<br />

surface of leukocytes. “Then we caught a lucky<br />

break,” she says.<br />

One of the patients on her clinical ward had<br />

an unusual blood disease called polycythemia<br />

rubra vera (PRV) – basically a form of<br />

cancer – in which certain types of precursor<br />

cells in the blood marrow divide too often.<br />

One sign of the disease is an overabundance<br />

of a protein called NB1 on the surface of blood<br />

cells.<br />

“At about the same time another patient<br />

turned up with an equally obscure disease<br />

called paroxysmal nocturnal hemoglobinuria,<br />

or PNH,” Friedrich Luft says. “These patients<br />

have red blood cells that burst in the night – a<br />

tell-tale sign is blood in their urine in the<br />

morning. Well, the white blood cells of these<br />

patients never have NB1 on their surfaces. So<br />

here was one patient with too much NB1,<br />

another with none. We thought that was<br />

interesting, so Ralph and I demanded that<br />

Sibylle learn everything ever known about<br />

NB1. And as things turn out, NB1 would end up<br />

being pivotal in a third disease, the one Sibylle<br />

wanted to study.<br />

115 Part Two: Identity crisis<br />

“This vignette is important because it reveals<br />

the ingredients that are required for clinical<br />

research to work,” Friedrich says. “You need to<br />

start with curious, industrious young clinicians<br />

who want to be scientists. Add to that<br />

mentors who have a successful (and funded)<br />

laboratory that can embrace young minds.<br />

And occasionally there is an old chief of service<br />

who has seen almost everything and<br />

made a lot of academic blunders, who can<br />

guide decisions and make things happen.<br />

Sibylle made a ‘triple play’ in this regard.”<br />

Healthy cells have to defend themselves from<br />

the bursts of white blood cells, and part of the<br />

protection comes from proteins on their surfaces<br />

that can deal with excess oxygen. Many<br />

of these proteins are attached to the membrane<br />

by anchors made of fat molecules<br />

called lipids. NB1 is one such protein. In PNH a<br />

lipid anchor called GPI is defective, particularly<br />

in red blood cells. It can no longer hold onto


NB1, and without an anchor the protein drifts<br />

away. When that happens the cell loses an<br />

important piece of protection against damage<br />

from leukocytes and it is chewed apart by<br />

the immune system.<br />

The patient with PNH had, as expected, no<br />

NB1 in the membranes of cells. Then the scientists<br />

thought to check for the presence of PR3,<br />

the protein involved in ANCA-vasculitis – and<br />

didn’t find any. Because the two proteins were<br />

both involved in autoimmune diseases, there<br />

might be a connection. Maybe they worked<br />

together in the cell.<br />

No one had checked NB1’s expression in<br />

ANCA-vasculitis, so Sibylle and her colleagues<br />

did so. “We discovered that NB1 was behaving<br />

just like PR3; in the disease, both molecules<br />

move to the membrane in higher amounts<br />

than normal. Particularly interesting was that<br />

the percentage of neutrophils that express<br />

both molecules on the outer cell membrane<br />

was higher than in controls. We marked the<br />

molecules with fluorescent tags and watched<br />

them under the microscope. You can’t see<br />

individual proteins under the microscope, but<br />

you can see the patterns that lots of them<br />

form, and they seemed to end up in the same<br />

positions on the surface of the cell.”<br />

Such close associations often mean that two<br />

molecules are bound to each other. Did the<br />

migration of PR3 somehow depend on NB1, or<br />

vice-versa? It was hard to tell. In diseases<br />

many things go wrong, and what looked like a<br />

connection might be a coincidence caused by<br />

something else. Interfering with NB1 might<br />

give some answers, so Ralph’s laboratory used<br />

an enzyme to cut NB1’s anchor in white blood<br />

cells. Without this protein, PR3 disappeared as<br />

well. Then they tried the opposite approach.<br />

They took a cell that didn’t produce either protein,<br />

and made it express NB1 and PR3, together<br />

and separately. PR3 did not appear on the<br />

membrane when it alone was made by cells.<br />

However, when PR3 was co-expressed with<br />

NB1, suddenly both proteins appeared on the<br />

surfaces of cells.<br />

“This was great evidence to support a direct<br />

connection between the proteins,” Ralph says.<br />

Part Two: Identity crisis<br />

116<br />

Microscope images revealed that NB1 (green) and<br />

PR3 (red) take up the same positions when they<br />

move to the surface of leukocytes. PR3 can only<br />

move there with the help of NB1. Normally PR3 is<br />

kept within the cell, but in ANCA-vasculitis patients<br />

the two molecules appear on the surface at a high<br />

rate. The body builds antibodies against PR3,<br />

causing a dangerous autoimmune reaction.<br />

“PR3 apparently needed NB1 to get to the<br />

membrane and stay there, because when we<br />

cut the anchor both proteins disappeared.<br />

This finding explained what we had seen with<br />

the PNH patients. They had lost their NB1, just<br />

as if the anchor had been cut, so their mambranes<br />

had no traces of either NB1 or PR3.”<br />

ANCA-vasculitis is an unusual disease, but a<br />

very serious one, and the connection between<br />

PR3, NB1 and its lipid anchor gives scientists<br />

new starting points to look for treatments. If<br />

PR3 is directly responsible for the dangerous<br />

symptoms of the disease, maybe scientists<br />

can find a way to cut NB1’s anchor, or break<br />

the connection between the two molecules.<br />

Most autoimmune diseases are hard to treat<br />

because it is hard to separate the healthy<br />

functions of the immune system from<br />

unhealthy ones. One strategy has been to use<br />

drugs that shut down parts of the immune<br />

system, but this leaves the body open to<br />

attacks from pathogens. Ralph and Sibylle<br />

know that finding the fundamental causes of<br />

a disease is only the first step toward a thera-


117 Part Two: Identity crisis<br />

py. But people who are equally comfortable in<br />

two cultures are always the best to solve<br />

problems of translation.<br />

And chance plays a role – if a scientist and her<br />

institute are prepared to take advantage of it.<br />

“Uncovering a crucial mechanism in this case<br />

stemmed from the fact that Sibylle cared for<br />

both the PRV patient and the PNH patient,”<br />

Friedrich says. “This clinical experience will<br />

surely guide her future work. Ironically,<br />

Sibylle’s findings are not likely to help patients<br />

with either PRV or PNH. Both patient groups<br />

can now look forward to acceptable and effective<br />

therapies. Instead, her work may benefit a<br />

completely different group, patients with PR3<br />

disease. The link to NB1 may break open PR3<br />

research.”


photo by Wayne McLean<br />

A pact with the Devil<br />

It’s not often that a cartoon character steps up to shed light on the<br />

immune system, cancer, and the potential of new therapies involving<br />

T cells. But the Tasmanian devil, best known as a star of children’s<br />

television, has suddenly caught the interest of scientists across the<br />

world. The real marsupial lives only on the island nation of Tasmania,<br />

where its odd shrieks startle tourists and keep people awake at night.<br />

Several hundred years ago the devil was also widespread on mainland<br />

Australia, but it became extinct there with the arrival of the dingo. Now<br />

it faces another threat. In 1996 scientists discovered an epidemic of cancer<br />

that was killing the animals in a northern region of the island. Since<br />

then the disease has spread to most of the island, so rapidly that<br />

researchers estimate it may have destroyed half the entire species.<br />

Tumors arise in the animal’s mouth and on its face, eventually making<br />

it impossible or too painful for the devil to eat, and it starves to death.<br />

What has been so puzzling about the disease is that it seems to spread<br />

directly from animal to animal. The devils often have fierce fights in<br />

which they bite each other on the face. This explains how cells might<br />

move between animals – but not why they cause a tumor in the animal<br />

Part Two: Identity crisis<br />

118<br />

Tasmanian Devil in defensive stance,<br />

at Tasmanian Devil Conservation Park,<br />

Tasman Peninsula.<br />

Use under the Creative Commons Attribution<br />

ShareAlike 2.5 License.


Thomas Blankenstein


that has bitten. Cancer cells aren’t like parasites,<br />

able to jump between hosts and directly<br />

cause a new disease.<br />

Thomas Blankenstein of the <strong>MDC</strong> and the<br />

Charité speaks of the devil during talks on<br />

cancer and the immune system, because the<br />

behavior of its cancer is an almost absolute<br />

exception. “Everything happens in nature,” he<br />

says. “But as far as we know this is only the<br />

second case ever of a directly transmissible<br />

cancer. Normally you could inject billions of<br />

tumor cells into animals – if they didn’t come<br />

from populations that had been heavily<br />

inbred in the lab – and those cells would be<br />

immediately rejected by the immune system,<br />

like any other foreign cell.”<br />

Several other types of cancer are thought of<br />

as “transmissible” for a different reason. They<br />

are linked to infections, the way the Epstein-<br />

Barr virus leads to Burkitt’s disease (described<br />

in the story “Where slow rivers meet”) or the<br />

way the human papilloma virus triggers<br />

tumors in the cervix. In those cases viruses<br />

bring along genes that disrupt the cell’s own<br />

molecules, causing it to spin out of control.<br />

The opposite scenario – that cancer cells<br />

themselves behaved like parasites, colonizing<br />

a new host and then reproducing – had only<br />

been seen once before, in a sexually transmitted<br />

disease affecting dogs.<br />

So the community was skeptical about the<br />

transmissible cancer hypothesis until 2006,<br />

when two Tasmanian scientists, Anne-Maree<br />

Pearse and Kate Swift, examined cells from<br />

animals with the devil-facial tumor disease.<br />

They discovered some extremely odd characteristics:<br />

cancer cells were missing five of the<br />

normal devil chromosomes and had four<br />

Part Two: Identity crisis<br />

120<br />

extra ones. It was extremely unlikely that such<br />

massive changes in the genome could spontaneously<br />

occur more than once, but tumor<br />

cells taken from animals all across the country<br />

had the same features. This virtually proved<br />

that all of the cases could be traced back to<br />

one source, rogue cells that evolved in a single<br />

animal. Like parasites, they were being passed<br />

from one devil to the next in a bite.<br />

��<br />

The immune system has evolved to fight<br />

bacteria, viruses, and other infectious<br />

agents. Likewise, animal immune systems<br />

normally recognize cells from other animals<br />

as foreign and reject them, whether those<br />

cells come from tumors or not. But devils<br />

don’t reject the tumors or many other types<br />

of cells transplanted from other members of<br />

their species. Usually this only happens


etween identical twins, and Pearse and Swift<br />

proposed that heavy inbreeding within the<br />

population had produced devils whose tissues<br />

were behaving like those of twins. A<br />

2007 study by other labs in Australia and<br />

Tasmania confirmed this and showed what<br />

aspect of the immune system was at fault.<br />

Most devils are so closely related that their<br />

cells produce very few types of MHC molecules.<br />

(These are the “vacuum sweeper” proteins<br />

on the surfaces of cells that combine<br />

with fragments of foreign molecules,<br />

described in “Traps along slow rivers.”)<br />

Receptors on T cells recognize combinations<br />

of MHCs and these fragments.<br />

Most species have three types of MHC 1 proteins,<br />

each of which picks up a different kind<br />

of fragment. And mutations and evolution<br />

have produced an enormous number of “flavors”<br />

of each type. Normally this variety plays<br />

an important role in the robustness of a<br />

species’ immune system. Since inbreeding has<br />

impoverished that system in devils, they can’t<br />

protect themselves from invasions by foreign<br />

cells.<br />

A normal healthy immune system is so critical<br />

that it sometimes even targets “self” and produces<br />

autoimmune diseases. “This raises the<br />

question,” Thomas Blankenstein says, “of<br />

whether the immune system recognizes cancers<br />

which arise spontaneously. The idea that<br />

Anna Kruschinski<br />

it should was a very attractive hypothesis<br />

when it was first proposed. A problem with<br />

very attractive hypotheses is that sometimes<br />

they hang on even when there is evidence to<br />

the contrary.” Or as another scientist named<br />

Thomas – Thomas H. Huxley – once put it,<br />

“The great tragedy of science – the slaying of<br />

a beautiful theory by an ugly fact.” He meant<br />

it ironically.<br />

Thomas Blankenstein is a lanky man with<br />

short-cropped black hair and glasses. If I didn’t<br />

know what he did, I could imagine him roaming<br />

gullies and desert ridges, digging for<br />

dinosaur bones. Instead he spends his days in<br />

the laboratory digging into the immune<br />

system.<br />

The themes he is working on stretch back over<br />

a century, to Paul Ehrlich’s realization that<br />

human cells likely underwent mutations all<br />

the time. Why didn’t more of them lead to<br />

cancer? He reasoned that the body had to<br />

have a system for dealing with them. When T<br />

cells were discovered in the 1950s, their func-<br />

121 Part Two: Identity crisis<br />

tions were unknown, but their behavior suggested<br />

that they were involved in immunity.<br />

Frank McFarlane Burnet and Lewis Thomas<br />

proposed that the role of T cells was to block<br />

tumors. The idea caught on and persisted as<br />

researchers unraveled the many functions of<br />

T cells in fighting infectious diseases.<br />

Most tumors begin when a human gene<br />

undergoes mutations. That happens more<br />

often the older a person gets, and it usually<br />

happens long after the immune system has<br />

learned the difference between self and notself.<br />

It is logical to suppose that the mutated<br />

form of a gene (and the new protein it<br />

encodes) would look foreign to the body’s surveillance<br />

system and thus trigger an immune<br />

response the way viruses or bacteria do. This<br />

was the standpoint of the “attractive hypothesis.”<br />

Since the body doesn’t reject all tumors,<br />

however, the hypothesis needed a second<br />

part: a mechanism that allowed cancer to<br />

escape immune control. A lot of laboratories<br />

began working hard to find it.<br />

The work of Thomas and many others suggests<br />

that the body treats tumors caused by<br />

viruses as infections; it handles those that<br />

arise spontaneously, usually late in life, differently.<br />

“Work by Georg and Eva Klein carried<br />

out 30 years ago clearly demonstrated that<br />

the immune system can cope with some<br />

kinds of tumors,” he says. “For example, T cells<br />

can very efficiently control B cell lymphomas<br />

that arise in association with the Epstein Barr<br />

virus. They also control other tumors associated<br />

with infections. This hinted that it should<br />

be possible to vaccinate animals against


tumors if a clearly defined target – an antigen<br />

specific to the tumor – could be found. And<br />

this is the case.”<br />

These findings make sense in the light of evolution,<br />

he says. The immune system is the<br />

product of natural selection, which helps<br />

shape organisms up to the point that they are<br />

finished bearing offspring. The immune system<br />

has evolved to fight infectious diseases –<br />

including those related to cancer – because<br />

they attack young mammals. But there is no<br />

pressure to protect organisms from old-age<br />

diseases, like most types of cancer that arise<br />

from spontaneous mutations. So these are<br />

probably fundamentally different kinds of diseases,<br />

and it makes sense that they should<br />

evoke different reactions from the body. That<br />

was what the Kleins’ experiments showed,<br />

and it might have put the Burnet hypothesis<br />

to rest for good.<br />

In 2006 Georg Klein wrote to Thomas and<br />

said he would like to visit the <strong>MDC</strong> as part of<br />

a trip he was making to Berlin. It was a rare<br />

opportunity; Klein is in his eighties and no<br />

longer travels very much. “We quickly organized<br />

a seminar,” Thomas said. “It was such<br />

short notice that the only time we could find<br />

was early on a Saturday morning. You can<br />

imagine how eager people are to come back<br />

to the campus at such a time, after a long<br />

work week. But the room was completely full.”<br />

The Kleins’ hypothesis received further support<br />

in the mid-1970s with the discovery of<br />

the nude mouse. This strain of animal had<br />

arisen through a chance mutation. As well as<br />

having no hair, the animal had no thymus –<br />

thus no T cells, and an immune system that<br />

couldn’t cope with infections. The animals<br />

must be kept in a carefully controlled environment<br />

or they quickly die.<br />

“Here was another chance to pit the Burnet-<br />

Lewis hypothesis against that of the Kleins,”<br />

Thomas says. “Nude mice undergo just as<br />

many cancer-causing mutations as other<br />

strains. Now suppose that the adaptive<br />

immune system played any role in suppress-<br />

Part Two: Identity crisis<br />

122<br />

ing such tumors. You’d expect, obviously, that<br />

nude mice would experience far more cancer<br />

than other strains. But that doesn’t happen.<br />

The conclusion that we draw is that T cells<br />

don’t fight spontaneous tumors.”<br />

A higher rate of tumors has sometimes been<br />

found in animals that have been immunesuppressed<br />

their entire lives. But this may well<br />

be due, Thomas says, to unknown “latent”<br />

infections or other environmental factors that<br />

play a role in the growth of tumors. “This<br />

wouldn’t be surprising,” he says, “because it<br />

supports a hypothesis of Rudolf Virchow, who<br />

discovered that cancer arises from defective<br />

cells. Virchow noticed a connection between<br />

long-term infections and cancer, and he was<br />

absolutely right.”<br />

In spite of such evidence, however, the idea<br />

that cancer provokes and then escapes an<br />

adaptive immune response has once again<br />

reared its “attractive” head – mostly on the<br />

basis of experiments with mice that had deficient<br />

immune systems. And in many of these


models, researchers used chemicals to induce<br />

cancer. Those results have to be interpreted<br />

with caution, Thomas says.<br />

“To demonstrate that the immune system is<br />

responding specifically to a tumor you need to<br />

find T cells which have been triggered by an<br />

antigen that belongs to the tumor,” he says.<br />

“Since no unique antigens had been identified<br />

in those tumors, you cannot prove this.”<br />

So many of his lab’s recent efforts can be<br />

summed up this way: Does our immune system<br />

offer any specific protection from tumors<br />

that arise spontaneously, from random mutations<br />

in genes? And do tumors escape by<br />

interfering with mechanisms of immunity? It<br />

has been challenging to phrase these questions<br />

in the language of rigorous experiments,<br />

particularly since most laboratory<br />

models of cancer immunology have relied on<br />

chemicals, or transplants of tumor cells<br />

between mice. With any transplant, there is a<br />

risk that the immune system will treat cancer<br />

like an infection. The results might not say<br />

SV40 virus<br />

123 Part Two: Identity crisis


Martin Textor<br />

anything about how it responds to tumors<br />

that develop spontaneously.<br />

Another issue is secondary infections.<br />

Burkitt’s lymphoma, described at the beginning<br />

of this part of the book, demonstrated<br />

that a human loses the ability to cope with<br />

the Epstein Barr virus – and the tumors that it<br />

leads to – if its immune system is simultaneously<br />

being challenged by the malaria parasite<br />

or other infectious agents. This means<br />

that cancer studies should ideally be carried<br />

out in an environment where animals are free<br />

from infectious agents.<br />

Despite these challenges, Thomas’ lab has<br />

found ways to approach the question. They<br />

have been working with a mouse bearing a<br />

gene called Tag, which is an oncogene – a molecule<br />

that promotes the development of<br />

cancer.<br />

Tag comes from a virus called SV40, which can<br />

infect rodents, primates, and humans.<br />

Researchers believe that Tag blocks the functions<br />

of at least two genes in mammals. One<br />

of these is p53, whose job is to launch a selfdestruct<br />

program in damaged cells. If there<br />

are mutations in p53, or if it can’t do its job<br />

due to interference by Tag or another molecule,<br />

aberrant cells may survive and cause<br />

cancer.<br />

These factors have made SV40 a tool to study<br />

cancer for 40 years. In the 1960s Galina<br />

Deichman, of the Cancer Research Center in<br />

Moscow, did a series of experiments that used<br />

the virus to create tumors in mice, in hopes of<br />

learning how the immune system responds.<br />

SV40 could infect cells in her rodents but<br />

could not reproduce itself. It functions by<br />

sneaking its genes into the cell’s genome.<br />

When the cell produces Tag it may become<br />

cancerous, divide, and build a tumor.<br />

Giving adult animals the virus would cause an<br />

infection that would provoke an immune system<br />

response. Instead, Deichman gave SV40<br />

to newborn hamsters whose immune system<br />

hadn’t yet been “switched on.” At this early<br />

stage their bodies could not treat it as an<br />

infection. Most of the hamsters infected as<br />

newborns developed tumors after a long<br />

latency period, similar to the slow pace at<br />

which spontaneous tumors arise in humans.<br />

Deichman discovered that the animals could<br />

be vaccinated against the molecule. Giving<br />

the virus to adult hamsters with a healthy<br />

immune system triggered a strong, immediate<br />

response. SV40 was fought like any other<br />

infection. T cells wiped it out. They even wiped<br />

out the disease in hamsters that had been<br />

given the virus as newborns. This meant that<br />

the immune system found cells that had been<br />

infected much earlier and destroyed them as<br />

it coped with the current infection. The hamsters<br />

given an SV40 “vaccine” never developed<br />

tumors. All the control animals, who were<br />

infected but didn’t receive the immunization,<br />

died from tumors.<br />

Vaccines are preventative tools; they don’t<br />

work if given too late in an infection. The<br />

same was true here. If Deichman waited too<br />

long with the second round of infection, until<br />

about two weeks before cells had grown into<br />

Part Two: Identity crisis<br />

124<br />

detectable tumors, the “vaccination” no<br />

longer helped. Either the body became tolerant<br />

when a certain number of cells became<br />

cancerous, or something else was going on.<br />

This shows, Thomas emphasizes, that giving<br />

the virus to newborns didn’t make them tolerant<br />

to SV40; they weren’t mistaking its<br />

genes for “self.” Even so, tumors weren’t being<br />

eliminated by the immune system.<br />

With the development of new genetic engineering<br />

techniques in the 1980s, scientists<br />

could directly implant Tag into the genome of<br />

mice; it was no longer necessary to use viruses<br />

to introduce cancer-causing genes. The<br />

results were the same. In young animals<br />

whose genomes contained Tag, very few cells<br />

– if any – produced Tag, so the mice did not<br />

learn to recognize it as “self.” If the animals<br />

were vaccinated early enough, they could be<br />

protected even from tumors arising from<br />

their own cells. This was getting very close,<br />

Thomas says, to a good model for human<br />

cancers.<br />

Another study helped set the stage for<br />

Thomas’ current work. In 1962, while working<br />

with cancers induced by chemical carcinogens,<br />

Lloyd Old and his colleagues at the<br />

Sloan-Kettering Institute for Cancer Research<br />

in New York took tumor cells from one mouse<br />

and transplanted them into another. They discovered<br />

that mice which received fewer transplanted<br />

cells were more likely to develop cancer<br />

than mice that received a lot of cells. You<br />

might expect things to be the other way<br />

around – unless high numbers of cells resembled<br />

an infection, while lower numbers


escaped. By staying under the radar of the<br />

body’s first wave of rejection, the tumors<br />

seemed to be avoiding the second.<br />

��<br />

Inserting the Tag gene into the mouse<br />

genome made a laboratory model that<br />

came close to imitating human cancer, but it<br />

could be refined further. Thomas and a postdoc<br />

in his group, Gerald Willimsky, added<br />

another feature to the artificial Tag gene. They<br />

intended to make a conditional form of Tag,<br />

which would allow them to switch it on in a<br />

specific tissue, at any time they chose. (This<br />

type of conditional gene is described in the<br />

story “How to roust a resting cell.”) In the<br />

process, something unplanned happened –<br />

one of those lucky accidents that sometimes<br />

make great science.<br />

Building the artificial Tag gene meant inserting<br />

it into the mouse genome with some additional<br />

instructions. One bit of code added to<br />

the gene was a promoter sequence, telling<br />

cells to make the molecule. A second bit of<br />

code acted as a brake, preventing that from<br />

happening. The combination meant that Tag<br />

wouldn’t be produced. The mouse needed to<br />

be mated with a second animal outfitted with<br />

another gene, one which released the brake in<br />

a specific tissue. Any offspring of the mice that<br />

inherited both genes could be used to study<br />

cancer in that tissue. By mating the Tag mouse<br />

with various other strains, the researchers<br />

could study tumors throughout the body.<br />

Then the accident happened: they noticed<br />

that all of the Tag animals were developing<br />

tumors even without being mated to the second<br />

strain. “At first this was very disappointing,”<br />

Thomas says. “It looked as if something<br />

had gone wrong, and because it takes months<br />

or years to develop such a mouse, it meant we<br />

might have to start all over again.”<br />

A closer look at the animals, however, showed<br />

that something else was going on.<br />

“Occasionally a rare random event, in random<br />

cells in the body, would release the brake and<br />

activate the Tag gene,” Thomas says. “It was<br />

impossible to predict when or where these<br />

mutations would happen. Most of the mice<br />

developed one tumor; a few had more than<br />

one, and in some cases there were metastases.<br />

This happened in all sorts of tissues: in<br />

the kidneys, liver, bone, spleen, and elsewhere.”<br />

In fact, the situation was close to what happens<br />

in spontaneous human cancers: somewhere<br />

in the body, a mutation in a gene leads<br />

a tumor to arise. Not completely by accident –<br />

but with a little luck – Thomas and Gerald had<br />

created a new mouse model that closely imitated<br />

the development of human tumors.<br />

They also had a perfect situation to study<br />

interactions between the tumors and the<br />

immune system. When cells became cancerous<br />

and produced Tag, they can potentially be<br />

noticed by the immune system. Are they? “We<br />

now had a way to check,” Thomas says, “by<br />

125 Part Two: Identity crisis<br />

looking to see if the body produced antibodies<br />

against Tag. And we found them.”<br />

The scientists repeated the types of vaccination<br />

experiments that Deichman had carried<br />

out with the SV40-infected mice. None of the<br />

mice immunized with Tag developed tumors<br />

over a two-year period. All of the mice in the<br />

control group, which had not received the vaccine,<br />

died from tumors. Additionally, when<br />

tumor cells were transplanted into young,<br />

tumor-free transgenic mice, they were rejected.<br />

It was further proof that the body treats a<br />

transplant like an infection, and that cancer<br />

cells themselves don’t come equipped with<br />

special characteristics that help them evade<br />

the immune system. They can be easily fought<br />

if the mouse body becomes aware of them in<br />

the right way. But tumors that arise spontaneously<br />

in the body are different. The immune<br />

system sees them, but then nothing happens.<br />

T cells don’t take further action.<br />

Thomas shows me a copy of the paper he<br />

wrote with Gerald, who has now left the <strong>MDC</strong><br />

and holds a post at the Charité’s Institute of<br />

Immunology, where Thomas has a second lab.<br />

“The move happened for formal reasons, and<br />

it was not always easy for him,” Thomas says.<br />

“Today to be a real success as a scientist, you<br />

have to publish often, and quickly. This project<br />

didn’t fit well into that scheme – it took seven<br />

years.”<br />

It was worthwhile, he says, because it is the<br />

strongest proof yet that there is a fundamen-


tal difference between the way the body<br />

rejects a spontaneously generated tumor and<br />

one which is transplanted – even if it is the<br />

same tumor. This is at least a partial answer<br />

to one of Thomas’ questions: cancer studies<br />

based on transplants do not necessarily shed<br />

light on some of the critical issues in immune<br />

responses to cancer.<br />

“Our results open a new set of questions<br />

because they show that tumors aren’t escaping<br />

by just hiding from the immune system,”<br />

Thomas says. “ The body sees Tag long before<br />

tumors appear and makes lots of copies of T<br />

cells that recognize it, just as humans produce<br />

T cells that recognize some kinds of tumors –<br />

you find them in the tumor itself, in lymph,<br />

and in blood – although there are no signs of<br />

the body rejecting the tumor. The body does<br />

not behave as if it has been immunized.<br />

Tumors grow by somehow creating a tolerance<br />

in the immune system, and this is something<br />

we are investigating quite intensively<br />

now.”<br />

��<br />

Iam having trouble getting used to this new<br />

way of thinking. If anything in today’s medical<br />

culture should be regarded as a devil, it is<br />

The Berlin Wall<br />

cancer. Strange that our bodies are able to<br />

come to a truce with this terrible disease. Too<br />

bad that it doesn’t last longer. On the other<br />

hand, the pact lasts longer than the 24 years<br />

given to Faust – long enough for most of us to<br />

have and raise our children. After that, the<br />

evolutionary peacekeeping force packs up and<br />

leaves.<br />

The last revolution in medicine began with an<br />

understanding of the causes of infectious diseases.<br />

Fighting those illnesses could take<br />

advantage of an immune system that had<br />

already been primed to cope with them. It<br />

seems there will be no such natural help for<br />

spontaneous cancers and the other threats<br />

that strike in old age. It’s the reason “molecular<br />

medicine” is necessary – if evolution hasn’t<br />

given the body defenses, they will have to be<br />

invented.<br />

On the other hand, human intelligence – and<br />

science – are also the products of evolution, so<br />

in a way biotechnology and modern medicine<br />

can be thought of as humans’ way of evolving<br />

new defenses. This idea was expressed eloquently<br />

a few years ago in a talk given by biologist<br />

Hubert Markl, former President of the<br />

German Research Council and of the Max<br />

Planck Society. He said:<br />

We should see ourselves not as some kind of<br />

fallen angel, alien intruder, some aberrant<br />

or deranged scourge of nature, but as its<br />

constituent and heir. And not only as one<br />

constituent part of nature among many<br />

others, just an arbitrarily chosen biological<br />

species, but as a unique, a quite extraordinary<br />

kind of natural species, through which<br />

nature entered into an entirely new stage of<br />

evolution; a species that not only participates<br />

in its future evolution like any other<br />

species, but that increasingly commands<br />

and determines this future, for better or<br />

worse. In evolving the human species,<br />

nature, as it were, began to take control of<br />

its own future, to give it purposeful direction,<br />

to assume responsibility for its own<br />

future development... From such a comprehensive<br />

evolutionary perspective, human<br />

technological and economic inventiveness is<br />

nothing other than nature’s way of intentionally<br />

acting upon itself and forming its<br />

own future.<br />

��<br />

Jens Reich


Thomas’ work suggests that it could very<br />

well be possible to retrain the immune<br />

system to fight spontaneous cancers – if<br />

markers can be found, and diseases are<br />

caught early enough. Putting this into practice<br />

is the goal of most of the efforts of<br />

Thomas, his lab, and some of his peers.<br />

One of those colleagues is Wolfgang Uckert, a<br />

former postdoc in Thomas’ lab, now head of<br />

his own group at the <strong>MDC</strong>. Although he’s still<br />

relatively young, Wolfgang has been a firsthand<br />

witness to more of the history of the<br />

Buch campus than most of the scientists who<br />

work here. He earned his PhD here when this<br />

area of Berlin belonged to the German<br />

Democratic Republic, doing most of his work<br />

in the oldest laboratory building on campus. It<br />

was originally built to serve as Oskar Vogt’s<br />

Institute for Neurological Research; it was this<br />

building that Vogt studied Lenin’s brain, looking<br />

for traces of genius; later it was integrated<br />

into one of East Germany’s prestigious<br />

Academies of Science.<br />

Practicing science was different in the GDR,<br />

Wolfgang says. Today’s researchers are world<br />

travelers, expected to attend conferences<br />

across the globe, often working as postdoctoral<br />

fellows in two or three countries before<br />

attaining a long-term job as an independent<br />

researcher. Things were much different in the<br />

East. It wasn’t necessarily impossible to travel<br />

outside the Iron Curtain under the East<br />

German regime, but it was extremely complicated.<br />

“At one point I was offered an opportunity<br />

to go to Belgium for three months,”<br />

Wolfgang says. “By that time I had a family.<br />

They stayed here, which was probably the reason<br />

I was allowed to go.” But when he was<br />

127 Part Two: Identity crisis<br />

Wolfgang Uckert and a piece of the Berlin Wall<br />

offered a longer-term position in California,<br />

the state wasn’t as forthcoming.<br />

Jens Reich, <strong>MDC</strong> emeritus and former head of<br />

the <strong>MDC</strong>’s biocomputing groups, had extensive<br />

first-hand experience with the Stasi. He<br />

remembers a case when a postdoctoral fellow<br />

in the group went abroad and failed to return.<br />

“This was considered extremely serious, and it<br />

had huge effects,” Jens says. “The lab was<br />

somehow thought to be responsible – how<br />

could this happen without our knowledge?<br />

Had he said anything? Had any plans been<br />

discussed?” The event led to months of questioning.<br />

Jens received regular summons for<br />

sessions of hour-long questioning.<br />

“The summons was always disguised as an an<br />

invitation to give an administrative report, in<br />

the 1970s to the Ministry of Science and<br />

Technology, and in the 80s to the headquar-


ters of the Akademie der Wissenschaften. You<br />

only learned that they were Stasi once you<br />

arrived at the first meeting and they introduced<br />

themselves.”<br />

“Things were also quite ‘exciting’ when foreign<br />

researchers visited the campus,” Jens<br />

says. “Once we had the exceptional opportunity<br />

to host an international conference here<br />

in Buch. The entire event, which lasted for several<br />

days, was ‘scripted’ and rigorously<br />

observed. Yet it had to be done in such a way<br />

that the visitors didn’t notice anything. This<br />

led to some ridiculous situations. It was a<br />

circus.”<br />

The issue, he says, was meetings between foreigners<br />

and East German citizens. “You were<br />

allowed to have contact with foreign guests<br />

only with explicit permission, and you had to<br />

report afterwards. The ‘inofficial staff’ had to<br />

report directly to the Stasi; everyone else sent<br />

their reports to the directorate, and if the<br />

Stasi had questions they came back the same<br />

way. The local authorities tried to keep things<br />

from getting too ridiculous by ‘not seeing<br />

everything,’ but if the Stasi got wind of this,<br />

they were in for rough times. A prominent<br />

incident happened in 1978 when the<br />

European Biochemistry Congress was held in<br />

Dresden. A young co-worker had to accompany<br />

a prominent guest from the United States<br />

on an excursion to the museums of Weimar.<br />

During the trip he bared his soul about the<br />

restrictions on international exchange and<br />

other nuisances. The American turned this<br />

into a very critical report in Nature, and that<br />

led to a major scandal. The organizers had to<br />

confess and repent and write a long, tortuous<br />

readers’ letter to Nature disclaiming the<br />

report.”<br />

Under those circumstances, people stuck<br />

together. The lab wasn’t only a workplace; it<br />

had an important social function. Wolfgang<br />

tells how members of his lab kept a collection<br />

jar for money. “You had to pay a little bit if you<br />

forgot to clean up or if you broke something,”<br />

he says. “At the end of the year we always took<br />

what was in there and went out to dinner in<br />

Part Two: Identity crisis<br />

128<br />

downtown East Berlin. There were really nice<br />

restaurants that were open late.”<br />

One year he got home from the event at<br />

around 11pm. He turned on the television – “It<br />

was easy to get the West Berlin stations here<br />

in the city” – and was amazed to see people<br />

dancing on top of the Berlin Wall. “At first I<br />

couldn’t believe it,” he says. “When I realized<br />

what was going on I told my wife that we had<br />

to go back downtown; we had to be there.”<br />

History was in the making in the city, an event<br />

not to be missed. He quickly asked a neighbor<br />

to come watch their small children and headed<br />

for the Kurfürstendamm. “Later I came<br />

back with my daughters and a hammer to get<br />

a piece of the Wall,” he says. “The nice part<br />

was on the Western side, where it had been<br />

painted with graffiti. We wanted to send<br />

pieces to all our friends and relatives. It turned<br />

out we didn’t get a big enough piece, and we<br />

had to go back again.”<br />

��


Can the immune system be taught something<br />

it hasn’t evolved to do – to combat<br />

spontaneous cancer? A major theme of<br />

Wolfgang’s work has been to develop “designer”<br />

T cells that might be able to do this. He has<br />

been working on the problem in collaboration<br />

with Thomas Blankenstein. The two men met<br />

as the <strong>MDC</strong> was established, as Thomas set<br />

up his new lab. He was looking for staff members<br />

and interviewed some of the scientists<br />

who had worked in the Academy on campus.<br />

“It was a bit of an odd situation at the time,<br />

hiring someone who was a year older than<br />

myself,” Thomas says. “But it turned out to be<br />

a very good fit. I was very pleased when<br />

Wolfgang received his professorship and<br />

could set up his own group, and we have<br />

worked closely all of these years.”<br />

“Adaptive immune therapy” aims to remove T<br />

cells from a patient’s body and outfit them<br />

with a new receptor protein that allows them<br />

to recognize cancer. A major hurdle is to find<br />

the right target for the receptor – a protein<br />

that only appears on the surface of tumor<br />

cells (the chapter “Capturing a firefly” shows<br />

how difficult that can be). But even before<br />

that happens, there are a number of technical<br />

challenges to solve in rebuilding T cells and<br />

making sure that they behave properly.<br />

To explain, Wolfgang shows me a greatly oversimplified<br />

drawing of a T cell. The receptors<br />

look like the letter “Y” and are made of two<br />

proteins. These are produced by the cell’s<br />

genes and then mounted on its surface with<br />

the assistance of other proteins called CD3s.<br />

The two parts of the receptor are called alpha<br />

and beta chains, he says, but it’s fine if I want<br />

to think of them as “left” and “right” arms. An<br />

artificial receptor also needs two arms, which<br />

means adding two new genes to the cell.<br />

“In real life things are much more complex,”<br />

Wolfgang says. “A typical T cell uses its genes<br />

to make about 40,000 identical copies of its<br />

receptor, that means about 40,000 left arms<br />

and 40,000 right. When we give that cell the<br />

genes to make a second receptor, it doesn’t<br />

129 Part Two: Identity crisis<br />

stop making those. Instead, it starts producing<br />

a lot more arms.”<br />

The cell assembles the proteins into pairs and<br />

moves them to the membrane. “Remember<br />

that it now has four ‘arms’ to make receptors,<br />

and that it can’t recognize where the different<br />

arms have come from. It starts pairing them<br />

up, a left and a right. A fourth of those pairs<br />

will consist of just the new left and right arms<br />

we have added. Another fourth will consist of<br />

the cell’s native left and right arms. The rest<br />

will be mixtures, one of our arms paired with<br />

one of the cell’s.”<br />

It is hard to predict how these different sets of<br />

receptors will behave. “Suppose you put one<br />

of these cells into the body and the new<br />

receptor recognizes a protein on a tumor,”<br />

Wolfgang says. “That will trigger the T cell to<br />

make billions of copies of itself. That’s what<br />

you want, because it creates the cells needed<br />

to fight the tumor. But at the same time,<br />

you’re making cells that may be looking for<br />

extra targets whose identity you can’t predict.<br />

The extra receptors on the cell might not recognize<br />

anything, in which case you’re safe. But<br />

if they come across a molecule anywhere in<br />

the body that they can dock onto, there’s likely<br />

to be trouble.”<br />

For example, the “hybrid” receptors that combine<br />

one foreign and one native arm might<br />

recognize one of the body’s proteins and trigger<br />

an autoimmune reaction. The native<br />

receptor might also recognize a “self” molecule.<br />

“Usually this is no real problem, because<br />

the body won’t produce many of these cells<br />

and they can be kept under control,”<br />

Wolfgang says. “But if you suddenly start having<br />

billions of them, because the body is using<br />

the cells to fight a tumor, things might get<br />

out of hand.”<br />

Is there any way to force the T cell to produce<br />

only the artificial receptors? I ask.<br />

“What we’ve discovered is that some receptors<br />

are very successful at moving to the surface<br />

of the cell and others are not,” Wolfgang


Martin Textor<br />

says. “In some cases when we engineer the T<br />

cells, they produce almost only the new receptor,<br />

and very little of the native one. In other<br />

cases it’s the other way around. One of the<br />

major activities of the group right now is to<br />

try to figure out what determines how successful<br />

a receptor will be. Because theoretically<br />

this means we can build a molecule that<br />

pushes aside the native one, and then we’ll<br />

have eliminated these kinds of concerns that<br />

we have about interference.”<br />

Successful receptors may be the ones that are<br />

best at recruiting partner proteins like CD3, he<br />

says. And in this regard, mice have amazingly<br />

potent molecules. If a T cell receptor is transplanted<br />

from a mouse into a human cell, it<br />

wins out over human receptors every time.<br />

His lab is now trying to figure out which features<br />

of the mice molecules make this happen.<br />

If they succeed, they should be able to<br />

custom design receptors on the surfaces of T<br />

cells that will replace the cell’s molecules.<br />

They may even be able to add more than one<br />

type of receptor – in some cases it might<br />

make the T cell more effective – and control<br />

the “dosage” at which each type is made by<br />

the cell.<br />

Even then, he admits, biological processes are<br />

so complex that there’s always a chance<br />

something unforeseen will happen. Wolfgang<br />

is well aware that making the jump from the<br />

lab to humans always involve risks. The first<br />

gene therapy experiments carried out in the<br />

U.S. and elsewhere – mostly using viruses to<br />

transport healthy genes into cells – have led<br />

to both successes and tragedy.<br />

“So while we have been designing these new<br />

types of cells, we thought we would also build<br />

in some safeguards,” he says. “We’re learning<br />

to equip our T cells with an emergency switch<br />

that allows us to shut things down fast if<br />

something unexpected happens.”<br />

PhD student Elisa Kieback has been pursuing<br />

that project, adding an extra bit of code to the<br />

receptor. The code allows the molecule to be<br />

Part Two: Identity crisis<br />

130<br />

recognized by antibodies that a doctor can<br />

give a patient if the cells start to misbehave.<br />

The antibodies tag the engineered T cells and<br />

call up a rapid immune system response,<br />

which destroys the cells.<br />

��<br />

ne conclusion from our work is that<br />

“Oif specific markers for tumors can be<br />

identified, and we intervene at an early stage<br />

in the disease, it is likely that patients can be<br />

immunized,” Thomas Blankenstein says.<br />

This may not be possible in the case of most<br />

spontaneous cancers that arise because of<br />

mutations in a random gene. But some types<br />

of tumors behave in a stereotypical way, produced<br />

by well-known oncogenes. If scientists<br />

can identify them, it may be possible to detect<br />

problems in patients at an early stage and<br />

prepare a defense.<br />

It may be easier in cancers associated with<br />

viruses, but for most of these diseases it has<br />

also been difficult to develop vaccines against


the viruses, and once they have settled in, the<br />

cancers they cause are difficult to treat. “Once<br />

classical therapies such as chemotherapy,<br />

radiation and surgery have failed, there are<br />

few options,” Thomas says. “Some recent successes<br />

using T cells trained against the viruses<br />

have encouraged us to continue to work in<br />

that direction, but developing T cells as therapeutic<br />

tools is a long-term, labor-intensive<br />

project.”<br />

In spite of that fact, clinical trials are currently<br />

underway with some of the T cells that<br />

Thomas and Wolfgang have engineered. Yet<br />

the real goal, making T cell therapies a standard,<br />

working tool in the doctor’s arsenal, is<br />

obviously beyond the work of a single lab. If<br />

you had a whole institute at your disposal, I<br />

ask him, and could devote it entirely to T cell<br />

therapies, what would you do?<br />

Thomas’ eyes light up. He doesn’t have a<br />

whole institute, but maybe he’s got something<br />

even better. In 2006 he made a proposal<br />

to the German Research Council to create a<br />

special, national research area dedicated to<br />

just this theme. The project unifies the work<br />

from laboratories in Berlin and Münich,<br />

including those of Thomas and Wolfang,<br />

other groups from the <strong>MDC</strong> and Charité, the<br />

Helmholtz Institute for Molecular<br />

Immunology, and a number of universities.<br />

Together they will try to develop specific<br />

therapies for some forms of leukemia, renal<br />

cell carcinomas, and other forms of cancer, as<br />

well as pursuing basic issues in the use of T<br />

cells.<br />

The ultimate aim, Thomas says, is to develop<br />

systems that can be used to produce T cells<br />

targeting any molecule that the researchers<br />

choose – ideally within a week. “The strategy<br />

would be to have a stock of T cell receptors on<br />

the shelf, which can be taken down and used<br />

to produce T cells targeting any molecule we<br />

choose – ideally within a week. They will be<br />

used to target life-threatening viral diseases<br />

or diverse forms of cancer, including some<br />

solid tumors.”<br />

131 Part Two: Identity crisis<br />

How long will it be until customized T cells<br />

become a widely-used tool in the fight<br />

against cancer? What will have to happen?<br />

“I’m always asked this question,” Thomas<br />

sighs. “I refuse to answer with a specific time<br />

frame. Reality will always be different than<br />

what we hope.”<br />

Still, he says, he strongly believes – as does<br />

nearly everyone working in the field – that in<br />

the long term, the strategy will work. When<br />

that day comes, it will be possible to add<br />

clauses to the body’s pact with cancer, or to<br />

rewrite the agreement, extending it by many<br />

more years. Or maybe the pact can be discarded<br />

entirely, in favor of therapies that prevent<br />

cancers from ever arising in the first place.


The decorators and<br />

engineers of Versailles<br />

It’s an amazing place to visit, but few would want to live there. To<br />

modern taste, the palace of Versailles on the outskirts of Paris is<br />

overwhelming. The French of the 17th and 18th centuries indulged themselves<br />

in an orgy of ornamentation that today seems oppressive.<br />

Decoration reaches into every cranny, every piece of furniture and picture<br />

frame, the vaulted ceilings, and even empty surfaces. Extravagent<br />

golden picture frames steal attention from the paintings they enclose.<br />

Walls, tabletops and floors are inlaid with intricate patterns. Gilded<br />

flowers are printed on wallpaper and stitched into cloth. Fireplaces are<br />

crafted from mottled marble rather than plain stone. Thousands of bits<br />

of glass in each of the chandeliers catch splinters of light and diffract<br />

them in every direction, where they are caught by thousands of mirrors<br />

and infinitely reflected.<br />

How could you walk in the shoes, or get anything done in those clothes?<br />

The lace on the sleeves would eventually drive you mad. In the court of<br />

Louis XIV, heads were wigged and faces were powdered; legs were<br />

stockinged and hands gloved, as if to protect the body from any contact<br />

with the real world. Historical sources say that the perfumes of<br />

Part Two: Identity crisis<br />

132


Hans Baumeister


Versailles were equally extravagent; it makes<br />

you glad that there have been two centuries<br />

since the Revolution for the air to recirculate.<br />

In the French Baroque, ornaments were as<br />

pervasive as bacteria in the ecosphere.<br />

Neither literature nor music escaped the<br />

infection. Nearly every note written by<br />

François Couperin is ornamented. Many of his<br />

trills and vibratos are unique; he included a<br />

handbook at the beginning of some of his<br />

works to explain how to play them. In this<br />

music, however, ornaments are more than<br />

style. They create textures; they glide between<br />

awkward intervals; they are fundamental to<br />

the harmony, leaving a listener hanging in dissonances<br />

that ache to be resolved. French<br />

Baroque compositions can’t be played without<br />

them; the music has no depth and makes<br />

no sense without the ornaments.<br />

The escape of art and life into an artificial<br />

world couldn’t go on forever. In music, popular<br />

taste shifted toward the Italian style of Vivaldi,<br />

whose virtuosity was based on acrobatics<br />

rather than subtlety. The Revolution stripped<br />

off the wigs and powders and forced the nobility<br />

to give up their cake for breads. It’s no sur-<br />

prise that a lifestyle so overburdened with<br />

style itself came crashing to an end.<br />

��<br />

Is there anything in the living world that is<br />

simply ornamental? Charles Darwin threw<br />

this idea into doubt. Variation is everywhere,<br />

but it is a motor of evolution: sooner or later,<br />

even tiny details come under the scrutiny of<br />

natural selection. A patch of color on an animal<br />

may attract a mate or a predator; the rich<br />

palette of flower petals serves as signposts for<br />

bees and butterflies, who see more colors<br />

than humans do.<br />

A type of chemical decoration found in cells –<br />

the attachment of sugars (carbohydrates) to<br />

proteins – probably began as an accident in<br />

ancient unicellular organisms. Over the<br />

course of evolution, this has expanded to a<br />

level of ornamentation worthy of Versailles. In<br />

the French Baroque, ornamentation rarely had<br />

a function. But in the cell, combinations of<br />

sugars and proteins (called glycoproteins)<br />

became useful; many glycoproteins, in fact,<br />

don’t function without carbohydrates.<br />

Organisms evolved hundreds of specialized<br />

molecules to do the decorating.<br />

Part Two: Identity crisis<br />

134<br />

Anything found so widely throughout nature<br />

has to be important, and every day<br />

researchers discover new ways that sugars<br />

influence the behavior of proteins. “We’re only<br />

now getting an idea of how versatile glycoproteins<br />

are,” says Hans Baumeister, of the<br />

company Glycotope on the Berlin-Buch campus.<br />

“In some cases carbohydrates’ functions<br />

are still not known, but in other cases they are<br />

essential, sometimes even vital for the development<br />

or survival of an organism.”<br />

Hans is the Chief Operating Officer of Glycotope,<br />

a company founded in 2001 by CEO Steffen<br />

Goletz and Andreas Eckert. Andreas is CEO and<br />

founder of another well-established business on<br />

campus: Eckert & Ziegler Strahlen- and<br />

Medizintechnik. Steffen was a group leader at the<br />

<strong>MDC</strong> until he and Andreas formed the two startups,<br />

first NEMOD and one year later Glycotope,<br />

with Steffen as CSO of both companies. Later<br />

Steffen became CEO of Glycotope. Hans joined<br />

NEMOD as a scientist and was one of the first<br />

employees at Glycotope. Over the past few years<br />

he has gravitated toward the business side of<br />

the company. Now he heads Glycotope's service<br />

activities as COO, and Franzpeter Bracht<br />

serves as the company's CFO/CBO.


The company fills an important niche in the<br />

biotech market, Hans says. “Most research and<br />

medical applications require the production<br />

of high quantities of proteins,” he says. “That’s<br />

usually done in cultures of bacteria or animal<br />

cells, sometimes in the animals themselves.<br />

These cells have to make human proteins<br />

properly – especially when those molecules<br />

are destined for medical applications.<br />

Achieving that hasn’t been as easy as people<br />

hoped, and that has partly been due to glycosylation.<br />

The company is based on unique<br />

technology that helps us fix this problem.”<br />

A century ago a lot of biochemists were<br />

deeply interested in sugars, but with the coming<br />

of molecular biology the focus shifted<br />

largely to DNA, RNA and proteins. One reason<br />

is the complexity of sugar molecules and the<br />

way they are built by cells. Sugars are not<br />

directly encoded in the genome but are constructed<br />

by other molecules in a complex<br />

series of chemical transformations. They are<br />

made from many different simpler sugars<br />

(monosaccharides) which form a much wider<br />

variety of sequences than are found in DNA,<br />

RNA and proteins. This makes their physical<br />

structures are very complex. All of these<br />

things have been obstacles to their investigation.<br />

“Basically this means that in most cases<br />

you know the protein you have in hand but<br />

you don’t know exactly what sugar is<br />

attached to the protein, or in some cases even<br />

where it is attached,” Hans says. “There aren’t<br />

the shortcuts you find with proteins – you<br />

can’t look at a DNA sequence and produce a<br />

complete catalogue of the sugars that a certain<br />

organism can make.”<br />

The surface of an animal cell is covered with<br />

carbohydrates – just about every molecule<br />

found on its surface or secreted from it has<br />

been modified through the attachment of<br />

sugars. “That includes protein hormones,<br />

growth factors, cytokines, blood factors and<br />

antibodies,” Hans says. “You know how important<br />

antibodies are in research and medical<br />

applications. Well, most antibodies for medical<br />

applications have been made by inserting<br />

genes into hamster cells, turning them into<br />

antibody factories. As the molecules are<br />

made, the cells attach sugars. But those are<br />

different sugars than the ones you find in<br />

human cells. The result is that the molecule<br />

you’ve made for therapy often works differently<br />

than it should. Even worse – it might<br />

induce an unwanted immune response<br />

against the therapeutic agent.”<br />

��<br />

Each organ and tissue is a composer, developing<br />

its own unique language of ornamentation;<br />

each has its own specialized types<br />

of sugars. The molecules change with every<br />

135<br />

new context. Cells glycosylate proteins differently<br />

as they specialize and when other<br />

changes happen – for example, if they start to<br />

build a tumor.<br />

Most antibodies are decorated with just one<br />

sugar, but it’s an important one. It is attached<br />

to the part of the antibody that does not bind<br />

to its antigen, but to receptors that activate<br />

effector cells such as macrophages or natural<br />

killer cells; these home in on the target cell,<br />

which is recognized by the antibody. This is an<br />

important mechanism for tumor therapy. “So<br />

an optimal glycosylation of an antibody is<br />

necessary for the therapeutic application of<br />

that antibody,” Hans says. Sugars are also<br />

important in the functions of hormones,<br />

growth factors and cytokines, powerful signaling<br />

molecules released by glands throughout<br />

the body.<br />

“An important part of their sugar decoration<br />

are sialic acids, the only sugars that are negatively<br />

charged,” Hans says. “Different numbers<br />

of these acids are added during the process of


glycosylation. Having a lot of sialic acids<br />

increases the molecule’s lifespan, because<br />

hormones that don’t have any are grabbed by<br />

the liver and removed from circulation. On the<br />

other hand, having too many of them may<br />

decrease the activity of the hormone. So it’s<br />

another thing that has to be done right when<br />

you try to produce a glycosylated molecule in<br />

a cell line.”<br />

Glycotope has found a way to address many of<br />

these problems through the development of<br />

new cell lines. The company’s scientists began<br />

with human cells and modified them to carry<br />

out different kinds of glycosylation. One type of<br />

cell sialylates molecules to a high degree.<br />

Another adds none, unless the cell is fed with<br />

with sugar. How much is added to molecules<br />

depends on how much the scientists feed it.<br />

“This gives us a very wide spectrum of possibilities,”<br />

Hans says. “Using human cells is<br />

already an improvement over animal systems<br />

that often glycosylate in the wrong way. And<br />

this system can be adapted to produce the<br />

molecules needed, with a range of carbohydrates<br />

attached to them.”<br />

There has been a huge demand for the company’s<br />

technology, called “GlycoExpress,” he<br />

says. Most of the clients are pharmaceutical<br />

and biotech companies developing their own<br />

therapeutic antibodies or other biotherapeutics;<br />

Hans says they can significantly improve<br />

their products with GlycoExpress by reducing<br />

the necessary doses to be injected and avoiding<br />

unwanted side effects.<br />

Glycotope has several products of its own that<br />

are currently in development. One is a new<br />

“therapeutic antibody” which recognizes and<br />

binds to target the carbohydrate of a protein<br />

on the surface of tumor cells. Scientists have<br />

known for over 40 years that a protein called<br />

mucin-1 acts abnormally during cancer and<br />

that sugars attached to that protein change<br />

during tumorgenesis.<br />

Mucin-1 was discovered when scientists<br />

examined a gelatinous substance that accu-<br />

mulated around breast tumors. An article<br />

written in 1965 by J.B. Adams of the Prince of<br />

Wales Hospital in Sydney, Australia, reported<br />

that the substance contained high levels of<br />

sugars and a protein. In the meantime, scientists<br />

have discovered that the protein is heavily<br />

loaded with sugars, and that large<br />

amounts of it accumulate inside and around<br />

many types of cancer cells.<br />

Scientists at Glycotope discovered a new antibody<br />

that very specifically recognizes the sug-<br />

Part Two: Identity crisis<br />

136<br />

ars on mucin-1 of diseased cells. It was an<br />

exciting find, Hans says, because only tumors<br />

seemed to glycosylate the protein in this particular<br />

way – not normal cells. “Any time you<br />

find a unique marker, it opens therapeutic<br />

possibilities. In this case the idea was to develop<br />

this antibody as therapeutic agent to<br />

destroy the tumor.”<br />

Antje Danielczyk and the other scientists of<br />

Glycotope know how to turn an antibody into<br />

a therapeutic device: they attached a radioac-


tive isotope to the antibody which will destroy<br />

the tumor cell when delivered to it.<br />

Alternatively, the antibody may be used as a<br />

therapeutic device as is (it’s called a “naked”<br />

antibody as opposed to the radioactively<br />

labeled antibody). That works if the antibody<br />

is able to mediate killing of the tumor cells<br />

through macrophages or natural killer cells.<br />

The next step was to test PankoMab, as the<br />

new drug candidate was called, in both the<br />

radiolabeled and “naked” format in tumor-<br />

loaded mice. (The name was inspired by<br />

Pankow, a neighborhood of Berlin that lies<br />

between the city center and Buch.) The tumor<br />

quickly absorbed PankoMab and was very efficiently<br />

killed using either form.<br />

In parallel, the scientists conducted tests of<br />

PankoMab in the test tube and discovered<br />

another important feature: it binds much<br />

more strongly to mucin-1 on the surface of living<br />

cells than to mucin-1 molecules that have<br />

been released by cells that have already died.<br />

137 Part Two: Identity crisis<br />

That is often a problem in antibody therapies,<br />

because loose protein fragments may suck up<br />

most of the injected antibodies.<br />

“So PankoMab has passed all the tests we<br />

could throw at it to determine whether it<br />

could be used in therapies,” Hans says. “First<br />

the antibody has been humanized. That’s an<br />

important step during development of therapeutic<br />

antibodies that are originally synthezised<br />

as mouse antibodies – they would<br />

induce strong immune reactions against the<br />

drug itself if you applied them in humans.<br />

Now it’s being produced under ‘good manufacturing<br />

conditions’ (GMPs) to give it the<br />

quality needed to be tested in clinical trials in<br />

patients. If it passes that hurdle, we’ll be getting<br />

closer to the day it can be used in any<br />

patient who is waiting for treatment. It has<br />

also been an important proof of principle that<br />

our approach – focusing on improving the<br />

sweet part of a protein – will lead to a much<br />

more powerful therapeutic drug.”<br />

��<br />

It’s a Friday afternoon in the spring and I am<br />

standing in the sun outside Hans’ office. He<br />

passes two coffee cups through the window,<br />

locks up his room, and joins me outside.<br />

Glycotope is housed in the Erwin Negelein<br />

building, which was built a few years ago as<br />

the campus set up the biotechnology park. It’s<br />

fitting that Hans works in a building named<br />

after Negelein, a biochemist and departmen-


138<br />

tal director during the GDR years. Negelein<br />

was a biochemist who had developed cell<br />

lines from mice into tools for the investigation<br />

of cancer.<br />

I ask if the campus research groups take<br />

advantage of the company’s services.<br />

“We have had a few collaborations with<br />

groups here,” he says. “Martin Lipp’s lab needed<br />

a human version of an antibody; the versions<br />

they had obtained from mice would have<br />

caused immune reactions. On the average I’d<br />

guess that we work with campus groups two<br />

or three times a year. Of course we would welcome<br />

more business directly from the institutes,<br />

but they’re not our typical target group.<br />

Making humanized and glycooptimised antibodies<br />

is usually too expensive for them. Most<br />

of our clients are pharmaceutical and biotech<br />

companies around the whole world.”<br />

There are many other advantages to being on<br />

campus. “Berlin itself is a good location,” he<br />

says. “The people we’d like to hire are generally<br />

eager to come; they like the city. Setting up<br />

labs is not expensive compared to other<br />

places – that’s very important for young companies<br />

trying to keep costs down.”<br />

The scientific life of the campus also has<br />

advantages. “You get into fruitful scientific<br />

discussions, and can get access to instruments<br />

and expertise that you might need.<br />

Obviously it’s a good place to be in terms of<br />

recruitment. A lot of the people who worked<br />

from the beginning in the company started<br />

off in one of the campus institutes.”<br />

We talk for a while about the differences<br />

between working in a basic research institute<br />

and working for a company. Is it hard for<br />

someone to switch to industry? I ask.<br />

“I guess I’m a good person to ask,” he says. “I<br />

began as a basic researcher and have now<br />

completely moved into the business side of<br />

science.”<br />

Hans grew up in Berlin – in the southwestern<br />

part of the city, where he still lives with his


family. After receiving his diploma from the<br />

Freie Universität of Berlin, he worked on his<br />

PhD in Hamburg, with a focus on cell receptors.<br />

His introduction to industry began with<br />

his next move, to New Jersey, where he<br />

worked for the Swiss pharmaceutical company<br />

Hoffmann-La Roche.<br />

“At that time the company was still heavily<br />

supporting fundamental science,” he says.<br />

“They had an institute which was really just<br />

doing basic research. It was an interesting<br />

place – the bosses were Americans, and just<br />

about everybody else was European or Asian.<br />

They were doing work on things like transcription<br />

factors, and I don’t think they had produced<br />

anything in the way of products for a<br />

market for about thirty years.”<br />

Eventually the institute went the way of so<br />

many research laboratories funded by pharmaceutical<br />

companies – they have been<br />

closed down, or the work has been pushed in<br />

a much more practical direction. This has left<br />

a gap that institutes like the <strong>MDC</strong> or FMP<br />

need to fill – if medical science hopes to take<br />

advantage of the wealth of new knowledge<br />

emerging from basic science.<br />

Maybe an increasing movement of students<br />

and other scientists into industry will help.<br />

The transition isn’t as hard as you might<br />

think, Hans says, although there are important<br />

differences between the cultures.<br />

“You have to get used to working on projects<br />

that are much more focused,” he says. “The<br />

projects we do don’t have the breadth of<br />

things going on in places like the <strong>MDC</strong> or FMP.<br />

And they have a different aim. In science the<br />

measure of success is publications; in industry<br />

it’s patents – at least for researchers.<br />

Publications come thereafter. And research is<br />

organized in milestones. When a project or<br />

potential product to be developed misses too<br />

many milestones, it is skipped.”<br />

139 Part Two: Identity crisis<br />

Companies have more of an engineering<br />

mentality, he says. It’s a different kind of creativity.<br />

“We’re less looking for some completely new<br />

thing than to take something that exists, fiddle<br />

with it, and improve it until it works,” he<br />

says. “It’s enormously satisfying to see that<br />

something you have worked hard on is used.<br />

The high point is when it is developed into a<br />

product that makes a difference in patients’<br />

lives.”<br />

One of the biggest contrasts, Hans says, is<br />

that the science carried out in companies has<br />

clients.<br />

“When you begin a project you immediately<br />

imagine its potential clients,” he says. “It’s not<br />

all about a patent and a product, but who will<br />

be using it. How much will it cost to develop<br />

something? Is there enough money and are<br />

there enough potential clients? Can you<br />

excite an investor about an idea?” One of the<br />

benefits when a company offers services, he<br />

says, is that income can be rolled into the<br />

development of new ideas.<br />

Later in the process come the real clients.<br />

“They’re trying something and having problems,<br />

and then your job becomes communication,”<br />

he says. “It’s a part of the job that I like<br />

very much. Sitting down with someone and<br />

figuring out how to solve technical issues.”<br />

From sugars on proteins to powerful therapeutic<br />

tools. Too bad the French never found a<br />

way to use their extravagent ornaments to<br />

heal the sick or feed the poor. The vast gardens<br />

devoted to plants for perfumes might<br />

have been used to raise medical herbs. That<br />

might not have changed anything significant<br />

in history, or the way science is practiced<br />

today. On the other hand, some elements of<br />

French style might have survived. Instead of<br />

the black suit and colored tie that Hans sometimes<br />

wears when he represents the company,<br />

he might have to wear clothing that<br />

looked like wallpaper. And a high powdered<br />

wig.


Waking a sleeping<br />

beauty and other tales of<br />

ancient genes<br />

Good stories, like living things, evolve over time. Nuances are<br />

added or taken away; tales become confused and merge; some<br />

are forgotten entirely, while others seem to live forever. The tale of the<br />

sleeping beauty has gone through most of these transformations. An<br />

early form of it appears in The Book of 1001 Nights, a collection of Arabic<br />

tales from the tenth century. Giambattista Basile, a soldier from Naples,<br />

inspired the modern story with a fable in his book Pentamerone,<br />

published in 1634. Charles Perrault revived it in the Mother Goose<br />

Tales (1697).<br />

The version collected by the Grimm brothers a hundred years later is the<br />

one most of us have heard. A young princess pricks her finger on a spindle,<br />

fulfilling a curse placed on her at birth. This sends her and everyone<br />

else in the castle into a deep sleep. They are frozen in time for a century,<br />

until a prince fights his way through the thorny thicket surrounding<br />

the castle to deliver his kiss. Everyone wakes up, and they all live happily<br />

ever after.<br />

The Grimms left out some of the darker details of the story. In Basile’s<br />

account the princess doesn’t wake up until she has borne two of the<br />

Part Two: Identity crisis<br />

140


Zoltán Ivics<br />

prince’s children. And the German version<br />

leaves out the entire second half, told by<br />

Basile and Perrault: one day the prince rides<br />

off to battle. He leaves his young family in the<br />

care of the queen mother, who is a descendant<br />

of an ogre, creatures with a taste for<br />

human flesh. She develops an insatiable<br />

hunger for the children and the princess and<br />

demands that they be cooked for her dinner. A<br />

quick-thinking cook switches the children for<br />

tender lambs. When it comes time to prepare<br />

the princess, he has a hard time finding an<br />

appropriate meat. (Perrault writes, “The young<br />

queen was twenty years old, not counting the<br />

hundred years she had been asleep, and the<br />

cook was hard-pressed to find a meat that<br />

would be tough enough.”) He finally settles<br />

on a cow. Naturally the queen discovers she<br />

has been deceived; naturally, in the nick of<br />

time, the prince returns to save the day.<br />

Genomes, too, are a sort of storybook, and<br />

some of the genes they contain have been put<br />

into a deep sleep by mutations. Even a prince<br />

can’t wake them – now it will take a genetic<br />

engineer.<br />

��<br />

Zsuzsanna Izsvák<br />

Mythopoeisis is the art of making and<br />

reshaping myths, and two labs at the<br />

141 Part Two: Identity crisis<br />

<strong>MDC</strong> are engaged in a sort of mythopoesis of<br />

genomes. Zsuzsanna Izsvák and her husband<br />

Zoltán Ivics head neighboring research groups<br />

on the second floor of the Walther Friedrich<br />

House. They are studying and waking ancient<br />

genes for use in research and medical applications.<br />

Genetic engineering is a collection of methods<br />

that allow scientists to remove genes,<br />

rebuild them, or transplant them from one<br />

organism to another. Usually this involves<br />

cutting parts from an existing DNA molecule<br />

and pasting them in somewhere else. In the<br />

1950s scientists found naturally-occurring


enzymes<br />

in bacteria able to<br />

do this cut-and-paste job.<br />

Genetic engineering was born when<br />

these molecules were adapted for use in<br />

other situations. In 1972 Janet Mertz and Ron<br />

Davis, of Stanford University, used them to<br />

remove a gene from one bacteria and insert it<br />

into another. A year later, their colleagues<br />

Stanley Cohen and Annie Chang, with Herbert<br />

Boyer of the University of California in San<br />

Francisco, used the procedure to transplant a<br />

gene between different species.<br />

Today these techniques are biology’s most<br />

powerful methods to learn what genes do<br />

and how they work. But they aren’t perfect,<br />

and they can’t be used as widely as scientists<br />

would like. “We need more tools, particularly<br />

to work with vertebrates,” Zoltán says. “At the<br />

time we started, the zebrafish was becoming<br />

an important model organism in the lab, but<br />

we didn’t have tools to manipulate its<br />

genome. Things are much better in the<br />

mouse, but some tools for mice don’t work in<br />

fish or mammals such as rats, or in many<br />

other important model organisms.”<br />

Zsuzsanna explains that they got started in<br />

the field almost by chance, under tragic circumstances.<br />

Zoltán had been awarded a sixmonth<br />

fellowship to work on zebrafish genetics<br />

in the United States, at the University of<br />

Minnesota. Zsuzsanna was working on her<br />

dissertation at the time – “which I could write<br />

anywhere,” she says – and she came along<br />

anticipating getting a job in the same lab. But<br />

immediately after getting off the plane at the<br />

airport, they were greeted with a piece of<br />

devastating news: the boss of their laboratory,<br />

Kevin Guise, had just died. He was<br />

one of many early victims of the AIDS<br />

epidemic, having received a transfusion<br />

of blood containing the virus. It<br />

had not yet become standard practice<br />

to screen blood supplies for<br />

signs of HIV. “We were shocked,”<br />

Zsuzsanna says, “because we didn’t know<br />

how seriously ill he was, and up until the last<br />

minute he had been arranging things for our<br />

stay. Apparently he had taken care of my visa<br />

paperwork from the hospital.”<br />

Arrangements were quickly made for them to<br />

join the nearby lab of Perry Hackett, who had<br />

helped introduce zebrafish at the university,<br />

and who hoped to find ways to do genetic<br />

engineering with them.<br />

“At that time we were getting our first real<br />

look at whole animal genomes and the huge<br />

amount of sequence – in humans, about 98<br />

percent – that doesn’t contain genes,” Zoltán<br />

says. “The amazing thing was that about half<br />

the genome consisted of the same little bits<br />

of material, repeated over and over again.<br />

Such repetition can be so substantial that<br />

particular sequences, such as one called the<br />

Alu repeat, are represented over a million<br />

copies in the human genome. A similar phenomenon<br />

is seen in fish. Over the course of<br />

evolution, fragments of sequence had somehow<br />

copied and pasted themselves all over<br />

the genome, like little viruses.”<br />

Some viruses were, in fact, known to behave in<br />

a similar way. Retroviruses such as HIV reproduce<br />

by squeezing their own genes into the<br />

DNA of the host cell. The cell treats these molecules<br />

like its own genes, using them to create<br />

new RNAs and proteins. But rather than working<br />

for the cell, these become the raw materials<br />

to make new copies of the virus.<br />

Cells had genes of their own that behaved<br />

this way. In the 1930s geneticist Barbara<br />

McClintock observed them in plants and animals<br />

and called them transposons, or jumping<br />

Part Two: Identity crisis<br />

142<br />

genes. One type of transposon cuts itself out<br />

of the DNA sequence and pastes itself in at a<br />

new place, sometimes after being copied. A<br />

huge amount of extra DNA has accumulated<br />

this way – apparently about half the human<br />

genome. Transposons were probably a major<br />

reason that the small genomes of the first<br />

cells eventually became so huge. This creates<br />

DNA sequences that may eventually be used<br />

in genes, which has been a major factor in the<br />

largest scale of evolution.<br />

But for an individual, or a single species, transposons<br />

are usually dangerous. “The fact that<br />

they are usually pasted into random places<br />

means they may land in the middle of another<br />

gene,” Zoltán says. “That scrambles up its<br />

information and usually destroys a molecule<br />

that the cell needs.”<br />

One goal of their research, Zoltán says, is to<br />

find out how cells protect themselves. The


most straightforward defense involves random<br />

mutations, which are just as likely to<br />

affect transposons as anything else. In most<br />

cases mutations have a negative effect on a<br />

gene, damaging its functions. Damage to a<br />

transposon would actually have a positive<br />

effect because it would protect other genes.<br />

The evolutionary pressure to deactivate jumping<br />

genes is so strong that nearly all of the<br />

transposons in animals have stopped working.<br />

In the mid-1990s active jumping<br />

genes had been found in bacteria, flies,<br />

worms and plants, but none were known in<br />

vertebrates.<br />

Mutations would probably wipe out all transposons,<br />

Zoltán says, if they didn’t find an<br />

escape route. That seems to involve jumping<br />

to other species, probably by riding as a hitchhiker<br />

inside a bacterium or virus. The new<br />

host offers fertile new DNA to colonize. The<br />

transposon will copy itself over and over until<br />

mutations catch up and disable it. Certain<br />

types of transposons can be found in a wide<br />

range of species, which is strong evidence for<br />

the hitchhiker hypothesis.<br />

Mutations don’t always wipe out a transposon;<br />

sometimes they spin one molecule off<br />

into a whole family of related forms. An<br />

ancient family called Tc1/mariner can be found<br />

throughout the living world, from bacteria to<br />

fish, frogs, and humans. But, at least in vertebrate<br />

species all of these transposons have<br />

been inactivated.<br />

“It would be nice to have one that works,”<br />

Zoltán says. “They make up so much of animal<br />

genomes that they need to be studied. And<br />

some transposons encode all the tools needed<br />

for cutting, copying and pasting DNA.<br />

They’re like entire genetic engineering toolboxes<br />

packed into the simplest form, into just<br />

one gene.”<br />

143<br />

��<br />

If nature hadn’t left any intact toolboxes<br />

lying around, Zsuzsanna and Zoltán began<br />

to wonder if they couldn’t make their own.<br />

There were instructions to be found everywhere<br />

in the transposon sequences that littered<br />

every genome. Something was wrong<br />

with nearly all of them – but each one had different<br />

things wrong. By comparing all the versions,<br />

the two scientists thought they might<br />

be able to reconstruct the sequence of the<br />

original gene, or at least a version that<br />

worked.<br />

Since the goal was to create tools that could<br />

be used to transfer genes from one species to<br />

another, it was best to start with a transposon<br />

that had already proven its ability to spread<br />

widely. The Tc1/mariner transposons seemed<br />

to be the perfect choice.<br />

“Mutations had inactivated all the versions<br />

we found,” Zoltán says. “But they still contained<br />

enough information to show how they<br />

had worked. Originally the gene encoded<br />

everything necessary to cut itself out and to<br />

paste itself back in elsewhere.”<br />

One goal of the analysis was to find the most<br />

recent working version of the transposon. It<br />

would have undergone fewer mutations,<br />

which would mean having to repair fewer<br />

broken parts to get it to jump again. A computer<br />

analysis of the genes of various types of<br />

fish identified one family of transposons in<br />

which the average difference was small – only<br />

five percent, despite the large evolutionary<br />

distance between their hosts. This means<br />

that this transposon family was very successful<br />

in invading several fish species and has<br />

been inactivated thereafter. The gradually<br />

accumulating average difference, five percent,<br />

functions as a molecular clock and the ”golden<br />

age” of this invader family was estimated<br />

to be around 10 million years ago.<br />

By comparing several copies, Zoltán and<br />

Zsuzsanna could make a hypothesis about<br />

the identity of the chemical “letter” that


Andrea Schmitt<br />

occurred most often in each position of the<br />

DNA code in the transposon family. This consensus<br />

recreated the genetic information of a<br />

hypothetical invader which had successfully<br />

colonized several salmonid fish species. Some<br />

parts of the genes had undergone more<br />

mutations than others. Evolution usually conserves<br />

the most crucial features of an animal<br />

– also of a gene – while less important characteristics<br />

rapidly change. Zsuzsanna and Zoltán<br />

found five regions that had been well conserved<br />

in the protein-encoding part of the<br />

transposon. These represented domains – the<br />

main functional modules of the transposase<br />

protein. As the lab reconstructed each module,<br />

they tested their functions one by one.<br />

As in all genes, the code in the transposon is<br />

used to make a messenger RNA molecule,<br />

which is then sent out of the cell nucleus to<br />

be made into protein. In this case the protein<br />

was the transposase, the scissors used to cut<br />

their own gene out of the genome. Once it<br />

has been made, the transposase has to get<br />

back into the nucleus. One of the transposase<br />

domains acts as a delivery code, telling the<br />

cell where to deliver it. Inside the nucleus, the<br />

transpose has to grip the DNA strand where it<br />

will make a cut. Another domain handles this<br />

job.<br />

When the labs had completed the new version<br />

of the molecule with the corrected<br />

spellings, it gave them a new, artificial transposon.<br />

They named it Sleeping Beauty –<br />

“because it had been sleeping in the fish<br />

genome a very long time” – and began testing<br />

whether it could be used to transfer genes<br />

into laboratory cultures containing human<br />

cells. After allowing the cells to grow, they<br />

checked to see if Sleeping Beauty had been<br />

inserted into their genomes. The results were<br />

positive. Not only had the delivery and insertion<br />

parts of the molecule worked, there was<br />

evidence that the gene had copied itself and<br />

jumped to new positions in the chromosomes.<br />

Part Two: Identity crisis<br />

��<br />

The couple brought along transposons<br />

when they joined the <strong>MDC</strong> in 1999.<br />

Zoltán was given a position as a group leader,<br />

and Zsuzsanna took some time off to give<br />

birth to twin boys – they both give a tired<br />

144<br />

smile (now) about the difficulty of managing<br />

a family and their careers. Shortly afterwards,<br />

Zsuzsanna won a position of her own as part<br />

of a prestigious award from the European<br />

Union: the European Young Investigator<br />

Award. Competition was tough, she says; of<br />

880 applications, only 25 researchers were<br />

ultimately given funding. The grant helped<br />

her establish her own laboratory at the <strong>MDC</strong>.<br />

The successful resurrection of Sleeping<br />

Beauty suggested that it might be possible to<br />

develop a palette of transposons which could<br />

be used in different organisms. This would<br />

widen their potential as genomic tools. The<br />

labs went on to try to isolate new active<br />

transposons from vertebrate species. They<br />

found that Tc1/mariner elements in the<br />

genome of the Northern leopard frog were<br />

unusually well conserved and contained only<br />

a few mutations. Csaba Miskey, a postdoctoral<br />

fellow in Zoltán’s group, needed to fix only<br />

two mutations in one of the transposon<br />

copies to make it jump again. The new transposon,<br />

keeping the Grimm brothers’ nomenclature,<br />

was named Frog Prince.


In 2006 the two labs produced yet another<br />

working transposon called Hsmar1. “We think<br />

that one of our distant ancestors, a primate,<br />

originally acquired this gene about 50 million<br />

years ago, probably from a virus,” Zoltán says.<br />

“It reproduced itself heavily for about 13 million<br />

years until mutations had disarmed all<br />

the copies. It had already spread quite far. The<br />

human genome alone contains about 200<br />

inactive copies.”<br />

Csaba decided to try to rebuild the original<br />

Hsmar1 gene. Csaba and Balázs Papp, a specialist<br />

in biocomputing at the University of<br />

Manchester (UK), began analyzing the<br />

“spelling” of human Hsmar1 genes.<br />

Comparing different versions revealed common<br />

features that had probably belonged to<br />

the first Hsmar1, and this allowed them to<br />

reconstruct the DNA sequence of the original<br />

gene. On the average, they discovered that<br />

about eight percent of each gene copy had<br />

changed through mutations.<br />

One of the molecules, however, had undergone<br />

far fewer mutations: a gene called SET-<br />

MAR had changed less than three percent.<br />

SETMAR is a member of the Hsmar1 family. It is<br />

an unusual case where cells still use it to<br />

make RNA and proteins, but it no longer possesses<br />

the cut-and-paste tools to allow it to<br />

jump. “Some transposons took a different evolutionary<br />

course,” Csaba says. “When they suffered<br />

mutations, instead of just becoming<br />

junk, they became domesticated – they<br />

remained genes and took on other functions.<br />

We’re not completely sure what those are in<br />

the case of SETMAR, but there’s some evidence<br />

that it helps repair breaks in DNA.”<br />

Recently a laboratory at the University of<br />

Oxford had shown that the SETMAR protein<br />

retained at least partial transposon activity.<br />

This suggested that rebuilding the original<br />

might create a functional transposon.<br />

Csaba and his colleagues repeated the strategy<br />

that had been used with Sleeping Beauty<br />

and Frog Prince: analyzing Hsmar1 copies in<br />

search of letters that had changed the least,<br />

a.<br />

b.<br />

The Sleeping Beauty transposon system.<br />

a. The components: Orange = the target gene that scientists want to move. It is engineered to lie<br />

between DNA sequences (white arrows) that can be recognized by the DNA-cutting transposase<br />

molecule (purple).<br />

b. When the transposase and target gene are produced in the same cells, the transposase recognizes<br />

the binding sites and docks on. It draws them together and cuts the ends, releasing the target gene.<br />

Cellular mechanisms repair the break in DNA and insert the transposon into a new site.<br />

then building a new molecule with the most<br />

frequent spellings. But their new version of<br />

the gene didn’t jump.<br />

Returning to the drawing board, Csaba and<br />

Zoltán thought they knew why. The situation<br />

was a bit like going to a city and counting the<br />

number of blue-eyed and brown-eyed people,<br />

finding a majority of citizens with blue eyes,<br />

then coming to the logical conclusion that<br />

the founders of the city must have had blue<br />

eyes. But that might not be true. If in the<br />

intervening generations, people with blue<br />

eyes had a lot more children, this might skew<br />

the Tobias results. Jursch The same thing could be true of<br />

Hsmar1. A particular mutant version of the<br />

145 Part Two: Identity crisis<br />

molecule might have reproduced more than<br />

others. If so, the “average” DNA sequence<br />

would be biased in that direction.<br />

Csaba tried a new type of computer analysis<br />

in which, instead of building an average<br />

sequence from all the Hsmar1 molecules, he<br />

reconstructed their family tree. This revealed<br />

four spelling differences between the “averaged”<br />

DNA sequence and the original. When<br />

the scientists incorporated the older spellings<br />

into their artificial gene, they now had a working<br />

transposon.<br />

Along with the 200 inactive versions of<br />

Hsmar1, the human genome also contains


about 2,500 copies of a fragment of the gene<br />

called MITE. Further experiments with the<br />

reconstructed gene showed that the ancestral<br />

Hsmar1 was probably also responsible for<br />

copying and shuffling these fragments<br />

through the genome. The new version of the<br />

gene made MITEs jump at a much higher rate<br />

than the entire gene, supporting a hypothesis<br />

that smaller elements can be moved easier<br />

than larger ones.<br />

The new transposon shows a high level of<br />

activity, Zoltán says; it is just as efficient as the<br />

two other transposons that the group has<br />

rebuilt. This means that it should be possible<br />

to turn the gene into a new tool to investigate<br />

gene functions in mammals and other organisms,<br />

just as Sleeping Beauty has been made<br />

into one.<br />

��<br />

Transposons have surely been a big force<br />

in the history of life. By expanding the<br />

size of genomes, they have introduced a lot of<br />

raw material for evolution to work on.<br />

Mutations have turned some of the<br />

sequences into new genes; over 100 human<br />

molecules are thought to have been domesticated<br />

from parts of transposons. One of them<br />

is HARBI1, a molecule that evolution has conserved<br />

in all vertebrate species. It originally<br />

evolved long ago as part of a transposon in an<br />

early ancestor, probably a fish.<br />

“HARBI1 stems from a family of genes called<br />

PIF/Harbinger that is found widely in vertebrates<br />

and plants,” Zoltán says. “There are<br />

copies in maize and rice which still work, but<br />

as far as we know there aren’t any active<br />

forms in vertebrates.”<br />

Reactivating the molecule might shed light<br />

on how it and other transposons have evolved<br />

into useful genes like HARBI1. There was<br />

another reason Zoltán, Zsuzsanna and their<br />

colleagues were interested in the transposon:<br />

it encoded a second molecule with unknown<br />

functions. Its architecture closely resembled<br />

another human protein called NAIF1. Any part<br />

of a transposon that wasn’t understood<br />

might give new insights into their behavior<br />

and evolution.<br />

HARBI1 was taken on by a French postdoc<br />

named Ludivine Sinzelle, whom the scientists<br />

had met in the course of an EU grant. “There is<br />

not a very large transposon community in<br />

Europe, and we are quite scattered about –<br />

such grants are great opportunities to bring<br />

people together,” Zsuzsanna says. “Ludivine is<br />

a wonderful person – extremely nice and<br />

cheerful. She has just left and everyone was<br />

sorry to see her go. I met her at a conference<br />

and we decided that she should come to the<br />

<strong>MDC</strong>. I found two months of money from one<br />

grant, and Zoltán put together four more<br />

months, and finally we had nine months and<br />

she could come.”<br />

The nine months stretched into two years;<br />

although she wanted to return to France, the<br />

project was going so well that Ludivine couldn’t<br />

bring herself to leave. Her reward was the<br />

completion of a working version of the<br />

Harbinger transposon. The labs have used it to<br />

analyze the functions of its two genes.<br />

One result is a better understanding of where<br />

the transposon becomes inserted into the<br />

genome. Some of the genes jump to random<br />

places; others insert themselves at specific<br />

DNA sequences. Vladimir Kapitonov, who had<br />

done most of the computational work on the<br />

project, had looked at the neighborhoods of<br />

the fish transposons and found a 17-letter<br />

code that seemed to be the target. But it was<br />

hard to be sure; these sequences, too, had<br />

undergone mutations.<br />

The new, intact Harbinger allowed the scientists<br />

to test the insertion code in living cells.<br />

The gene had jumped 46 times in their experiments,<br />

and Ludivine carefully inspected the<br />

sites where it landed. Her computer analysis<br />

revealed a sequence of 15 letters that told the<br />

gene where to insert itself. “Twelve of those<br />

are the same in the 17-letter code in<br />

zebrafish,” Zoltán says. “This is interesting<br />

because it means that even when you trans-<br />

Part Two: Identity crisis<br />

146<br />

plant the gene from one species to another, it<br />

looks for the same target.”<br />

This fulfills one aim of the project: to produce<br />

tools that can be used to insert foreign or<br />

altered genes into new species. It also puts<br />

them one step closer to another of the scientists’<br />

goals – to turn transposons into tools for<br />

medical therapies.<br />

“Molecular biologists all over the world hope<br />

to develop gene therapies – methods of delivering<br />

healthy versions of genes to people with<br />

genetic diseases or other illnesses,” Zoltán<br />

says. “Perhaps the greatest barriers we have to<br />

overcome are the problem of delivery – getting<br />

foreign DNA or RNA into the right types<br />

of cells – and then getting molecules to<br />

behave the right way once they have arrived.<br />

This means that healthy molecules need to be<br />

inserted into the genome at specific places, so<br />

that they don’t interfere with the activity of<br />

another gene.”<br />

A clinical trial conducted in hospitals<br />

throughout the world in 2002 and 2003<br />

showed how crucial this is. Researchers were<br />

using a retrovirus to deliver a protein called<br />

the interleukin 2 receptor to patients suffering<br />

from a severe immune system disease<br />

called X-SCID. People with defective versions<br />

of this molecule are unable to produce B and<br />

T cells, which allows even trivial infections to<br />

run rampant in their bodies. They have to<br />

spend their life in a sterile “bubble” unless a<br />

bone marrow transplant from a close relative<br />

takes hold.<br />

But in several patients the virus inserted the<br />

gene’s DNA into the worst possible position,<br />

near the control region of a gene called LMO2.<br />

This molecule is a proto-oncogene, which<br />

means that it is known to cause cancer if it<br />

undergoes changes. Somehow the arrival of<br />

the gene triggered an uncontrollable wave of<br />

replication among T cells – basically a<br />

leukemia-like form of cancer.<br />

“The only way to avoid these problems is to<br />

know exactly where a gene will land in a


patient’s DNA,” Zoltán says. “That could be difficult<br />

with a virus, which is very complex – transposons<br />

are much simpler.” Ideally, a transposon<br />

could be “programmed” to insert itself into precise<br />

places in the genome by rewriting the tools<br />

that told it where to go. This is a crucial step in<br />

turning the molecules into therapeutic tools. In<br />

June 2007, Zoltán and Zsuzsanna used Sleeping<br />

Beauty to prove that this could be done. That was<br />

an important step toward obtaining a new grant<br />

from the European Union, which is uniting the<br />

efforts of transposon researchers from several<br />

countries to develop the genes into therapeutic<br />

tools.<br />

In a completely different approach, transposons<br />

are not simply used as harmless delivery vehicles,<br />

but are designed to be highly mutagenic. Indeed,<br />

transposable elements can insert themselves<br />

into many places in the genome that can cause<br />

mutations. This is how Barbara McClintock discovered<br />

them in maize: jumps of transposons<br />

resulted in changes of the color of the kernels.<br />

After the completion of several genome sequencing<br />

projects, the new challenge is to understand<br />

the function of each gene. “Transposons can be<br />

exploited as discovery tools,” Zsuzsanna says,<br />

“and we are adapting them to do systematic<br />

studies of the genome.”<br />

Most of the studies in this book involve disturbing<br />

a gene’s functions through knock-outs and<br />

then studying an organism to see how it gets<br />

along without the molecule. Many of the largescale<br />

methods for doing so involve chemicals that<br />

cause DNA to mutate; when a gene is non-functional,<br />

researchers can study its effects on the<br />

animal. But chemical mutagens often alter only a<br />

few letters in the code of a gene, which makes<br />

these approaches time-consuming, Zsuzsanna<br />

says. “It takes a lot of work in the laboratory to<br />

identify the mutated gene– and often you need a<br />

very large screening facility to look at the animals.<br />

That is different if you disrupt the gene<br />

with a transposon, because the mutated gene<br />

will be tagged and you can use this label to find<br />

out where it has gone just by scanning the<br />

genome for the tag.”<br />

147 Part Two: Identity crisis


Andras, Zuzsanna, Zoltán, and Zsombor<br />

Once the transposon has been inserted into<br />

an organism’s genome, it can be used to spin<br />

off new mutants. It can be made to jump to a<br />

new location simply by breeding an animal<br />

that carries the transposon with another that<br />

has the molecules needed to cut it out and<br />

paste it somewhere else. Each offspring that<br />

inherits both parts of the toolbox will be like a<br />

new genetic experiment: the transposon may<br />

disrupt a new gene by jumping somewhere<br />

else. One can imagine the power of this<br />

approach: each sperm can carry a new mutation.<br />

Extra information can be packed into the<br />

transposon so that it becomes activated only<br />

in certain tissues or at particular stages of an<br />

animal’s development. Zsuzsanna and Zoltan<br />

are adapting their transposon tools for systematic<br />

discovery of genes that are thought<br />

to play a role in human disease. They are concentrating<br />

their efforts on genetic screens to<br />

uncover genes involved in different types of<br />

cancer and cardiovascular disease.<br />

��<br />

Even for a couple working on common<br />

projects in neighboring labs, life is not all<br />

about science at the bench. In the little spare<br />

time they have, Zoltán and Zsuzsanna have<br />

Part Two: Identity crisis<br />

148<br />

been encouraging the development of molecular<br />

biology in their native country. In 2007<br />

the journal Nature printed a letter they wrote<br />

about the current situation in Hungarian science.<br />

The letter was inspired by a news article<br />

about a government plan to cut budgets,<br />

restructure its science academies, and privatize<br />

parts of the national research infrastructure.<br />

Both scientists feel a strong attachment to<br />

their home country. “On the positive side, we<br />

received an excellent education there,”<br />

Zsuzsanna says. “Hungary continues to produce<br />

very good students and postdocs. But<br />

then they all leave to take jobs abroad. The


country does not offer good opportunities for<br />

senior scientists, and the proposals that are<br />

circulating now will further cut the already<br />

limited amount of money that is spent on<br />

basic research. This is amazing in a country<br />

which has received the highest number of ‘per<br />

capita’ Nobel prizes.”<br />

The Hungarian academy was founded in 1825<br />

to ensure the long-term scientific interests of<br />

the nation. The new policies run directly<br />

counter to that aim. “With its short-sighted<br />

outlook,” their letter says, “the government<br />

does not realize that its ‘reform’ activities<br />

undermine the development of a healthy<br />

innovation chain, and drown the personal creativity<br />

that has been an asset and a source of<br />

pride for Hungarian science.” Zoltán and<br />

Zsuzsanna were among 2,500 scientists who<br />

signed an open letter to the prime minister<br />

urging that budgets not be cut.<br />

Running labs, watching out for Hungary’s<br />

interests, and taking care of twins leaves very<br />

little time for anything else. Still the couple<br />

tries to keep a balance. We talk for a while<br />

about the difficulty of managing a family life<br />

and two scientific careers.<br />

“We never discuss science after ten p.m.,”<br />

Zsuzsanna says.<br />

149 Part Two: Identity crisis<br />

“Actually it’s earlier than that,” Zoltán says.<br />

“Usually.”<br />

“Yes, if we talked about it all the time, the<br />

twins would be left out.”<br />

I try to imagine sitting at the dinner table<br />

with two seven-year-olds when the discussion<br />

turns to the nuances of jumping genes.<br />

As good as Zoltán and Zsuzsanna are at<br />

explaining their work in simple terms, that<br />

would surely make the twins do some jumping<br />

of their own.<br />

Fairy tales would surely work better.


Interlude:<br />

These rats don’t dance<br />

to the Red Hot Chili Peppers<br />

If you’re ever on the road in Central Eastern Africa with a shovel and<br />

feel the need to dig a hole, you’re likely to strike a tunnel or burrow<br />

made by one of the strangest creatures in the world: the naked mole rat,<br />

which is neither a mole nor a rat. Its hairless, wrinkled appearance is<br />

only the first of its oddities. Where you would expect to see an eye there<br />

is only a small black hole – the rodent is blind. Ears are knobbed bumps<br />

with no openings. It darts around frenetically on thin legs that carry it<br />

equally quickly whether moving forward or backwards.<br />

The naked mole rat prefers to live out its life huddled with its brothers<br />

and sisters in huge masses. It has a social life like that of bees. A colony<br />

has one queen and two or three fertile males; the rest are drone workers.<br />

It is the world’s only cold-blooded mammal. It is capable of shutting<br />

down its metabolism to a low rate, which is probably responsible for its<br />

very long lifespan; the animals have been known to live for up to 25<br />

years. The time that isn’t spent sleeping, or huddling, is usually spent<br />

digging. That’s how the colony finds the roots and fleshy tubers that it<br />

feeds on. (It also eats its own feces, and nourishes its offspring on<br />

them.)<br />

Interlude: These rats don’t dance to the Red Hot Chili Peppers<br />

150


Gary Lewin


Underground, in a confined space that may be<br />

inhabited by up to 300 densely-packed<br />

rodents, the air is quickly exhausted. Huge<br />

amounts of carbon dioxide accumulate,<br />

enough to kill any other mammal. But the<br />

naked mole rat survives. This suffocating<br />

atmosphere may contribute to one the<br />

strangest things of all about the animal. It<br />

lives in a world without sight – and largely<br />

without pain.<br />

Pain is so fundamental that sometimes it<br />

seems the world is divided into two categories:<br />

things that cause it and those that<br />

don’t. Interestingly, most animal species find<br />

the same things painful – a pin-prick, a hot<br />

flame, or acid. We are all too familiar with the<br />

sharp pains of wounds and inflammations.<br />

And then there is capsaicin, the active ingredient<br />

in chili peppers. When it comes in contact<br />

with the tongue, the skin, or any other tissue,<br />

it burns. Birds don’t notice it, though, which is<br />

good for the plant – they propegate its seeds.<br />

Christiane Wetzel, M. Benkovic, Jing Hu, Gary Lewin<br />

And the naked mole rat barely notices it at all,<br />

even when the substance comes in contact<br />

with an open wound.<br />

��<br />

Gary Lewin’s interest in chili peppers is<br />

more scientific than culinary. The main<br />

topic of his work at the <strong>MDC</strong> is how nerves<br />

detect and pass along sensory information<br />

like pressure and pain. “One reason that capsaicin<br />

is interesting is that we think it stimulates<br />

the same system, at least partly, which<br />

transmits the pain of inflammation,” he says.<br />

“We knew that some of the elements of that<br />

system were lacking in the naked mole rat. So<br />

alongside our work with more traditional laboratory<br />

animals such as mice, we’ve recently<br />

been developing this new unusual creature as<br />

a model to study pain.”<br />

So if you want to see one of the animals, you<br />

don’t have to go to Africa. There are some in<br />

the basement of the Max Delbrück house.<br />

Gary leads us down long, dimly lit corridors, a<br />

labyrinth of grey walls and doors. Somehow<br />

he finds the right one and lets us into a small,<br />

cramped room. Two naked mole rats scurry<br />

around in a glass cage. It’s warm in the room<br />

and warmer in the cage – a prerequisite to<br />

keeping the animals alive. He reminds us that<br />

they are cold-blooded.<br />

Seven or eight of us have accompanied him<br />

downstairs, a spontaneous mid-day field trip<br />

organized by Elisabeth Kujawa-Schmeitzner<br />

of the director’s office. There are two administrators<br />

from the PhD program, two more from<br />

the conference office, and the grants officer.<br />

None of them scientists, but they have all<br />

heard Gary talk about the animals and their<br />

curiosity has been piqued.<br />

After having seen pictures of the creatures –<br />

which don’t conform to any standards of<br />

human beauty I know of – it’s a surprise to<br />

meet them in the flesh. They are about eight<br />

centimeters long and, well... sort of cute.


“Oh, they’re great,” Gary says. “You can watch<br />

them all day. Here, want to hold one?”<br />

They dart around in the cage fast, somehow<br />

reminding me of hyperactive lizards because<br />

of their thin legs and their odd way of dashing<br />

forward, stopping, and zipping off backwards<br />

at the same pace. He reaches in and grabs<br />

one, picking it up unceremoniously by the tail<br />

and letting it scrabble up and down his<br />

sweater. When he puts it on my hand it darts<br />

for an open sleeve. Gary grabs it by the tail<br />

again.<br />

“They’d make great pets,” he says. “People have<br />

considered it. But they’re not easy to take care<br />

of because of the temperature issue.”<br />

They would also never come when you call –<br />

being almost deaf. On the other hand, they<br />

would have no problem with my son’s stereo,<br />

which has been getting louder and louder<br />

over the past few months. He’s trying to make<br />

me enthusiastic about the new band he likes.<br />

Ironically, they’re called the Red Hot Chili<br />

Peppers.<br />

��<br />

Understanding why the rodents have lost<br />

their eyes and some of their pain<br />

requires a bit of time travel. It was almost<br />

exactly 150 years ago that a group of English<br />

scientists shook the water off their umbrellas<br />

and shed their soggy coats as they entered<br />

the Royal Geological Society in London. They<br />

were gathering for a meeting of the Linnean<br />

Society, the prestigious British science club,<br />

whose members included some of the world’s<br />

best naturalists. They met every month to<br />

read papers to each other and brag about<br />

new species of beetles they had discovered<br />

under moldy logs. Few of them had any idea<br />

of what they were about to witness at the<br />

meeting held on July 1, 1858. It took place in<br />

the Burlington house on Picadilly, in an ornate<br />

upstairs room whose walls and high ceiling<br />

were decorated with gilded carvings. Charles<br />

Darwin was not there – he had retired with<br />

his family to mourn the death of a young son<br />

– and Alfred Russel Wallace was chasing the<br />

“bird of paradise” on the opposite side of the<br />

globe, hoping to save himself from financial<br />

ruin by selling it to some rich British collector.<br />

In their absence, friends read their papers to<br />

the Linnean Society, and that is how evolution<br />

was announced to the world.<br />

One of the things said that night was that<br />

animals don’t acquire new hereditary traits<br />

because they need them, or lose features like<br />

eyes because they live in the dark. The naked<br />

mole rat has not become blind because it<br />

spends most of its life in an underground burrow.<br />

Instead, in a cramped world occupied by<br />

hundreds of expert diggers with sharp claws,<br />

packed close together, eyes are easy targets<br />

for scratches and infections. Ears are also likely<br />

sources of infection, more trouble than they<br />

are worth in the underworld. An animal in<br />

which the organs had disappeared or become<br />

covered over with skin would probably have<br />

better chances of surviving to pass along its<br />

genes.<br />

The lifestyle would also confer benefits, Gary<br />

Lewin says, to an animal that had lost its sensitivity<br />

to certain types of pain. “The problem<br />

was the air,” he says. “When so much carbon<br />

dioxide builds up in the environment, the<br />

body becomes very acidic. Acid is painful to<br />

every other known species of mammals, fish,<br />

and birds. But living the way the naked mole<br />

rats do, there would be strong evolutionary<br />

pressure to tune down or even eliminate the<br />

systems that cause such pain.”<br />

The same mechanisms that made the naked<br />

mole rat insensitive to acid might reduce<br />

other types of pain. “There are two main types<br />

of pain associated with inflammations,” Gary<br />

says. “You know how it is if you get an infected<br />

finger that gets red and swells up? That<br />

finger becomes extremely sensitive to pressure<br />

and heat, much more than the other fingers.<br />

That’s missing in the naked mole rat. It<br />

feels pressure and the pain of being pinched.<br />

But heat doesn’t hurt, and it doesn’t become<br />

hypersensitive.”<br />

Finding out why hasn’t been easy. There are a<br />

lot of steps between the skin, where pain originates,<br />

and the brain, where it is registered.<br />

Gary’s research shows that evolution has tinkered<br />

with the system at several levels.<br />

��<br />

Pain begins when wounded cells release<br />

molecules that dock onto neighboring<br />

neurons. This usually opens channels in their<br />

membranes that let in charged atoms like calcium<br />

or sodium. The result is a change in the<br />

balance of charges between the inside and<br />

the outside of the cell, turning the neuron’s<br />

membrane into a fantastic conductor of electricity.<br />

An electrical impulse shoots along the<br />

nerve, sometimes traveling fantastic distances<br />

to the other end of the cell. When they<br />

reach the synapse, a small gap that separates<br />

the nerve from its neighbor, the cell releases<br />

small molecules called neurotransmitters. By<br />

docking onto the next nerve, they cause it to<br />

open and close its own channels, allowing it<br />

to forward the message on to the brain.<br />

Cells that receive pain are called nociceptors.<br />

The main body of each cell, the part containing<br />

the nucleus, lives near the spinal cord in a<br />

cluster of nerves called the dorsal route ganglion,<br />

or DRG. It is wired into the nervous system<br />

in two directions. It extends an axon into<br />

the spine, where it transmits signals to nerves<br />

that carry them on to the brain. Another<br />

extension grows huge distances to the surface<br />

of the skin – sometimes a meter or more<br />

153 Interlude: These rats don’t dance to the Red Hot Chili Peppers


to reach a person’s toes. (The nervous system<br />

is built the same way in a giraffe, where single<br />

nerves can be over four meters long.) At the<br />

tips of these cells are receivers which sense<br />

the molecules that cause pain.<br />

Most of what is known about pain in the<br />

naked mole rat has come from comparing this<br />

system to a more traditional laboratory model<br />

– the mouse. Gary’s group has been equally<br />

successful at uncovering new sensory mechanisms<br />

in this other small mammal – mechanisms<br />

that are usually just as important in<br />

humans. At the end of 2007 a collaboration<br />

with the lab of Alistair Garratt turned up an<br />

important mechanism behind the pain detection<br />

systems in mice and other “normal”<br />

mammals.<br />

Alistair is a British scientist who came to the<br />

<strong>MDC</strong> in 1996, as a postdoctoral fellow with<br />

Carmen Birchmeier. “I had studied in<br />

Cambridge and was looking to leave the UK. I<br />

happened to see an advertisement that<br />

Carmen had published in the journal Nature –<br />

she was moving from Cologne to the <strong>MDC</strong><br />

and was looking for people to join the new<br />

lab.” Coincidentally, Alistair had studied<br />

German – inspired by a Swiss girl he used to<br />

Alistair Garratt<br />

take dancing lessons with. And Berlin seemed<br />

like a good place to be.<br />

Alistair has known he wanted to be a scientist<br />

for a long time. One of his first memories is of<br />

an experiment done by his father, a chemistry<br />

and biology teacher. “He combined lithium<br />

and water – it fizzed and caught fire,” Alistair<br />

says. “I couldn’t have been more than three<br />

years old, but I couldn’t take my eyes off it.”<br />

That first spark of interest was nurtured<br />

through his childhood by, among other<br />

things, television programs. The Open<br />

University of London broadcast lectures about<br />

science. After graduating from high school,<br />

Alistair went on to study science at the<br />

University of Cambridge, then got his PhD in<br />

Manchester.<br />

In 2002, after several very successful years<br />

working with Carmen, Alistair received one of<br />

the much sought-after Helmholtz fellowships.<br />

This allowed him to set up a small, independent<br />

group in the <strong>MDC</strong>’s program on<br />

Functions and Dysfunctions of the Nervous<br />

System. Part of independence meant finding<br />

new projects to work on. He had been interested<br />

in the body’s pain circuitry and talked to<br />

Gary Lewin and Carmen about how to find<br />

new genes involved in the sensory system.<br />

One possibility was to use DNA chips<br />

(described in the story “Stem cells in a jar.”)<br />

The technology would compare the genes<br />

active in nociceptors to those in other nerves,<br />

hopefully revealing those that were important<br />

in pain sensitivity.<br />

There were a few standouts. When Alistair<br />

brought the list along one day to lunch, Gary<br />

immediately recognized one of them: a receptor<br />

protein called c-Kit. “There’s a story behind<br />

c-Kit – it actually has to do with a cat,” Alistair<br />

says. “A cat from the state of Missouri in the<br />

U.S. – someone’s pet – that had cancer. This is<br />

one of those forms of cancer caused by a<br />

virus. Among other things, the virus brings in<br />

its own, defective form of the Kit gene. That<br />

pushes out the cell’s version of Kit, which<br />

causes problems, because the virus molecule<br />

can’t understand signals that the cell needs<br />

to receive.”<br />

c-Kit especially interested Alistair because it<br />

closely resembled other receptors in nerve<br />

cells that he had worked on; it might function<br />

the same way. Gary was interested because it<br />

might be another link in the pain network. In


1993, as a postdoc at Stony Brook University in<br />

New York, he had discovered that nociceptors<br />

had to be stimulated by a protein called NGF<br />

(which stands for nerve growth factor) to survive<br />

as the embryo developed. Later in life NGF<br />

switched functions and seemed to be<br />

involved in making animals hypersensitive to<br />

pain. This condition, called hyperalgesia, is<br />

what happens during inflammations.<br />

Gary discovered that there were two very similar<br />

types of nerves in the dorsal root ganglion.<br />

NGF was one of the few differences:<br />

only half of the nerves produce it. These<br />

seemed to be the same nerves that transmit<br />

hyperalgesic sensations. Identifying the molecules<br />

and pathways that make such similar<br />

nerves behave differently has been high on<br />

Gary’s research agenda ever since.<br />

Alistair suggested that one of his PhD students,<br />

Christina Frahm, take on the project. She was<br />

working on a protein that was a lot like c-Kit,<br />

and it would make a good backup in case her<br />

findings weren’t important enough to write up<br />

as a dissertation. Gary also put one of his students,<br />

Nevena Milenkovic, on the project. One<br />

of the first things the lab needed was a strain<br />

of mouse that didn’t produce c-Kit.<br />

“We could have made the knockout ourselves,”<br />

Alistair says. “But that takes time, and<br />

strains already exist. They don’t survive too<br />

well, because signals from c-Kit have several<br />

different functions. The molecule that activates<br />

c-Kit is called SCF, for stem cell factor,<br />

and you can imagine from the name that it<br />

has important jobs. For example, SCF triggers<br />

c-Kit to guide the development of some types<br />

of blood cells. Animals that don’t have it<br />

become quite anemic and usually die by the<br />

time of birth or very shortly thereafter.”<br />

But Max Gassman’s group in Zürich had been<br />

working with these mice and found a way to<br />

keep them alive. The animals could be “doped”<br />

with a molecule called erythopoeitin, a hormone<br />

that stimulates the production of new<br />

red blood cells. The treatment would disqualify<br />

the mice for the Tour de France, but it<br />

extended their lifespans. Alistair had recently<br />

invited Gassman to the <strong>MDC</strong> to give a talk;<br />

now he contacted the researcher again to ask<br />

to borrow some mice. Gassman was willing to<br />

help out.<br />

“Sometimes the animals that we use are<br />

more international than the scientists,”<br />

Alistair says. “The mice we got came from the<br />

University of Ulm, who had obtained them<br />

from a group in Japan, who had originally<br />

received them from the Jackson Laboratories<br />

in Maine.”<br />

��<br />

One job of SCF, the trigger for the c-Kit<br />

receptor, is to help guide the development<br />

of the skin. So the protein is produced by<br />

cells in the skin which lie right alongside the<br />

nerves that sense pain. Everything seemed to<br />

suggest that the two molecules were involved<br />

in pain transmission, so Christina put<br />

Gassman’s mice to the test. She found that<br />

they didn’t respond strongly to heat pain, but<br />

were oversensitive to pressure.<br />

The next step was to have a look at the nerves<br />

themselves. They were structured the same<br />

way in mice with and without c-Kit, so the loss<br />

of the molecule didn’t seem to be causing any<br />

rewiring of the nervous system. By staining<br />

cells in normal mice, Christina discovered that<br />

the nerves which produced c-Kit were the<br />

same as those that required NGF signals. And<br />

those cells also bear a molecule called TRPV1 –<br />

a channel protein that lets charged atoms<br />

into a nerve. This is what creates the impulse<br />

that turns the cell into a pain transmitter.<br />

Gary had been working on TRPV1 for a long<br />

time. Without the molecule, mice aren’t bothered<br />

by heat or capsaicin, the substance from<br />

chili peppers, and they don’t become hypersensitive<br />

to pain.<br />

Nevena and Christina examined the animals<br />

and discovered that they didn’t seem to feel<br />

heat pain caused by inflammations. On the<br />

other hand, they were more sensitive to pressure<br />

caused by inflammations than normal<br />

mice. These effects might be due to any of a<br />

number of things. The loss of c-Kit might<br />

change the types of nerves that developed or<br />

cause other structural changes to the nervous<br />

system. But dissections under the microscope<br />

revealed that this was not the case. The<br />

nerves that transmit pain are called C-fibers.<br />

There seemed to be the same number in the<br />

mutant mice, and they seemed to be wired<br />

into the nervous system the same way.<br />

“If those things were true, what we needed to<br />

check was whether the nerves were able to<br />

155 Interlude: These rats don’t dance to the Red Hot Chili Peppers


transmit signals at the same strength,” Gary<br />

says. One of his group’s specialties is the ability<br />

to extract nerves and test how strongly<br />

they conduct particular types of impulses, like<br />

determining the strength of a battery by<br />

measuring the current that passes through a<br />

wire. When Nevena performed this experiment,<br />

she discovered that heat signals in mice<br />

without c-Kit were much weaker than those<br />

in normal mice.<br />

The data suggested that the scientists were<br />

dealing with two types of pain receivers. Some<br />

responded very strongly to heat and others<br />

didn’t. The behavior of the mild responders<br />

didn’t change much without c-Kit, but there<br />

was a big change in the strong responders.<br />

“This offered a possible explanation for the<br />

difference between mild reactions to heat<br />

and the strong pain that is felt in hyperalgesia,”<br />

Gary says. “There are two types of C-fiber<br />

nerves that transmit heat information. Both<br />

are activated the same basic way, but one<br />

responds much more strongly. It’s a bit like<br />

having the same CD player hooked up to two<br />

sets of speakers. Both get the same signal. The<br />

cheap speakers never get very loud, but on the<br />

expensive set, turning up the sound just a little<br />

bit boosts the noise so high that it keeps<br />

the neighbors awake at night.”<br />

Was c-Kit turning on an amplification system<br />

in the second set of nerve fibers? One way to<br />

test the hypothesis was to shut down c-Kit in<br />

normal mice. Gary knew of a drug that could<br />

accomplish this. When it was administered to<br />

normal animals, they no longer experienced<br />

heightened pain from heat during infections.<br />

A second test took the opposite approach. SCF<br />

seemed to be passing the signal during an<br />

inflammation. What would happen in a<br />

mouse whose nerves detected high amounts<br />

of SCF even without an inflammation? When<br />

the experiment was performed, mild<br />

amounts of heat caused intense amounts of<br />

pain.<br />

The results raised more questions – for example,<br />

whether SCF and c-Kit directly provoked<br />

heightened pain. “We already knew that<br />

hyperalgesia depended on the TRVP1 receptor,<br />

the sensor for capsaicin. We had a strain of<br />

mouse that didn’t have the receptor, so we<br />

started to look for a connection.”<br />

The researchers repeated the experiments<br />

with mice that lacked TRVP1. If the receptor<br />

were lacking, the pain level didn’t rise, even<br />

with more SCF. This meant that the signal<br />

through c-Kit only worked if the receptor did<br />

its job.<br />

“The experiments have shown several new<br />

links in the chain that causes intense pain<br />

associated with heat,” Gary says. “In healthy<br />

animals and humans, those links are intact.<br />

Each one of them is a point at which it might<br />

be possible to use a drug to intervene and<br />

block the signals.”<br />

He is quick to point out that SCF and c-Kit are<br />

not the holy grail of pain research. “Removing<br />

the receptor greatly reduces sensitivity to<br />

heat, but it has another effect that seems<br />

contradictory,” he says. “Mice actually become<br />

Sören Markworth


much more sensitive to pain caused by pressure.<br />

This is equally interesting. It shows that<br />

parts of the same system regulate a second<br />

type of pain in a different way. We’re still<br />

working on how it does that.”<br />

��<br />

How much of what had been learned in<br />

mice could explain what was going on<br />

in the naked mole rat? Thomas Park and his<br />

colleagues in Gary’s group have discovered<br />

that evolution has worked at several levels to<br />

reduce the rodent’s sensitivity to pain.<br />

They are the only mammals known to lack<br />

two proteins called SP and CGRP in pain receptor<br />

nerves in the skin. Other labs had shown<br />

that if these molecules were removed from<br />

mice, the animals were much less sensitive to<br />

some types of pain. “This convinced us to take<br />

a systematic look at the behavior and biology<br />

of pain in the rats,” Gary says. “We discovered<br />

that they are the only mammals that don’t<br />

sense acid; they don’t become oversensitive to<br />

Cells stimulated by capsaicin, the<br />

ingredient that makes chili peppers<br />

hot, lie near the surface of the dorsal<br />

horn in the mouse. In the naked mole<br />

rat most of the nerves like deeper.<br />

Where the stimulation takes place<br />

may determine whether a sensation<br />

is painful or not.<br />

heat during inflammations, and they don’t<br />

respond to capsaicin.”<br />

But naked mole rats have TRPV1, the capsaicin<br />

receptor, and it works – the nerve that is stimulated<br />

by the substance receives and passes<br />

along a signal. The lab’s measurements<br />

showed that the strength of the impulse is<br />

just as strong as in mice. This meant that<br />

pain from the peppers – and from inflammations<br />

– was being interrupted somewhere<br />

higher in the network.<br />

This was different from the situation in birds.<br />

They also have TRPV1, but its architecture is<br />

somewhat different. The small changes<br />

means that capsaicin doesn’t dock on – the<br />

nerve isn’t told to feel pain, and it doesn’t generate<br />

an electrical impulse.<br />

So where does the rodent’s insensitivity come<br />

from? Thomas followed the signal to the<br />

spinal cord – maybe the sensory nerves<br />

weren’t wired up properly. Dissections<br />

showed that there was a difference in the way<br />

sensory nerves linked to their partners. In<br />

157<br />

These images show a cross section<br />

of the dorsal horn region of the<br />

spine. Synapses that bear the<br />

pain-transmitting molecule TRPV1<br />

are stained in red. Left: In the mouse,<br />

nearly all of these synapses from<br />

sensory neurons are found near the<br />

surface of the dorsal horn. Right:<br />

Cells producing TRPV1 are found<br />

much deeper in the naked mole rat.<br />

mice most nociceptor nerves – the pain<br />

receivers – lie near the surface of the spinal<br />

cord. In naked mole rats, most are much<br />

deeper.<br />

Previous studies have shown that when most<br />

mammals experience pain, the greatest stimulation<br />

of these nerves happens close to the<br />

surface. Gary hypothesizes that balance may<br />

be the crucial factor: if more signaling happens<br />

at the border of the spinal cord, information<br />

from the skin will be interpreted as pain.<br />

If the most activity is deeper inside, as with<br />

the naked mole rats, it will not.<br />

All of us would welcome a method to reduce<br />

pain from inflammations and other sources.<br />

These studies reveal several different junctions<br />

at which pain is perceived and passed<br />

along; any of them might make good places<br />

to interrupt its transmission. Understanding<br />

why the naked mole rat doesn’t hurt, Gary<br />

thinks, may reveal why the rest of us do.<br />

Interlude: These rats don’t dance to the Red Hot Chili Peppers


Part Three:<br />

Frontiers and ferrymen


The Swiss Army knife<br />

of the cell<br />

My personal favorite is the CyberTool 34, with the “translucent sap-<br />

Part Three: Frontiers and ferrymen<br />

phire” handle. It is outfitted with two knife blades (small and<br />

large), and a corkscrew, can and bottle openers, screwdrivers, a wire strip-<br />

per, a reamer, toothpick, sewing eye, bit wrench/case, tweezers, nine sock-<br />

ets and bits, a ball point pen, pliers, a wire cutter/crimper, scissors, hook,<br />

160


Thomas Willnow<br />

mini screwdriver, straight pin, and key ring.<br />

What else could you possibly need?<br />

Admittedly, the laser pointer and flash memory<br />

stick, which come with the “Swiss memory”<br />

knife, might be useful.<br />

The tool has come a long way since Karl<br />

Elsener made the first one in 1891; he was<br />

upset that his country’s army had outsourced<br />

its knives to Germany. His original model<br />

came with two screwdrivers, a can opener and<br />

an awl. (Officers were given an enhanced version<br />

with a corkscrew.) Today it has become a<br />

standard part of the repertoire on expeditions<br />

to Mount Everest and the space shuttle,<br />

although you have to wonder how astronauts<br />

open the knife while wearing those thick<br />

gloves. Art historians apparently find them<br />

useful because they’re standard equipment at<br />

the Museum of Modern Art in New York City.<br />

And then there is the case of John Ross, doctor<br />

in Uganda, who had to use one to perform six<br />

emergency amputations when someone stole<br />

his surgical saw.<br />

The Wenger company, which shares the manufacture<br />

of Swiss Army knives with Elsener’s<br />

company Victorinox, recently produced a collector’s<br />

edition called the Giant, which is all<br />

versions wrapped into one. It contains all 85<br />

tools ever found in a Swiss army knife, including<br />

instruments to clean the cleats of golfers’<br />

shoes and the heads of their clubs. It’s 22 centimeters<br />

thick, weighs nearly a kilogram, and<br />

costs about a thousand Euros. You’d need a<br />

wheelbarrow to carry one around. I think I’ll<br />

stick with the CyberTool.<br />

161 Part Three: Frontiers and ferrymen<br />

��<br />

As large as the Giant is, it’s a fraction of<br />

the size it would be if Karl Elsener hadn’t<br />

invented a special spring mechanism that<br />

allowed him to pack tools into both sides and<br />

both ends of the knife. That innovation is<br />

what allows a soldier (or an art historian) to<br />

walk around with one knife instead of five or<br />

six. And it’s one reason that Thomas Willnow,<br />

researcher at the <strong>MDC</strong> and professor at the<br />

Charité, sees the Swiss army knife as a<br />

metaphor for his favorite family of proteins.<br />

The molecules Thomas has been working on<br />

are related to a protein called the low-density<br />

lipoprotein receptor. So they are called lowdensity<br />

lipoprotein receptor related proteins<br />

(LRPs). It’s a daunting name, Thomas admits.<br />

One of the functions of the low-density<br />

lipoprotein receptor and its LRP relatives is to<br />

regulate the amounts of cholesterol and<br />

other lipids (fats) in the bloodstream by drawing<br />

them into cells.<br />

“Lipids are the major component of cell membranes<br />

and are used for many other things,”<br />

Thomas says. “Cholesterol, for example, is<br />

modified to make vitamin D and several types<br />

of hormones, including the sex hormones<br />

estrogen and testosterone. It’s also used as a<br />

glue to hold molecules in the membrane<br />

together. And you know of its potential negative<br />

effects: if your cells don’t remove it from<br />

the bloodstream, it builds up in the linings of<br />

blood vessels and causes a clogging of your<br />

arteries and many other problems. So lipid


The LDL receptors (left) and Sortillin gene families have many members.<br />

Evolution has modified their structures by adding on and removing<br />

modules, giving the proteins a range of functions in many different tissues.<br />

and cholesterol metabolism are fundamental<br />

to a wide range of biological processes.”<br />

Most of the body’s cholesterol is synthesized<br />

by cells in the liver and a few other organs;<br />

some is absorbed through food. It is usually<br />

transported from these sources to the rest of<br />

the body through the blood. Because fat does<br />

not mix with water, cholesterol and other<br />

lipids have to be wrapped into specialized<br />

cargo complexes for transport. The cargo<br />

complexes for fat are low-density lipoproteins<br />

(LDLs) or similar types of protein carriers.<br />

A single LDL can hold about 1500 cholesterol<br />

molecules. When the package arrives, it is recognized<br />

by a receptor on the target cells – the<br />

LDL receptor or an LRP. The cell engulfs the<br />

protein with its cargo, and carries it to internal<br />

compartments called endosomes, where it is<br />

unpacked. This process of drawing in molecules<br />

is called endocytosis.<br />

Part Three: Frontiers and ferrymen<br />

162<br />

“Endocytosis is the main way that cells control<br />

the entry of molecules they need,”<br />

Thomas says. “A receptor binds to an outside<br />

molecule and draws it inside. The cell builds a<br />

wide range of receptors that behave this way.<br />

One reason there are so many is that various<br />

tissues need to import different types of molecules,<br />

and specialized receptors help them<br />

do so. But we don’t yet have a detailed understanding<br />

of what many of these proteins do.”<br />

For instance, human cells, as well as those of<br />

other mammals, make nine types of LRPs. They<br />

are all spin-offs of a common ancestor, arising<br />

through evolution and taking on new functions.<br />

When Thomas arrived at the <strong>MDC</strong> 12<br />

years ago, the first LRPs just had been identified,<br />

but little was known about its functions.<br />

“Each of the receptors is able to bind to combinations<br />

of proteins and lipids such as cholesterol,”<br />

Thomas says, “but it wasn’t known<br />

where in the body and under what circumstances<br />

they did so. Like the LDL receptor, they<br />

may play an important role in recycling and<br />

using fats. That’s a crucial process throughout<br />

our bodies, and it is disturbed in a number of<br />

diseases. So we decided to take a systematic<br />

look at the whole family.”<br />

He has a diagram of the nine LRP molecules<br />

shown side-by-side. Each is lodged in the cell<br />

membrane, pushing small, root-like tails into<br />

the cell. Outside, they stretch away from the<br />

cell like beanstalks. The LDL receptor is the<br />

smallest of the family; some of its cousins<br />

have much larger external modules. But they<br />

have similar features, Thomas points out –<br />

evolution has removed some subunits,<br />

stretched others, and copied yet others, like<br />

adding new tools or changing the shapes of<br />

existing ones so they perform different jobs.<br />

Some regions of the molecules help them<br />

bind to the cargo. Others have roles to play<br />

while escorting their partners into the cell.<br />

One module lets them release a partner once<br />

it has been delivered inside to endosomes.<br />

It’s another parallel to the Swiss army knife,<br />

although evolution didn’t start with the sim-


plest form of the molecule and develop<br />

increasingly complex types. Instead, the oldest<br />

LRP seems to be a huge protein called<br />

megalin, which is more like the Giant. Parts of<br />

it have been removed or rearranged to make<br />

specialized tools for particular types of cells,<br />

the way there are specialized knives for mountain<br />

climbers and golfers. “That’s also how cell<br />

types evolved,” Thomas says. “Things started<br />

with a primitive unicellular organism that<br />

gradually developed into a multicellular<br />

organism with many specialized types of cells.<br />

“As our lab and other groups have knocked out<br />

the LRPs in mice, we discovered that they have<br />

many jobs in addition to binding combinations<br />

of lipids and proteins. That multi-functionality<br />

might seem strange because the<br />

LRPs are so similar. The versatility comes from<br />

the fact that each part of the protein can bind<br />

to various partners, with different effects on<br />

the cell. The external region might bind to<br />

lipid carriers, or signaling molecules, or a virus.<br />

Or the LRP might bind to another receptor in<br />

the cell’s membrane, changing the external<br />

partners they can dock onto. And the tail<br />

region of the molecule, inside the cell, can also<br />

bind to different proteins. Each combination<br />

has unique effects.”<br />

I know of one lab that has spent six years<br />

working on just the tail region of one receptor.<br />

Isn’t it a daunting task to take on a whole<br />

family of molecules?<br />

“It is,” he admits. “We couldn’t do it without<br />

the help of colleagues with expertise in various<br />

systems, many of whom are on campus.<br />

We’ve done most of our projects in collaboration<br />

with other groups. Everything we’ve<br />

learned tells us that to deeply understand one<br />

of these molecules, we’ll have to understand<br />

several of them, and their functions in various<br />

parts of the body. It’s easier to understand the<br />

CyberTool 34 if you’ve seen other types of<br />

Swiss army knives.”<br />

��<br />

163 Part Three: Frontiers and ferrymen


Thomas and his colleagues have found<br />

that some of the molecules have an<br />

impressive range of jobs. Megalin is a good<br />

example. High numbers of the receptor are<br />

found in the proximal tubules, a crooked<br />

pipeline in the kidney. Since this region of the<br />

organ is responsible for capturing sugar,<br />

water, amino acids, and a range of proteins,<br />

then recycling them for use by the body, it<br />

made sense that an LRP should be found<br />

there. Exactly what the receptor recycled,<br />

however, was unknown.<br />

In 1996 the lab developed a strain of mouse<br />

without the molecule in hopes of finding out.<br />

They expected to see defects in the absorption<br />

of molecules in the kidney. That turned<br />

out to be the case, but first another issue had<br />

to be dealt with.<br />

“For some reason, without the protein the<br />

brains of the mice didn’t develop properly,”<br />

Thomas says. “The forebrain was so seriously<br />

malformed that most mice died within a few<br />

minutes of birth. This specific type of deformi-<br />

ty is called holoprosencephaly – in which the<br />

two hemispheres of the brain and the features<br />

of the face become fused. Sadly, this<br />

defect is quite common in humans. The most<br />

extreme form of this condition is cyclopism, in<br />

which an organism develops a single eye and<br />

often no nose.”<br />

A project carried out by Robert Spoelgen,<br />

Annette Hammes, and Uwe Anzenberger in<br />

Thomas’ lab has offered a possible explanation<br />

for the problem. The brain defects arise<br />

partly because of megalin’s ability to recognize<br />

combinations of proteins and cholesterol,<br />

and partly from its binding to other<br />

receptors in the membrane.<br />

“To make a proper brain with two hemispheres,<br />

cells in the early embryo have to<br />

receive a signal from a small protein called<br />

sonic hedgehog,” Thomas says. “Sonic hedgehog<br />

is secreted by cells. For some reason, cholesterol<br />

has to be attached to it for the successful<br />

transmission of a signal. The reason<br />

hasn’t been completely clear, but it makes<br />

sense that an LPR protein might be involved in<br />

Part Three: Frontiers and ferrymen<br />

164<br />

recognizing such a signal from a cholesterolmodified<br />

protein. After all, these are the<br />

favorite cargoes of the receptors. Robert discovered<br />

that the sonic hedgehog signal<br />

requires the presence of megalin to work in<br />

the embryonic brain. We think that megalin<br />

teams up with another molecule – perhaps<br />

the main receptor for sonic hedgehog, called<br />

patched – to allow cells to recognize and act<br />

on the signal.”<br />

The study showed where the molecules meet<br />

up – in a surface layer of the developing brain.<br />

Signals from sonic hedgehog help tell cells<br />

whether they are at the front or back of the


animal and where they should migrate.<br />

Another function of the signal is to tell some<br />

cells to die, which sculpts and separates layers<br />

of brain tissue. Without the signal, these<br />

processes don’t happen normally, and the<br />

brain doesn’t form two hemispheres.<br />

The capture of sonic hedgehog by the cell is<br />

only the first step in sending a message to<br />

cells’ genes, telling them to change their<br />

behavior. Megalin’s tail may pass the signal<br />

into the cell, and in 2004 Helle Petersen, a<br />

postdoc in Thomas’ group, figured out<br />

how that might happen. To do so she had to<br />

“fish” for proteins that might bind to this<br />

region of megalin by presenting it with cellular<br />

molecules and watching what it grabbed<br />

onto.<br />

One partner she found was a new protein she<br />

called megalin-binding protein or MegBP (for<br />

obvious reasons) “That was interesting<br />

because MegBP can interact with other regulators<br />

in the cells, transcription factors. We<br />

suspect that megalin has gene-activating<br />

functions, and Helle showed that this is probably<br />

the case. Its tail binds to MegBP. When<br />

megalin receives a signal, its tail module<br />

releases MegBP. The factor is now free to<br />

move to the nucleus, where it may bind to<br />

165 Part Three: Frontiers and ferrymen<br />

Kristin Kampf<br />

transcription factors and regulate activation<br />

of a specific set of genes.”<br />

��<br />

Not all animals are equally affected by<br />

the loss of megalin. In about five percent<br />

of the knockout mice, the effects on the<br />

brain are mild and the animals live to adulthood.<br />

The survivors gave the lab a way to<br />

study the receptor’s functions in the kidney.<br />

Work in the test tube had shown that megalin<br />

could bind to many molecules that are<br />

supposed to be absorbed by the proximal<br />

tubules. In 1999 PhD student Jörg-Robert


Salim Seyfried<br />

Leheste and other members of the group<br />

showed that megalin is the major receptor in<br />

this part of the kidney that draws in hormones,<br />

vitamins and other molecules bound<br />

to carrier proteins. Many of these molecules<br />

accidently slip through the filter of the kidney<br />

because they are small – but the body invented<br />

a backup mechanism to capture them<br />

before they are lost in the urine.<br />

That finding also suggested a link between<br />

megalin and disease. If megalin is defective,<br />

mice are unable to absorb crucial vitamins and<br />

proteins which then leave the body in the animal’s<br />

urine. Soon vitamins and hormones are<br />

depleted from the body. The consequence is<br />

severe vitamin deficiency. The same sequence<br />

of events is seen in patients suffering from a<br />

condition called Fanconi’s syndrome.<br />

“But how can you identify the true binding<br />

partners for megalin in the kidney in the first<br />

place if you have no clue what they could be?”<br />

Thomas asks. He draws a diagram in which he<br />

represents the kidney as a simple tube.<br />

“Substances enter on one side from the bloodstream<br />

and leave on the other in the urine. The<br />

receptors line the inside of the tube. Anything<br />

removed by the receptors doesn’t come out<br />

the other side. So we thought of a very simple<br />

experiment to do with our knockout mice. We<br />

compared their urine to that of healthy mice,<br />

expecting to find extra substances which<br />

would normally be removed by megalin. We<br />

discovered that vitamin D3 and its carrier were<br />

no longer being absorbed.”<br />

Vitamin D3 is a hormone which helps cells<br />

control their levels of calcium. “At the time it<br />

Part Three: Frontiers and ferrymen<br />

166<br />

was still thought that most hormones simply<br />

slip through the membrane on their own,”<br />

Thomas says. “Our experiments proved that<br />

vitamin D3 needs help. At first this finding was<br />

considered an exception.” Follow-up studies<br />

by another student, Regina Burmeister,<br />

showed that megalin was binding to a second<br />

receptor, called cubulin, to carry out some of<br />

its jobs in the kidney.<br />

One thing that might help clarify the receptors’<br />

functions would be to compare their<br />

behavior in the mouse and other animals.<br />

Thomas turned to Salim Seyfried, a junior<br />

group leader at the <strong>MDC</strong> who has introduced<br />

the zebrafish to the lab. Over the past 15 years,<br />

this tiny animal has become one of the most<br />

important laboratory organisms for the study<br />

of developmental processes. It is almost com-


pletely transparent, which allows scientists to<br />

watch organs form over time without having<br />

to dissect the animal. It is so small that its<br />

cells can feed themselves for a long time by<br />

absorbing nutrients directly from the environment.<br />

This means that the fish can survive<br />

even if they have defects in the heart or circulatory<br />

system. It is an ideal subject for Salim’s<br />

lab, which focuses on heart development, and<br />

needs to be able to study animals whose<br />

organs grow in a defective way.<br />

Uwe Anzenberger and colleagues from<br />

Salim’s and Thomas’ lab carried out a study to<br />

see how easy it would be to compare LRPs in<br />

fish and mice. They found that zebrafish have<br />

close relatives of megalin and cubulin, that<br />

the molecules are produced together in the<br />

fish’s kidney, and that they play an equally<br />

important role in drawing hormones into<br />

cells.<br />

“Many of the molecules that are captured by<br />

LRPs are delivered to internal compartments<br />

in cells called endosomes,” Thomas says. “Uwe<br />

found that without megalin, substances were<br />

not drawn into cells in the fish – and there<br />

weren’t any endosomes. So the system closely<br />

parallels what happens in mammals and<br />

will allow us to get a closer look at how the<br />

kidney works.”<br />

These varied studies have changed Thomas’<br />

mind about the textbook view that hormones<br />

simply slip into cells. That is surely true at<br />

times, because some of the tissues that need<br />

them do not produce LRPs. But where the<br />

receptors are present, the lab kept discovering<br />

exceptions. Studies in cell cultures revealed<br />

that megalin was also needed to take up sex<br />

hormones. That didn’t prove it also happened<br />

in animals, but it was likely to. Megalin is also<br />

found at high levels in the target cells for sex<br />

hormones.<br />

In 2005 this led postdoc Annette Hammes,<br />

collaborating with scientists from the<br />

University of Aarhus in Denmark, to take a<br />

new look at the megalin knockout mice. She<br />

discovered abnormalities in the development<br />

of the reproductive organs of both males and<br />

females, as if cells were no longer being stimulated<br />

by sex hormones. That didn’t fit with<br />

the hypothesis that hormones could simply<br />

slip through membranes; it only made sense<br />

if megalin was detecting them and escorting<br />

them inside.<br />

“The vast majority of cholesterol-derived hormones<br />

and lipids are bound to proteins or<br />

packaged in lipoproteins as they move<br />

through the bloodstream,” Thomas says. “In<br />

those cases, we think that their entry into the<br />

cell is managed by receptors, usually the LRPs.<br />

That hypothesis is giving us a handle on many<br />

diseases in which what is broken – and what<br />

needs to be fixed – is an endocytic receptor.<br />

So, for example, we’ve found that signals from<br />

another LRP called the ApoE receptor 2 are<br />

needed to prompt cells to develop into<br />

mature sperm. That signal never arrives when<br />

this LRP is defective. Infertility is the consequence.<br />

Seven percent of all human males<br />

have fertility problems. LRPs may well play an<br />

important part in this condition. Under -<br />

standing why the signal goes missing may<br />

one day help to restore it.”<br />

��<br />

The flat topography of Berlin and its spacious<br />

streets make the city ideal for bikers,<br />

if you’re willing to risk getting caught in<br />

the tempests that sometimes arise without<br />

warning in the fall and winter. One of them<br />

flashed over the campus in January 2007 with<br />

such ferocity that it stripped the bright red<br />

façade off the Max Delbrück Commu nications<br />

Center, or <strong>MDC</strong>C. The timing couldn’t have<br />

been worse – in less than 24 hours the campus<br />

would hold its New Year’s party – in the<br />

<strong>MDC</strong>C. At two a.m. residents of the campus<br />

guest houses had heard what sounded like an<br />

explosion. When security personnel came to<br />

investigate, they were greeted by a sight so<br />

alarming that they immediately called<br />

Gudrun Erzgräber and told her that the entire<br />

roof had collapsed. She called Walter<br />

Birchmeier.


Ulrike Förster<br />

By three a.m. Walter was surveying the damage<br />

with Stefan Schwartze, the <strong>MDC</strong>’s administrative<br />

director. Things were bad enough, but<br />

not as bad as they looked. The roof hadn’t fallen<br />

in. What appeared to be stone rubble littering<br />

the ground – and the bust of Max<br />

Delbrück, at the front door – was actually a<br />

layer of light, foam-like blocks that had<br />

covered the building. Workers swept the<br />

debris out of the way and the New Year’s celebration<br />

went on as planned, with a talk by<br />

Nobel laureate Aaron Ciechanover and a<br />

per formance by jazz pianist Rich Bairach to<br />

take our minds off the roof. The <strong>MDC</strong>C spent<br />

most of the year covered in plastic and scaffolding.<br />

Thomas is one of those who braves the Berlin<br />

weather; his family doesn’t even own a car.<br />

Every day he rides his bike from home to the<br />

S-Bahn, catches a ride to Buch, then gets back<br />

on the bike for the last bit to campus. We met<br />

once at the S-Bahn station in Pankow, where<br />

there is a shop that sells Swiss army knives.<br />

The front window displays a larger-than-life<br />

knife, even larger than the Giant, with rotating<br />

blades.<br />

“I’ve been using this metaphor for quite a few<br />

years, and the more we learn the better it fits,”<br />

he says. “LRPs can do so many things because<br />

of their complexity – single molecules contain<br />

several tools – and their ability to link up with<br />

other tools. Besides proteins bound to lipids,<br />

these molecules recognize viruses, antibiotics,<br />

toxins, signaling molecules, and a range of<br />

other partners. The result may be to draw the<br />

partner into the cell, or to allow an influx of<br />

calcium, or trigger cell migration, or change<br />

the behavior of synapses... I could go on, but<br />

you get the point.”<br />

Thomas is willing to follow the LRPs wherever<br />

they lead him. “Sometimes it is helpful not to<br />

be preoccupied with a certain theory but to<br />

watch what nature tells you,” he says. It’s one<br />

of the advantages of being at the <strong>MDC</strong>. “If you<br />

knock out a molecule and it leads to a problem<br />

in the brain, we can find partners in the<br />

neurobiology groups to help us follow up. Our<br />

own expertise lies in the cardiovascular<br />

system, but these systems are controlled<br />

by fundamental processes which occur<br />

throughout the body. So the processes and<br />

the molecules involved in them have to<br />

be looked at in a more holistic way. You have<br />

Part Three: Frontiers and ferrymen<br />

168<br />

to look at many facets of a molecule to<br />

truly understand it. You have to explore how<br />

it has changed through evolution by observing<br />

its functions in different model organisms.<br />

“Everyone is talking about ‘systems biology’<br />

these days. Well, hormone signaling is a good<br />

example of a ‘systems’ problem. Hormones<br />

are made and released by individual cells, and<br />

taken in by individual cells, but in a way that<br />

has to be coordinated all over the body. It’s<br />

what raises blood pressure when there has<br />

been an injury and blood has been lost, for<br />

example. Hormone signals change the behavior<br />

of blood vessels, the heart, and a number<br />

of other bodily systems. So it’s the type of<br />

problem that you have to try to approach in a<br />

global way.”<br />

Recently the lab has begun studying a second<br />

family of proteins, the sortilins. Although the<br />

members of this family evolved independently<br />

from LRPs, they have a similar structure.<br />

“Many regions of the brain don’t produce<br />

LRPs, and they don’t have access to the bloodstream<br />

where the fats circulate, but they need<br />

cholesterol and other lipids,” he says. “So


there’s been an idea that sortilins might act<br />

as lipid receptors as well.”<br />

You never know what will happen when you<br />

open a new biological box. The lab’s studies of<br />

a sortilin called sorLA brought them faceto-face<br />

with an entirely new problem: the<br />

development of Alzheimer’s disease. If there’s<br />

a biological equivalent of the roof falling in,<br />

that’s it.<br />

��<br />

In 1901 a 51-year-old woman named<br />

Auguste Deter was admitted to the State<br />

Hospital for Epileptics and the Mentally Ill in<br />

Frankfurt. At home her behavior had become<br />

progressively more bizarre. She had paranoid<br />

fantasies in which she accused her husband<br />

of having affairs; she was losing her ability to<br />

read, write, and reason. Finally the family was<br />

unable to cope and admitted her to the<br />

hospital.<br />

She was examined by a 37-year-old neurologist<br />

named Alois Alzheimer. He conducted a<br />

medical and behavioral examination. His case<br />

notes, which were found in the archives of the<br />

University of Frankfurt in 1995, report:<br />

She sits on the bed with a helpless expression.<br />

What is your name? Auguste. What is<br />

your husband’s name? Auguste. Your husband?<br />

Ah, my husband. She looks as if she<br />

didn’t understand the question. Are you<br />

married? To Auguste. Mrs D? Yes, yes,<br />

Auguste D. How long have you been here?<br />

She seems to be trying to remember. Three<br />

weeks. What is this? I show her a pencil. A<br />

pen. A purse, key, diary and cigar are<br />

identified correctly. At lunch she eats cauliflower<br />

and pork. Asked what she is eating<br />

she answers spinach. While eating meat she<br />

answers potatoes and horseradish. After a<br />

short time she forgets objects that have<br />

been shown to her. She constantly speaks of<br />

twins. When she is asked to write, she<br />

holds the book in such a way that one has<br />

the impression that she has a loss in the<br />

right visual field. Asked to write Auguste<br />

D., she tries to write Mrs and forgets the<br />

rest. It is necessary to repeat every word...<br />

By the time Deter died in Frankfurt in 1906,<br />

Alzheimer had moved on to Munich, to a posi-<br />

169 Part Three: Frontiers and ferrymen<br />

tion at the Royal Psychiatric Clinic. He had not<br />

forgotten the case, and when he learned of<br />

her death, he wrote to the Frankfurt hospital<br />

to ask permission to examine her brain.<br />

Alzheimer had become one of Germany’s foremost<br />

experts in the pathology of the nervous<br />

system, and the hospital director granted the<br />

request.<br />

Alzheimer discovered bundles of “a peculiar<br />

material” in the cortex of the brain. These<br />

dense clusters of fats and proteins are called<br />

amyloid plaques. In Alzheimer’s patients they<br />

collect between neurons, blocking their communication<br />

with each other and causing the<br />

cells to die. Over time, this leads to the symptoms<br />

of Alzheimer’s disease: a loss of cognitive<br />

functions and a person’s control of his or<br />

her body, and finally death.<br />

In the 1980s, Konrad Beyreuther’s lab at the<br />

University of Heidelberg identified the main<br />

component of the plaques: a small peptide<br />

called β-amyloid. This turned out to be a fragment<br />

of a much larger molecule called APP, for<br />

amyloid precursor protein. APP is a membrane<br />

protein found on the surface of everyone’s<br />

neurons. Although its functions in the healthy


ain are not well understood, it seems to play<br />

an important role in memory and learning.<br />

For most of people’s lives they produce β-amyloid<br />

fragments without causing much harm.<br />

However, by old age so much of it has accumulated<br />

in the brain that it forms the conspicuous<br />

deposits – first encountered by Alois<br />

Alzheimer in his autopsy studies. Similar to the<br />

build-up of “debris” in the vessels in cardiovascular<br />

disease, the built of these plaques in the<br />

brain harms the nerve cells, causing mental<br />

problems typical of Alzheimer’s disease. If scientists<br />

could find out why this happens, they<br />

might be able to interrupt the process and<br />

block the development of the disease.<br />

Could it be a coincidence that diseases of<br />

blood vessels and diseases of the brain share<br />

so much similarity? Thomas suspects a connection,<br />

and the connection may be receptors<br />

for lipid-loaded lipoproteins. That’s why his<br />

group was investigating sorLA, one member<br />

of the sortilin family found in neurons, when<br />

he read a study by James Lah’s laboratory at<br />

Emory University in Atlanta, USA. Lah and his<br />

colleagues had conducted a survey of the<br />

molecules produced in the brains of<br />

Alzheimer’s patients, looking for molecules<br />

that behaved differently than in healthy peo-<br />

ple. They found that people with the disease<br />

had abnormally low levels of sorLA.<br />

“This happened at a time when a flurry of discoveries<br />

had been made connecting<br />

Alzheimer’s disease and cholesterol,” Thomas<br />

says. “A study by Elizabeth Corder from Duke<br />

University in the US showed that sequence<br />

variations in the protein ApoE were the highest<br />

single predictor of a person’s developing<br />

Alzheimer’s disease. I’ve already told you<br />

about our work on ApoE – it’s a lipid-carrying<br />

protein that is recognized by LRPs – and by<br />

sortilins as well. So finding that some forms of<br />

ApoE are highly likely to lead to Alzheimer’s is<br />

one link between LRPs and the disease.”<br />

An important piece of evidence for the<br />

hypothesis came from the discovery of a<br />

direct link between the receptor LRP1 and APP.<br />

Brad Hyman’s lab in Harvard and others discovered<br />

that a third protein was able to grab<br />

both molecules, gluing them together.<br />

“When this happens on the cell surface, the<br />

two molecules are drawn into the cell and<br />

delivered to endosomes,” Thomas says. “These<br />

compartments contain the molecule that<br />

slices up APP into the β-amyloid fragment. If<br />

APP isn’t delivered, it won’t be cut that way.”<br />

Part Three: Frontiers and ferrymen<br />

170<br />

What prevents LRP1 and APP from coming<br />

together in every cell? I ask him.<br />

“Part of it seems to have to do with another<br />

LRP receptor called LRP1B,” Thomas says.<br />

“There are only so many copies of APP on the<br />

surface of the cell, and other molecules compete<br />

with each other to bind to them. LRP1B is<br />

drawn into the cell at a much slower rate than<br />

LRP1. So if it binds to APP, it holds it on the cell<br />

surface, where it eventually gets processed in<br />

other ways, releasing ‘healthy’ fragments.”<br />

SorLA comes in, he says, during the process of<br />

delivery. “SorLA has the classic features of an<br />

LRP. The main difference between it and other<br />

LRPs is an extra module. This unit acts as a<br />

sort of global positioning system that directs<br />

the molecule toward specific addresses within<br />

the cell. It’s a bit unusual for an LRP because<br />

it operates mainly inside the cell, rather than<br />

on its surface.”<br />

Olav Andersen, a postdoc in Thomas’ lab from<br />

Denmark, worked on the connection between<br />

sorLA and Alzheimer’s disease for about five<br />

years. “What we wanted to understand was<br />

the ‘mechanical’ connection between the<br />

activity of the two molecules, sorLA and APP,”<br />

Thomas says. “We needed to prove that they<br />

actually met in the cell and could influence<br />

each other’s activity.”<br />

Olav came to the lab along the same route as<br />

several other talented young scientists;<br />

Thomas recruited him from the University of<br />

Aarhus, where he has long-standing collaborations.<br />

“It’s a good partnership because<br />

Anders Nykjaer’s group there is strongly<br />

focused on the biochemistry and cell biology<br />

of these proteins,” he says. “We concentrate<br />

more on the physiology side. So if their postdocs<br />

start getting interested in looking at<br />

organs and animals, I can attract them here.”<br />

Thomas says that he had to get adjusted to<br />

Olav’s work habits. “He left at three or so in<br />

the afternoon,” Thomas says. “I’d go into the<br />

lab and the only thing there would be his<br />

Birkenstock sandals and an empty bench.


Scientists have flexible work hours, but I<br />

thought leaving at three was a bit extreme.<br />

Then I found out he was coming into the lab<br />

at five in the morning. He’s always done that.<br />

It’s just the way he works.”<br />

Olav’s first experiments proved that the modules<br />

of sorLA and APP lying outside the cell<br />

could bind to each other. Next he tagged the<br />

molecules with fluorescent markers and<br />

tracked their movement through the cell<br />

under the microscope. The proteins seemed to<br />

be moving together to certain internal compartments.<br />

An additional experiment gave<br />

cells a mutant form of sorLA which sat in the<br />

cell membrane but could stay in the cell. This<br />

left APP stranded on the surface as well.<br />

The opposite approach – giving healthy cells<br />

much higher amounts of sorLA than normal –<br />

had an equally interesting outcome. “APP is<br />

kept in the cell along with the receptor and<br />

gets trapped in large compartments,” Thomas<br />

says. “But this time it is a different compartment<br />

than the endosome. It doesn’t get<br />

chopped into the harmful β-amyloid fragment.<br />

So sorLA has an interesting job. It delivers<br />

APP to places in the cell where it is locked<br />

up and remains harmless. It is no longer delivered<br />

to the cell membrane to meet the culprit<br />

LRP1.<br />

“That’s what happens in ‘normal’ neurons, at<br />

least as they get old, and it happens more<br />

often when there is not enough sorLA. This<br />

made us think that the cause of the accumulation<br />

of β-amyloid fragments might be a<br />

drop in the production of sorLA in patients.”<br />

The scientists put the hypothesis to the test<br />

by examining tissue samples from ten<br />

Alzheimer’s patients. All of the patients – and<br />

none of the healthy subjects – showed low<br />

levels of sorLA in the brain.<br />

The findings led Thomas’ lab to develop a<br />

strain of mouse without sorLA. If the hypothesis<br />

was correct, the mice’s APP ought to<br />

behave like that of Alzheimer’s patients. By<br />

ten months of age, the animals’ brains experi-<br />

171


Daniel Militz<br />

Part Three: Frontiers and ferrymen<br />

enced a 30 percent increase in levels of the dangerous beta-amyloid<br />

fragment, and they developed those notorious plaques.<br />

The way APP is shuttled through the cell directly influences whether it<br />

is processed into the harmful fragments that have been linked to<br />

Alzheimer’s disease. Several molecules are involved in this process, and<br />

pharmaceutical companies are attempting to find drugs that can alter<br />

their activity. “SorLA looks like it would make a good target,” Thomas<br />

says. “A drug that increases its activity could have the same effects in<br />

patients we have observed in the lab: it might lock up APP and prevent<br />

the formation of amyloid fragments and plaques.”<br />

��<br />

Another Danish alumnus of the Willnow group, Anders Nykjaer, is a<br />

professor at the University of Aarhus. He spent almost two years<br />

on a sabbatical in Berlin; while there he introduced the lab to sortilins.<br />

Like many former Danish colleagues who spend time in Berlin, he continues<br />

to collaborate with Thomas on themes started at the <strong>MDC</strong>. In<br />

2006 a collaboration between the two labs and Weill Cornell Medical<br />

College in New York City revealed new functions for sortilin, a protein<br />

closely related to the sorLA receptor. They found that sortilin binds to<br />

another membrane receptor called p75NTR . When that happens, the two<br />

proteins build a platform that can dock onto a signaling molecule<br />

called proNGF. The combination of the three molecules triggers a selfdestruct<br />

program in the cell. This signal likely participates in the widespread<br />

death of cells that follows strokes or other types of brain injury.<br />

Blocking this signal prevents the cells from dying, as the scientists have<br />

shown by knocking out sortilin in mice.<br />

Megalin, sorLA, and sortilin all look like promising leads in the search<br />

for drugs to curb diseases. A few years ago Thomas and Anders founded<br />

a company called ReceptIcon to bring the studies to the point that<br />

pharmaceutical companies might step in.<br />

“Megalin is particularly interesting because it plays a role in the toxic<br />

side-effects that arise from antibiotics and a huge range of other therapeutic<br />

drugs,” Thomas says. “As the receptor filters vitamins passing<br />

through the proximule tubules of the kidney, it sometimes also draws<br />

in substances it should not – such as therapeutic drugs. It does so<br />

because the receptor accidentally confuses them with its true binding<br />

partners. Thus, instead of being excreted, these drugs build up in the<br />

kidney and cause organ failure and death.”<br />

One of ReceptIcon’s projects has been to help find substances that will<br />

block megalin’s activity – at least temporarily – until drugs have passed<br />

through the kidney. One such inhibitor has now passed animal trials<br />

and is scheduled for clinical trials in humans next year.<br />

It’s gratifying, Thomas says, to make a discovery with clear potential<br />

applications and health benefits. “This is a real example of molecular<br />

172


Anne-Sophie Carlo Chandresh Gayera<br />

medicine,” he says. “It happened because we<br />

found a molecule which plays a direct role in<br />

processes of disease and organ damage. But<br />

in many cases, for example the sortilin story, a<br />

solution will probably have to be more subtle.<br />

We have a good track record of uncovering<br />

new and surprising functions of previously<br />

unknown receptors in processes like the<br />

uptake of binding partners or organ development,<br />

but all these pieces have to be understood<br />

and put into place before we understand<br />

certain organs as a whole and how they<br />

are connected to other body systems.”<br />

The “genome age” has brought a mass of data<br />

to the table, he says, but it is not well integrated.<br />

“Results of experiments are most useful<br />

when we can see their place in the whole –<br />

from simple things like where particular molecules<br />

are expressed in the kidney to more<br />

complex questions like what signaling pathways<br />

they participate in, and ultimately how<br />

that perturbs the physiology of the kidney,” he<br />

says. “But there haven’t been tools to ‘navigate’<br />

the data that way. This prompted us to<br />

propose an EU project that would capture<br />

data from groups throughout Europe and<br />

elsewhere to be plugged into a ‘Kidney Atlas.’<br />

In a way, it is a virtual organ that we’re building<br />

on the web now.”<br />

The atlas is compiling information from laboratories<br />

working with a variety of model<br />

organisms including mice, fish, frogs, and rats.<br />

Direct participants include 18 research groups<br />

from universities and other institutes, plus six<br />

university clinics, spread across nine European<br />

countries. The goal is to create a four-dimensional<br />

map of the kidney, showing how molecules<br />

guide its development and activities<br />

from its origins in the early embryo to adulthood.<br />

During a disease all of these levels are<br />

affected, Thomas says, sometimes over an<br />

organism’s entire lifespan. Solving the health<br />

problems related to kidneys will require scientists<br />

to develop an integrated view of the kidney<br />

as a whole and its links to the rest of the<br />

body.<br />

173<br />

You’d need a fat user’s manual to keep track of<br />

all the parts and possible functions of the<br />

Giant Swiss army knife. “And that’s a knife<br />

with only 85 tools,” Thomas says. “Imagine the<br />

manual you would need to describe the thousands<br />

of molecules that build and operate the<br />

kidney, or other organs for that matter. At the<br />

moment that information exists in bits and<br />

pieces, not assembled into a coherent picture.<br />

That’s the main task of the atlas. It’s a big job,<br />

but every bit of knowledge added to this virtual<br />

kidney will pay off in understanding the<br />

real one.”


Bernd Reif


Bad origami<br />

Even with step-by-step instructions, some of us would never be<br />

able to fold a sheet of paper into an elegant origami flower or<br />

crane. Let alone the model of an ankylosaurus, a dinosaur, which was<br />

designed by Ronald Koh of Singapore and can be found on the Internet.<br />

It has 61 steps, and that’s not really counting right, because most of<br />

some of them tell you things like, “Repeat the fold you just made 20<br />

times.” I gave up on the third step.<br />

But suppose you had to learn out how to make the origami ankylosaurus<br />

by unfolding a finished one. That would give you an idea of<br />

how Bernd Reif feels when he looks at β-amyloid plaques, the culprits<br />

in Alzheimer’s disease, introduced in the previous story. In trying to figure<br />

out how they form, the situation is like that of the crane: even<br />

knowing a lot about the starting material (paper) and seeing the finished<br />

product (a paper flower) doesn’t explain how to get from one to<br />

the other. Some of us have wastebaskets full of failed origami projects<br />

to prove it.<br />

“Amyloid plaques accumulate between nerve cells, leading to cell death<br />

and the symptoms of Alzheimer’s,” Bernd says. “But we understand very<br />

175 Part Three: Frontiers and ferrymen


The Reif group at the NMR spectrometer.<br />

Mangesh Joshi, Rasmus Linser, Bernd Rief<br />

and Veniamin Chevelkov


Part Three: Frontiers and ferrymen<br />

178<br />

Mangesh Joshi<br />

little about what causes those effects. We think<br />

that protein fragments assemble into amyloids in<br />

an orderly way, going through many steps in<br />

which they take on different structures. Some of<br />

those structures are more toxic than others,<br />

although they are made of the same subunits.<br />

Finding out what makes the plaques so dangerous<br />

will probably require getting a step-by-step look at<br />

how the protein fragments assemble.”<br />

Bernd’s laboratory at the FMP is trying to uncover<br />

some of the basic principles underlying<br />

Alzheimer’s, Huntington’s Diseases, prion diseases,<br />

and other neurodegenerative conditions<br />

that are caused by accumulations of proteins.<br />

They all begin when a protein needed by the<br />

healthy body begins to behave in an unusual way.<br />

In the case of Alzheimer’s disease, brain cells produce<br />

protein fragments all the time, but normally<br />

the body can get rid of them.<br />

“The same small bits of protein – made of just 40<br />

amino acids – are sometimes completely harmless,<br />

sometimes fatal,” Bernd says. “If you watch<br />

them bind to each other in the test tube, you get<br />

a variety of forms: single fragments that don’t<br />

accumulate, or stringy fibers, or clumps that are<br />

not very organized. What you get depends on the<br />

experimental conditions.”<br />

The plaques may assemble differently in the test<br />

tube than in the brain, and even under controlled<br />

conditions it has been difficult to get a glimpse of<br />

the stages. The battle, he says, has been to find or<br />

invent methods to get a look at single stages of<br />

how protein fragments interact with each other<br />

and other molecules found in the clumps. Call it a<br />

folding manual.<br />

��<br />

If β-amyloid isn’t yet a household word, it probably<br />

will be soon. More and more families are<br />

confronting Alzheimer’s disease, the most common<br />

age-related neurodegenerative disease in<br />

the world.<br />

The previous story describes how β-amyloid<br />

begins as part of a much larger molecule called<br />

APP, found in high numbers in the membranes of<br />

nerve cells. What it normally does there is still a


mystery. Mice born without the molecule develop smaller brains, problems<br />

of body coordination, and impaired memory and learning, but<br />

scientists do not know how APP contributes to these processes in<br />

healthy animals or humans.<br />

Like most proteins, APP is processed by other molecules in several<br />

ways. First it is cut by an enzyme called β-secretase, or BACE. Another<br />

protein called γ-secretase cleaves the chain for a second time, releasing<br />

the β-amyloid fragment. This is a normal, life-long process that<br />

undoubtedly has a role in the healthy brain, but the fragment’s function<br />

is not yet known. Normally it is broken down and disposed of, but<br />

with age our cells and molecules behave differently. Later in life the<br />

fragments accumulate to form fibers and plaques in the space<br />

between cells.<br />

“There are a few theories about why the aggregations are dangerous,”<br />

Bernd says. “One is that they create holes in cell membranes that allow<br />

charged calcium atoms to pass through. That would be bad because<br />

179<br />

calcium ions have to be carefully regulated for the cell to behave properly<br />

and survive. Another idea is that plaques stabilize or produce free<br />

radicals, highly reactive atoms that are dangerous because they can<br />

modify a wide range of molecules. Whatever the answer is, β-amyloid<br />

only does this in a specific type of cluster, or what we call an oligomer.<br />

That’s simply a complex in which more than one copy of the same<br />

molecule is bound together. For example, if copies assemble in<br />

oligomers of six or eight they might not be dangerous, whereas<br />

groups of ten could be fatal. Determining which of these forms is the<br />

deadly one has been extremely challenging.”<br />

And amyloid plaques contain more than β-amyloid. (“The pleasures<br />

and powers of green tea,” at the end of the book, describes an <strong>MDC</strong><br />

group’s efforts to find other molecules that dock onto the fragment<br />

and influence its folding.) Recently another laboratory discovered that<br />

a chaperone protein called α-B-crystalline could bind to the fragments<br />

in the test tube as well as in living cells.<br />

Part Three: Frontiers and ferrymen


Chaperones are “assistants” whose normal job is to help keep proteins<br />

folded in their proper forms – if they become unfolded, the cell usually<br />

destroys them. Folding gives a protein the shape it needs to interact<br />

with other molecules and snap into “molecular machines.” A protein<br />

begins as a long string; folding happens because of chemical interactions<br />

between the amino acids that make it up. Sometimes it needs<br />

help to achieve or maintain the right form, which is where chaperones<br />

come in. Experiments showed that the brains of some Alzheimer<br />

patients contained chaperones, including α-B-crystalline, in abnormal<br />

places.<br />

“That’s interesting because α-B-crystalline is disturbed in people who<br />

suffer from Down Syndrome,” Bernd says, “and nearly all of them<br />

develop Alzheimer’s disease. They develop it early, at around 40 years<br />

of age, which is a fairly dramatic acceleration of the disease process.”<br />

Experiments in the test tube revealed that α-B-crystalline blocks the<br />

formation of amyloid fibers. Fragments can still bind to each other in<br />

the molecule’s presence, but they don’t take on the long, orderly shape<br />

of fibers. Instead, copies of β-amyloid cluster together, in groups of<br />

twos (dimers), threes (trimers), or higher combinations. Some remain<br />

alone.<br />

“It is unclear why only one oligomeric structure should be toxic,” Bernd<br />

says. “To find out, we’ll have to watch the process of assembly. And that<br />

has been a huge technical challenge. What is very interesting is that<br />

the combination of β-amyloid with α-B-crystalline is more deadly to<br />

Part Three: Frontiers and ferrymen<br />

β-amyloid fibers like those found in the<br />

brains of Alzheimer’s patients (upper left).<br />

The presence of certain proteins or other<br />

substances can change the types of larger<br />

structures they form (shown in the other<br />

images). Most of these aggregations would<br />

probably not have the deadly effects seen in<br />

the disease.<br />

cells than β-amyloid on its own. Maybe if fibers could form, the situation<br />

wouldn’t be so dangerous; α-B-crystalline is keeping that from<br />

happening.”<br />

Were fibers a transition state – were they involved at all in the development<br />

of disease symptoms? Wondering why α-B-crystalline<br />

increased the deadliness of β-amyloid, Bernd’s lab applied several<br />

innovative techniques to study the nature of the contact between the<br />

two molecules and its relationship to the formation of fibers and<br />

plaques. The first question to answer had to do with basic architecture:<br />

what regions of β-amyloids were linked to each other, and which<br />

regions could bind to α-B-crystalline? Next the scientists wanted to<br />

know how many copies of β-amyloid could be found linked within a<br />

single cluster, and whether this changed because of the presence of α-<br />

B-crystalline. If changes occurred, they might happen in several steps.<br />

Sometimes in origami, a fold is made and then undone, so that the<br />

crease can serve another function.<br />

��<br />

Saravanakumar Narayanan, a PhD student from India, used a combination<br />

of methods to find some of the answers. First, β-amyloid<br />

and α-B-crystalline were mixed together at changing concentrations.<br />

Presumably a small number of copies of α-B-crystalline would have<br />

different effects on the fragment than a high number. As the amounts<br />

changed, the structures were examined by NMR, which provided an<br />

atom-by-atom view of some regions of the molecules.<br />

180


The experiments showed Saravanakumar and his colleagues the docking<br />

sites of the two molecules. A tiny region of β-amyloid permits two<br />

or more copies to bind to each other. The same region serves as the<br />

binding site for α-B-crystalline. In fact, α-B-crystalline binds much<br />

more strongly to β-amyloids than the fragments bind to each other.<br />

“That is extremely interesting,” Bernd says, “because it suggests a way<br />

by which α-B-crystalline can change the groupings and the function<br />

of β-amyloid fragments. Without α-B-crystalline, there is nothing to<br />

keep the fragment from forming large groups, including structures<br />

like fibers. When the molecule is present, however, it competes for the<br />

binding site. Fewer copies of β-amyloid will bind.”<br />

The findings closely echo what the group found while investigating<br />

another protein, Sup35, using similar methods. Sup35 is a yeast prion<br />

protein, similar to the human protein that causes prion diseases (like<br />

Mad Cow Disease). Both yeast and human prion proteins have functions<br />

in healthy cells that they can no longer perform when they have<br />

accumulated in clumps. Saravanakumar and Bernd discovered that a<br />

chaperone protein called Hsp104, known to interact with Sup35, links<br />

to clusters of Sup35 molecules and tosses out some of them, leading<br />

to smaller clusters.<br />

α-B-crystalline might be doing something similar, but the scientists<br />

needed to know if it was changing the number of β-amyloid mole-<br />

cules attached to each other. While investigating the yeast prions, the<br />

scientists had developed a method to filter different sizes of molecules<br />

by placing them in a dialysis bag and allowing them to leak out. Such<br />

bags are used to filter the blood of people with kidney diseases; they<br />

are full of holes so tiny that they trap large molecules while allowing<br />

small ones to pass through. By using bags with holes of different sizes,<br />

the scientists could test amounts of Sup35 molecules that were<br />

detached from Sup35 fibrils. They could also track Hsp104’s effects on<br />

the size of these groups.<br />

Similar experiments revealed that α-B-crystalline binds weakly and<br />

only for a very short time to single copies of β-amyloid. But it binds<br />

very strongly to multiple copies of the fragment that have formed<br />

fibers or are in the process of doing so. “When that happens, it causes<br />

a structural change,” Bernd says. “We believe that α-B-crystalline shifts<br />

the preference of the fragments – instead of linking up into long,<br />

stringy fibers containing thousands of copies of the fragment, they<br />

collect in smaller clusters.”<br />

The association between the molecules has an additional effect: one<br />

of the amino acid subunits of β-amyloid becomes oxidized. This means<br />

that the molecule picks up additional electrons, changing its overall<br />

charge and the way it interacts with other molecules. That, Bernd says,<br />

might help explain why amyloid plaques are so deadly.<br />

181 Part Three: Frontiers and ferrymen


“Oxidation seems to destabilize fibers so that they can dissolve,” he<br />

says. “That might also explain why this form of β-amyloid is toxic to<br />

neurons, because it produces highly reactive forms of oxygen.”<br />

To test the idea, the scientists used a probe that could measure<br />

α-B-crystalline’s oxidizing activity. They put α-B-crystalline into a mixture<br />

of single molecules called GSH. On its own, the oxygen that is dissolved<br />

in the solution provides an electron that oxidizes GSH. Adding<br />

another protein may slow down the process, because it also undergoes<br />

oxidation and captures some of the electrons that would otherwise<br />

be used by GSH. That’s what happened when Bernd and his colleagues<br />

mixed the two types of molecules. Mixing GSH with β-amyloid<br />

fragments, however, had almost no effect. This meant that the<br />

fragments alone weren’t producing the oxidizing reactions.<br />

To “turn up the heat,” Bernd and his colleagues added copper to the<br />

mixture of GSH and α-B-crystalline. “Copper doesn’t change the oxidation<br />

state of GSH on its own,” Bernd says, “but in combination with α-<br />

B-crystalline, it triggers much faster reactions. Our interpretation is<br />

that copper binds tightly to α-B-crystalline and pushes its oxidizing<br />

activity into higher gear.”<br />

��<br />

Bernd is just as helpless as I am when it comes to origami; his artistic<br />

talents lie elsewhere. He has always loved singing, but things<br />

started to get serious when he began his undergraduate degree in<br />

physics in Beyreuth, Germany. (If you aren’t familiar with that town,<br />

you’ve never heard of Richard Wagner, who established an opera<br />

dynasty there.)<br />

Bernd won a Richard-Wagner scholarship and took singing lessons<br />

from Manfred Hegen, a school teacher and singing coach. Some of<br />

Hegen’s students have made it into the choirs of the Beyreuth music<br />

festivals or gone on to have impressive careers. Hegen founded Musica<br />

Vocalis, a concert choir. First Bernd joined the choir, then graduated to<br />

singing arias. He has performed in excerpts from operas and Bach’s<br />

Christmas Oratorio, and recently Creation, by Haydn. There is a rumor<br />

that he sometimes performs at scientific conferences, so we’ll have to<br />

Part Three: Frontiers and ferrymen<br />

follow up on that. And as if there weren’t already enough music in his<br />

life ... he is married to a violinist in the Berlin Opera orchestra.<br />

Bernd is an example of the possibility of combining a scientific career<br />

with a love for music, but it isn’t always easy, and there have been<br />

some dry spells. When he spent three years at the Massachussetts<br />

Institute of Technology in the USA on a postdoctoral fellowship, it was<br />

hard to find a choir.<br />

Back in Germany that would be less of a problem, but at the beginning<br />

he had his hands full setting up his own group in Munich, at the<br />

Institute for Organic Chemistry and Biochemistry II of the Technical<br />

University. He began thinking about ways to combine methods to get<br />

a look at the way proteins cluster. Most questions about protein structure<br />

were being solved by crystallography, a process that teases highly-purified<br />

proteins into a lattice-like array, like stacking shoeboxes<br />

onto shelves. If the molecules formed neat enough rows, they could be<br />

examined by bombarding them with X-rays. But not all proteins<br />

behave this way. Some, particularly membrane proteins like APP, don’t<br />

form tidy crystals. Besides, the process locked molecules into stiff<br />

arrangements – not a very good way to watch a delicate process that<br />

might involve many proteins. While in the USA, Bernd had been introduced<br />

to another method, NMR, which might provide answers.<br />

��<br />

We still have a long way to go in understanding the molecular<br />

processes that lead to Alzheimer’s disease, Bernd says. “We’re<br />

asking questions that tax all of our current methods. Finding answers<br />

has required combining techniques in new ways.”<br />

For example, the laboratory would like to get a look at how molecules<br />

bind to β-amyloid in the fiber form. NMR, the technique of choice at<br />

the FMP for getting structural information, works best with molecules<br />

that float freely in solution. It applies a strong magnetic field to protein<br />

samples. This alters the energy levels in the nucleus of every atom<br />

in the sample in a way which is slightly different for each proton in a<br />

protein. The slight energy difference allows scientists to determine the<br />

182


identity of each proton – in other words, to tell what type of atom it<br />

belongs to. When the strength of the magnetic field is lowered, or<br />

“relaxed”, they can probe the distances between protons, permitting<br />

the creation of atom-by-atom maps and thus a structure of the protein.<br />

But for the method to work, molecules need to swim around<br />

freely, which they can’t do if they are locked into a fiber.<br />

In addition, Bernd says, one can try to get a lot of molecules in the<br />

sample to align themselves in the same direction. He knew that some<br />

of his colleagues had been using magnetic fields to force fibers into a<br />

more organized structure, allowing them to obtain higher-resolution<br />

images using X-rays. Maybe the same thing could be done using the<br />

powerful magnetic field of the NMR machine.<br />

“The amino acid subunits of proteins have different magnetic properties,”<br />

Bernd says. “Think of a fibril as a shoestring that several small<br />

magnetic beads have been strung onto – representing the most magnetic<br />

subunits of the protein. If you lay the string alongside an iron<br />

pipe, it will attach itself and hold on. This is an analogy of how the protein<br />

fiber can be oriented in a certain way by the magnetic field. Now<br />

if other proteins come along and dock on, they can only attach themselves<br />

in certain ways. Those constraints give us enough information<br />

to see some of the details.”<br />

It often takes months or years to get such ideas to work, Bernd says,<br />

and in many institutes scientists don’t have the freedom. “Methods<br />

development is dry compared to some of the science stories,” he says,<br />

“and it is sometimes difficult to publish interdisciplinary papers where<br />

the focus is on methods. But the β-amyloid studies are a good example<br />

of how techniques have to be adapted or used in creative new<br />

ways to solve the kinds of questions we are asking right now. Walter<br />

Rosenthal, the FMP’s director, has been very supportive of this<br />

approach.”<br />

��<br />

With method developments, a wide range of scientific interests<br />

and his love of music, Bernd clearly has his hands full, but he<br />

recently took on another major project with the creation of a new<br />

graduate school in Molecular Biophysics at the FMP. The program has<br />

now been launched in collaboration with universities and other institutions.<br />

The Leibniz Association has financed the creation of several of<br />

graduate schools, on a competitive basis, over the past few years.<br />

Walter Rosenthal says that Bernd’s proposal to create a department<br />

devoted to biophysics – specifically the investigation of protein interactions<br />

– was successful for several reasons.<br />

“In the first place the topic fills a gap in the landscape of German academic<br />

programs,” Walter says. “And it was a very strong proposal.<br />

Bernd developed the concept, convinced all of us and the partner institutes,<br />

and wrote the grant application.”<br />

Currently there are 15 students enrolled in the program, about half<br />

doing their work in Berlin and the rest at other institutions. Each<br />

receives a healthy, competitive stipend – to which some expectations<br />

are attached. The students present their work once a year in an annual<br />

report, and participate in two major “block” courses whose purpose<br />

is to give them a broader background than students normally receive.<br />

“Biophysical techniques are acquiring a growing role in the investigation<br />

of biological processes,” Bernd says. “We’re also in the midst of a<br />

very exciting time in which new methods are being invented all the<br />

time. Young scientists will need a very wide exposure to the palette of<br />

methods that are available and should think creatively about new<br />

ones.”<br />

The proposal captures the best experts in particular areas from the<br />

region, Bernd says. It will draw on campus expertise in X-ray crystallography,<br />

NMR, and signal transduction; institutes within the Charité will<br />

contribute in areas such as electron microscopy and biochemistry; the<br />

Institute for Experimental Physics and other departments of the Freie<br />

Universität will offer spectroscopy and other types of expertise, and<br />

the Humboldt-Universität adds expertise in membrane biophysics<br />

and several other fields. The University of Potsdam is another partner<br />

in the areas of CD spectroscopy and thermodynamics.<br />

“In putting together this program, we’ve collected a wide range of<br />

techniques practiced by groups that have a lot of experience working<br />

together,” Bernd says. “Scientific themes will include the biophysics of<br />

protein aggregations, such as our work on Alzheimer’s disease; signal<br />

transduction in processes like visual perception; the molecular<br />

machines that build proteins, and viral infections. The questions in all<br />

of these fields are extensive and require a combination of approaches.<br />

Our plan is that each student in the program will be mentored by two<br />

group leaders, to increase interdisciplinarity.”<br />

What about a PhD program choir? That would certainly be interdisciplinary.<br />

“Maybe someday,” he laughs. “We’d need a pianist. And we might have<br />

to change the criteria by which we select our students.”<br />

183 Part Three: Frontiers and ferrymen


Trouble in the<br />

waterworks<br />

It’s hard to imagine an advanced society without electricity. Hard to<br />

imagine that progress and luxuries could be achieved without this<br />

magical river that stems from the turbines of coal plants, dams and<br />

windmills, flows into conduits underneath our streets and the engines<br />

of factories, into secret spaces inside the walls of our houses. Yet without<br />

electricity, the ancient Romans achieved a standard of living – at<br />

least their rich did – that seems sumptuous even by today’s standards.<br />

They managed it by building a society based on water.<br />

Visitors to Pompeii and Herculaneum can view a film in which, thanks<br />

to computer animation, the towns are restored to their full glory in the<br />

days before Vesuvius erupted and covered them in meters of ash and<br />

lava. Everywhere in this world is water: springing up through fountains,<br />

pouring into public baths, irrigating colorful gardens. Without it, Roman<br />

society would have crumbled.<br />

Robert Harris’ novel Pompeii takes us underground in this ancient world.<br />

In one scene a Roman engineer climbs the slopes of Mount Vesuvius<br />

and enters a vast subterranean cavern, trying to discover why water has<br />

stopped flowing. The problem stems from an earthquake that has<br />

Part Three: Frontiers and ferrymen<br />

184


Enno Klußmann


Walter Rosenthal and Enno Klußmann<br />

shifted the ground and shoved aqueducts out<br />

of alignment. As the water slows to a trickle in<br />

the towns around the volcano, people begin<br />

to panic. Life is utterly dependent on water,<br />

and its loss seems much more threatening<br />

than rumblings on the mountain.<br />

��<br />

Our own existence is completely dependent<br />

on water, mostly on recycling it.<br />

While a human body is 70 percent water, only<br />

a fraction of that comes from our daily intake<br />

of liquids and food. The rest – 99 percent – is<br />

won through recycling. The kidneys process<br />

about 180 liters per day. Recovering nearly all<br />

of that water is the job of cells in the body’s<br />

waterworks – the kidney – and they only manage<br />

because they are leakier than other cells.<br />

If it weren’t for proteins called aquaporins, we<br />

would quickly die of dehydration even without<br />

any disruptions in the water supply. That<br />

partly explains the interest of two researchers<br />

at the FMP: Enno Klußmann and Director<br />

Walter Rosenthal.<br />

In the 1920s researchers figured out that the<br />

membranes of cells were made of fat molecules<br />

called lipids, put together in such a way<br />

that small quantities of water could simply<br />

slip through. But that amount was clearly not<br />

enough to supply the body’s big drinkers, like<br />

kidney and red blood cells, so there had to be<br />

another way for water to enter. Fifty<br />

years later, this “water channel” still hadn’t<br />

been found. Then along came Peter Agre, a<br />

young hematologist who was working at<br />

the John Hopkins School of Medicine in<br />

Maryland.<br />

Part Three: Frontiers and ferrymen<br />

186<br />

“The field was essentially stuck,” Agre says,<br />

“but following the well known scientific<br />

approach known as ‘sheer blind luck,’ we<br />

stumbled upon the protein that is the answer<br />

to the question: do water channels exist?”<br />

Agre and his colleagues had found the first<br />

aquaporin, a discovery so important that he<br />

was awarded the Nobel Prize in 2003.<br />

He was also the FMP’s first choice as a speaker<br />

when the institute was thinking about who<br />

should give the main talk at the campus New<br />

Year’s party; in 2008 it was the FMP’s turn to<br />

host the event. In his introduction, Walter<br />

Rosenthal commented on the man as well as<br />

his work, mentioning the fact that Agre was<br />

an Eagle scout and a cross-country skier. He<br />

also dwelt for a moment on Agre’s political<br />

activities:


“I have also been deeply impressed by your<br />

political activities, which are inspired by your<br />

strong belief in humanity. You jumped to the<br />

defense of Thomas Butler, who voluntarily<br />

reported to the university safety office that 60<br />

vials of plague bacteria were missing and had<br />

probably been autoclaved. You defended his<br />

refusal to plead guilty, which ultimately led to<br />

him receiving a two-year prison sentence. You’re<br />

also a founding member of a movement called<br />

Scientists and Engineers for Change and the<br />

slide shows one paragraph of the Bill Of Rights<br />

of this initiative. It reads: ‘The federal government<br />

shall not support any science education<br />

program that includes instruction in concepts<br />

that are derived from ideology and not science.’”<br />

Agre is widely known as a modest man, and<br />

when he took the podium he talked about his<br />

career in science and its relationship to the<br />

rest of his life, particularly the support he had<br />

received from his family and colleagues. A<br />

high point of the talk was his solo performance<br />

of “the Elements song,” by Tom Lehrer,<br />

dashing through the entire periodic table of<br />

the elements in rhyming verse.<br />

��<br />

Since Agre’s key discovery in the 1980s, scientists<br />

at the FMP and elsewhere have<br />

learned a lot about how aquaporins work.<br />

These proteins float in the plasma membrane<br />

– the barrier between the cell and the world –<br />

where they create passageways. Proteins are<br />

folded in specific ways as they are made, and<br />

the folds of aquaporins leave a gap through<br />

the middle that permits the passage of water<br />

molecules. When the body needs water, aquaporins<br />

are added to the plasma membrane of<br />

cells; when enough has been absorbed, they<br />

are removed again. Cells know how much<br />

water to let in because of hormone signals<br />

that get released when the supply gets low.<br />

The hormone doesn’t directly communicate<br />

with aquaporins; first it has to trigger another<br />

signal, called cAMP, which does that job.<br />

Enno Klußmann began looking at aquaporins<br />

as a postdoctoral fellow in the research group<br />

of FMP Director Walter Rosenthal. He joined<br />

the lab in 1997 – “on loan” from the Günther<br />

Schultz’s group at the Freie Universität Berlin,<br />

where he had attempted to identify a receptor<br />

for steroid hormones in the cell membrane.<br />

University regulations permitted postdoctoral<br />

fellows to remain no longer than five<br />

years at a time, and Schultz recommended<br />

that he join the group of another former student,<br />

Walter Rosenthal, at the FMP. “The idea<br />

was that I would leave for a few months and<br />

then go back,” Enno says. “But Walter’s lab was<br />

working on some fascinating things, and ten<br />

years later, I’m still here.”<br />

One of the group’s projects was the study of<br />

aquaporins. “The way that aquaporins behave<br />

in the kidney is important and interesting,”<br />

Enno says, “but it opens the door on some<br />

187 Part Three: Frontiers and ferrymen<br />

even more interesting questions. cAMP is a<br />

strong signal, found almost everywhere in the<br />

cell. It triggers a wide range of other processes<br />

often at the same time, in the same cell.<br />

That situation raised the question of how the<br />

cell controls the signal – how it manages to<br />

respond in specific ways, in specific places.<br />

Without controls, everything would be happening<br />

everywhere, all the time.”<br />

Many processes triggered by cAMP are linked<br />

to diseases – including hypertension, gastric<br />

ulcers, and heart disease. It turns out that<br />

some common drugs work by preventing<br />

other proteins from “hearing” cAMP signals.<br />

That’s desirable if the drug blocks only one<br />

kind of activity. But there may also be unintended<br />

side effects. In the 1990s doctors discovered<br />

that lithium, commonly used to treat<br />

people with a psychiatric problem called bipolar<br />

disorder, was interfering with aquaporin.<br />

Patients on the drug developed a form of diabetes<br />

in which their bodies lost their ability to<br />

recycle most of the water. Even though they<br />

drank all the time, most liquids just passed<br />

right through without being absorbed. This<br />

was bad news because lithium was helping<br />

many bipolar patients, and it was also being<br />

considered as a supplement in the treatment<br />

of other diseases. A better understanding of<br />

aquaporins might help researchers find a way<br />

to prevent the side effects.<br />

A drug that blocked all of cAMP’s activity<br />

would be fatal, but maybe there were ways to


stop a single process. Learning to do that<br />

might show how to block other processes.<br />

Enno and Walter hoped that an aquaporin<br />

called AQP2 would make a good place to start.<br />

��<br />

Things went so well for Enno’s work at the<br />

FMP that he could set up his own lab in<br />

2001. He brought along the aquaporin project,<br />

and has continued to work on it in collaboration<br />

with Walter and a network of researchers<br />

across Europe. He heads a project funded by<br />

the European Community to identify or create<br />

small molecules that can control cAMP signaling<br />

in processes related to human diseases.<br />

Enno and Walter have also been awarded a<br />

Part Three: Frontiers and ferrymen<br />

188<br />

number of grants from the German Research<br />

Council (DFG) to study particular aspects of<br />

cAMP signaling and aquaporins.<br />

Enno says that understanding how cells control<br />

water channels – and thus the body’s<br />

water supply – boils down to several questions.<br />

Those have to be tackled one by one,


and over the past few years the FMP has<br />

made progress on them all. The first question<br />

is how the molecules are moved to the locations<br />

where they are needed.<br />

“The molecules triggered by cAMP have to get<br />

to the right place to respond properly to a signal,”<br />

Enno says. “One way to do that is to tie<br />

them up to particular membranes. In addition<br />

to the plasma membrane that surrounds the<br />

cell, there are a lot of internal membranes<br />

that often serve as duty stations. Tethering is<br />

how the system that activates aquaporins<br />

works. AQP2 is first stored in internal membranes<br />

before it is moved to the cell surface.<br />

Inside the cell it undergoes a kind of pre-fabrication;<br />

it is installed alongside the machinery<br />

that controls it. That includes a switch, which<br />

is the actual receiver of the cAMP signal,<br />

called a PKA protein. It also includes the<br />

anchor that attaches the PKA to the membrane,<br />

and other molecules. There are lots of<br />

types of anchors – so far, about 50 have been<br />

found. The anchors for PKA are known as<br />

AKAPs (for A-kinase anchoring proteins). The<br />

cell has so many because each one is used to<br />

anchor a particular type of molecule to its<br />

proper place.”<br />

In 2001 it wasn’t certain that aquaporin functions<br />

depended on the interaction of PKA with<br />

an anchor, so Enno and Walter decided to find<br />

out. For their experiments they used cells<br />

taken from rat kidneys, raised in laboratory<br />

cultures, which produced aquaporins.<br />

Microscope studies had shown that water<br />

channel proteins were held in the interior of<br />

the cells, in small membrane-wrapped compartments<br />

called vesicles, until they were<br />

needed. When the cAMP signal was received,<br />

the aquaporins moved to the plasma membrane.<br />

The scientists began shutting down<br />

proteins in the cells, hoping to block this<br />

movement. Success would mean that they<br />

had found a protein that had something to do<br />

with positioning the aquaporin. Since it was<br />

likely that an AKAP was involved, one thing<br />

they tried was to prevent the anchor from<br />

binding to PKA. To do this they used a small<br />

molecule, a synthetic peptide, known to block<br />

the contact between the two molecules.<br />

Cutting an anchor sets a boat adrift, and if an<br />

AKAP is disturbed, signaling molecules float<br />

away from their duty stations – or never<br />

become attached there in the first place. This<br />

renders the anchor an interesting target for<br />

drugs. “You wouldn’t want to block cAMP<br />

itself to stop a process – that would disrupt a<br />

lot of other important things going on in the<br />

cell that rely on the signal,” Enno says. “The<br />

anchor is a much better point of attack.”<br />

Enno and Walter introduced the synthetic<br />

peptide that disrupts anchors into the kidney<br />

cells; they discovered that aquaporins were no<br />

longer arriving at the membrane. An anchor<br />

was obviously involved, but which one? In<br />

2006 Postdoc Volker Henn, PhD student<br />

Bayram Edemir found a candidate: a protein<br />

called AKAP18-δ. Microscope studies and<br />

other methods showed that AKAP18-δ was<br />

installed alongside AQP2 in the cell interior<br />

and moved along with water channels when<br />

they moved to the plasma membrane – “guilt<br />

by association.”<br />

Another version of the AKAP18-δ protein influences<br />

the contraction of the heart; there,<br />

cAMP signals control a different type of channel<br />

which allows the passage of calcium ions<br />

into the cell. These channels are stimulated<br />

through another hormone. “The situation is<br />

similar to that of aquaporins,” Enno says. “The<br />

hormone signal triggers a wave of cAMP signaling<br />

that opens channels in the plasma<br />

membrane. But only if the channel proteins<br />

are connected to anchors.”<br />

Beta-blockers, which are commonly used in<br />

the treatment of cardiovascular disease, work<br />

because they tune down the heart’s response<br />

to the hormone. The same thing might be<br />

achieved by blocking the connection between<br />

anchors and their partners, such as calcium<br />

channel proteins. “This makes the anchor a<br />

very good target for drugs,” says Christian<br />

Hundsrucker, another PhD student in Enno’s<br />

group. “The peptides we had been using to<br />

189 Part Three: Frontiers and ferrymen<br />

block AKAP-PKA interactions don’t do so very<br />

effectively. So we began trying to make new<br />

ones.”<br />

The scientists began by taking a closer look at<br />

the docking site. Normally two PKAs form a<br />

pair when the DD domain, a module in one<br />

protein, binds to its counterpart in the tother.<br />

Two domains joined in this way create a surface<br />

called RII that plugs into the anchor.<br />

Structural pictures of anchors had already<br />

been obtained using NMR and X-rays,<br />

providing a model of one of the surfaces. This<br />

could be used in “docking” simulations –<br />

twisting and turning the RIIs until the molecules<br />

snapped together on the computer<br />

screen.<br />

Christian says that the anchor has a coiled<br />

structure which fits into a pocket formed by<br />

RIIs. Whether the coil fits at all and how<br />

strongly it binds depend on the shape and<br />

chemistry of the two structures, which are<br />

determined by chemistry of the the amino<br />

acid building blocks that make up the proteins.<br />

Christian hoped to make a molecule<br />

that would fit better and bind even more<br />

strongly to RII, which would then lock onto<br />

the domains of PKAs and prevent them from<br />

acquiring an anchor. Christian began with a<br />

fragment of AKAP18-δ and started to change<br />

its recipe.<br />

Understanding how the surfaces of two proteins<br />

bind is an enormous chemical puzzle<br />

that would probably be impossible to solve<br />

without the help of sophisticated computer<br />

techniques. Christian and his colleagues<br />

enlisted the help of the lab of Gerd Krause, a<br />

colleague at the FMP; they were using software<br />

to study protein binding. By scanning<br />

features of several different RII domains and<br />

their binding sites on AKAP proteins, the scientists<br />

could detect specific amino acids that<br />

seemed to be crucial to linking the molecules.<br />

Michael Beyermann’s group, experts in building<br />

small proteins, helped design new versions<br />

of the binding domain by replacing those subunits<br />

with others that might make a better fit.


The resulting artificial molecules were<br />

screened in heart muscle cells. The strategy<br />

had worked – they were taking the place of<br />

anchors, breaking the connections between<br />

AKAPs and their targets. “What this showed<br />

us is that we have identified the key points<br />

that permit binding between the anchor and<br />

its PKA partner,” Enno says. “This gives us a<br />

much clearer strategy as we develop new synthetic<br />

molecules.”<br />

The dozens of known AKAPs may have evolved<br />

from a single ancestral molecule, inheriting a<br />

common “docking station” that allows them<br />

to bind to their main partner, PKA. The situation<br />

is a bit like the way a boarding ramp at an<br />

airport fits the hatches of different airplanes.<br />

You could design something that would prevent<br />

a plane made by one manufacturer from<br />

docking onto the ramp, but it would probably<br />

Michael Gomoll<br />

also interfere with other types of planes. In<br />

the same way, a drug that stopped one anchor<br />

from binding to its PKA would probably stop<br />

others. So the binding site between anchors<br />

and PKA proteins probably wouldn’t be the<br />

best choice to interfere with.<br />

But the experiment is important for another<br />

reason, Enno says. “The approach is valuable<br />

as a proof of principle to verify that pharmacological<br />

interference with AKAP function is a<br />

feasible concept. We know a lot about the<br />

function of AKAP-PKA interactions. Less is<br />

known about functions of AKAP interactions<br />

with other proteins.”<br />

Instead of disrupting the contact between the<br />

anchor and PKA, the scientists figured it<br />

would be better to aim a drug at another part<br />

of AKAP18-δ – its targeting region. “This is the<br />

‘address label’ that sends the anchor to the<br />

same location as aquaporins,” Enno says.<br />

“Interfere with that and the anchor gets lost,<br />

so it can’t help aquaporins move.”<br />

One goal is to produce potential drugs; another<br />

is to give scientists tools to dissect other<br />

aspects of the molecules’ activity. Aquaporins<br />

don’t work forever, even when the cell is flooded<br />

with cAMP. A study begun by PhD student<br />

Eduard Stefan revealed one of the reasons. As<br />

the anchor links to an aquaporin, it also brings<br />

other molecules on board. One of them, called<br />

a PDE, is a sort of volume control for cAMP. It<br />

works by cutting a chemical link inside the<br />

cAMP molecule – like disconnecting a wire<br />

inside a radio – leaving pieces that can no<br />

longer pass the signal.<br />

PDE molecules are found throughout the cell,<br />

and they help tune down the signal. This fact,<br />

plus the cell’s ability to install different sen-


sors for cAMP in different places, is what<br />

keeps the cell from becoming overstimulated<br />

by the signal all the time. That they can bind<br />

to an AKAP means that the anchor collects an<br />

entire toolbox of molecules that give the cell<br />

a fine level of control over the water channels.<br />

��<br />

One of the lab’s ongoing projects is to<br />

clear up yet another question about<br />

how aquaporins take up their proper positions<br />

in the cell. cAMP delivers the signal to<br />

move the water channel protein to the cell<br />

membrane, but what actually carries the<br />

aquaporin there? This is really a two-part<br />

problem: if you want to move something<br />

through the cell you need a track to move<br />

things along and a motor to do the pulling.<br />

Enno says that scientists had been focusing<br />

on microtubules, a network of fibers that act<br />

as a scaffold to give the cell its shape and also<br />

function as tracks along which molecules are<br />

delivered throughout the cell. Motor proteins<br />

act as the locomotives, traveling down the<br />

microtubules with other proteins in tow.<br />

“There was experimental evidence that artificially<br />

breaking down microtubules stopped<br />

the transport of aquaporins,” Enno says. “And<br />

motors that use microtubules had been<br />

found in the same vesicles that contain aquaporins.<br />

This looked like another case of preassembly,<br />

in which aquaporins and the molecules<br />

that were needed to transport them<br />

were being wrapped up in the same package.”<br />

However, another lab had recently done a<br />

study of all the proteins found in aquaporin<br />

191 Part Three: Frontiers and ferrymen<br />

vesicles, and they failed to turn up the motor<br />

proteins. Enno began to wonder whether<br />

microtubules were really the delivery route –<br />

after all, they weren’t the cell’s only transport<br />

system. In 2005 a Japanese group did an<br />

experiment in which a drug was used to break<br />

down the microtubules in cells taken from<br />

kidneys. They discovered that even without a<br />

microtubule railway system, aquaporins could<br />

be delivered to the membrane. Pavel<br />

Nedvetsky, a postdoctoral fellow in Enno’s<br />

group, hoped to find an explanation.<br />

“Some papers from 20 or 30 years ago had<br />

already hinted that another type of filament<br />

called actin fibers might be involved,” Pavel<br />

says. “So we began looking for another transport<br />

system, one that could move along these<br />

filaments.”


Working with other members of the group<br />

and international partners, Pavel made an<br />

important discovery: a motor protein called<br />

myosin Vb, known to shuttle other types of<br />

vesicles back and forth along actin fibers, was<br />

consistently found in the neighborhood of<br />

aquaporins. Motors often require adaptor proteins<br />

and other helpers to find and attach a<br />

cargo, and one of myosin Vb’s usual assistants<br />

is a protein called Rab11. This gave the scientists<br />

two ways to investigate whether myosin<br />

Vb was acting as a transporter for aquaporins<br />

– by interfering with the motor and with the<br />

adaptor.<br />

Pavel first used a non-functioning version of<br />

the motor – it didn’t work because it was<br />

missing the regions that normally attach it to<br />

filaments. He introduced it into the kidney<br />

cells and watched what happened when they<br />

were stimulated with the “thirst” signal.<br />

Normally they would take in extra water. Now<br />

they didn’t, and a look under the microscope<br />

showed why: aquaporins were trapped in<br />

internal vesicles, and were no longer being<br />

transported to the cell membrane. In another<br />

set of experiments, Pavel interfered with<br />

Rab11’s ability to bind to the motor. The same<br />

thing happened; aquaporins became stranded.<br />

“These experiments gave us an idea of some<br />

of the proteins responsible for moving aquaporins<br />

around,” Enno says, “but they still didn’t<br />

tell us exactly what was happening. To understand<br />

what we found next, you have to realize<br />

that most of us believe that aquaporins get<br />

recycled. That means that an aquaporin isn’t<br />

finished once it has moved to the plasma<br />

membrane. It helps absorb water as long as<br />

the signal to do so continues, but when the<br />

signal stops it has to be removed from the<br />

membrane. We think that it gets transported<br />

to an internal compartment, where it gets put<br />

on hold, and then it can be called up for duty<br />

again.”<br />

Returning to the microscope, the scientists<br />

found that some aquaporins managed to<br />

Control<br />

AVP<br />

MyoVb tail AQP2 ZO-1 Overlay<br />

Interfering with the tail of the “motor protein” myosin Vb (green) affects the normal location of<br />

aquaporins (red) and how they are recycled in the cell. The top row shows cells with normal<br />

myosin Vb, and the bottom row shows how the locations of the motor and aquaporins if the<br />

motor doesn’t function. Second column: Aquaporin 2. Third column: labeling ZO-1, a protein found<br />

in tight junctions, clearly shows where the borders of the cells are.<br />

make it to the cell membrane without myosin<br />

Vb, but once there, they couldn’t be shuttled<br />

back inside or out again. The cell handles new<br />

aquaporins a bit differently than those it recycles<br />

– other motors and helpers probably lend<br />

a hand. But myosin Vb and Rab11 appear to be<br />

essential in handling the recycling of the<br />

motor channels. Myosin Vb is thought to have<br />

a similar role in other recycling events.<br />

“The aquaporin story is interesting when you<br />

think about diseases in which the body<br />

retains too much water, like hypertension or<br />

congestive heart failure,” Enno says. “In those<br />

cases the body still needs to be able to absorb<br />

water – just not so much. Maybe you could<br />

temporarily shut down the recycling center<br />

alone, without also turning off all the intake<br />

valves. To do that you’d have to get a good look<br />

at the recycling machinery, and find something<br />

you could perturb without disturbing<br />

the rest. That’s where these experiments are<br />

taking us.”<br />

193 Part Three: Frontiers and ferrymen


Bribing the ferryman<br />

In Greek mythology, the dark river Styx separates the world of the living<br />

and the dead. It is traversed by a ferryman named Charon, a<br />

gloomy, glowering giant who floats up and down the river, guarding the<br />

entrance to the underworld. To appease him the Greeks buried their<br />

dead with a coin under the tongue – a bribe, or a fare – to secure a passage<br />

to the afterlife. Those that didn’t have the coin, or who had been<br />

improperly buried, were cast out of the boat.<br />

Smugglers throughout history have used bribes to move their wares<br />

cross borders, so it shouldn’t be surprising that scientists are trying<br />

something like a bribe to slip drugs and other substances into cells. In<br />

this case the barrier that has to be crossed is also a sort of river: the cell<br />

membrane, two liquid layers of fats and proteins. One of its key functions<br />

is to keep out toxins, parasites, and a wide range of other foreign<br />

substances.<br />

“It wouldn’t do much good to design the perfect drug,” says Sandro<br />

Keller of the FMP, “if you couldn’t get it to a cell that was diseased, and<br />

then get it inside the cell to do its job. Managing that means designing<br />

a delivery vehicle that can get substances through the cell membrane.”


Sandro Keller


Once inside the cell, drugs have to survive other cellular defenses that<br />

target and break down foreign substances. And ideally the substance<br />

should last a long time; otherwise, the drug would have to be administered<br />

continually, in such high doses that it would probably be toxic.<br />

The task is tricky enough when the cargo to be delivered is a small protein<br />

or chemical compound, like most of today’s drugs. Scientists are<br />

now hoping to insert other things into cells: DNA, RNA, large proteins,<br />

or complex molecules that can serve as probes of the inner workings<br />

of the cell. Getting these past the membrane will require understanding<br />

how it is built and how it functions, and that’s the task Sandro and<br />

his laboratory have devoted themselves to.<br />

��<br />

It’s not often that a young scientist goes directly from receiving his<br />

PhD to his first independent research position, without moving<br />

abroad and passing through a stint or two as a postdoctoral fellow,<br />

but that’s the case with Sandro. It has been two years now since the<br />

Part Three: Frontiers and ferrymen<br />

transition, and he seems entirely at home in his office at the FMP; he<br />

has brought in some plants, and the bookshelves are full. But those<br />

who don’t know him – including the organizers of some of the conferences<br />

he gets invited to speak at – still mistake him for a student. He’s<br />

young, and looks young for his age, which might make it difficult to<br />

pick him out in pictures of the group. Sandro’s fast track to leading his<br />

own team is the result of a success story: while working on his PhD,<br />

Sandro began taking an interesting approach to the study of the cell<br />

membrane and how to get cargos inside.<br />

The membrane has two liquid layers, like a soap bubble. Both layers are<br />

built of fat molecules called lipids, interspersed with proteins. Many<br />

membrane proteins have a head region that hangs outside the cell, a<br />

tail that dangles in the interior, and a complex folded region that<br />

spans the two layers of the membrane. What happens in this zone has<br />

been difficult – often impossible – to observe, but that will probably be<br />

necessary to solve the problem of drug delivery.<br />

Sandro would like to get a look at the regions of proteins embedded in<br />

the membrane, to understand how they are linked to each other and<br />

to lipids, and to watch what happens to those membrane structures<br />

as molecules pass through on their way into the cell. The problem is<br />

that the inner space of the membrane is a bit like the zone of radio<br />

silence that astronauts pass through as they reenter the earth’s<br />

atmosphere. It is hard to get any direct information from the zone.<br />

“The classical methods used to study the structures of proteins don’t<br />

work with proteins that span the membrane. We have to be very creative<br />

and use less direct, biophysical techniques.”<br />

One method that the group is using involved creating artificial molecules<br />

that can function as probes. The goal is to understand how the<br />

subunits of proteins interact and cause the protein to fold in the presence<br />

of lipids. The folds determine the molecule’s shape and functions.<br />

“We started with a simple bacterial protein called Mistic, which has<br />

four domains that pass through the membrane,” Sandro says. “We<br />

have created four separate, artificial molecules – one corresponding to<br />

each domain – that can be used as a model to study the behavior of<br />

the protein. These regions are called -helical domains; they are found<br />

in many different types of proteins, and they play a crucial role in the<br />

196


structure and organization of the membrane. But we know very little<br />

about how internal forces in the molecules cause them to fold and<br />

interact in certain ways. We can manipulate these artificial molecules<br />

to test hypotheses about these processes and to expand what we<br />

have learned from Mistic to many other molecules.”<br />

The project is the subject of a grant proposal that Sandro submitted to<br />

the German Research Council (DFG) and which was approved in Spring<br />

2007. The grant has given Sandro two extra positions to pursue the<br />

project in his lab. Getting a glimpse of protein structures and functions<br />

in the membrane would help solve questions about how other<br />

molecules enter the cell. Most methods used to observe this process<br />

have drawbacks, Sandro says. Some leave “artifacts” that make it seem<br />

as if a protein has crossed into the cell when it hasn’t. Other techniques<br />

are restricted to one aspect of the problem; for example, they<br />

either show how a protein binds to the cell surface, or give hints about<br />

how it passes through – but not both. Different kinds of experiments<br />

give contradictory results.<br />

“All of this confusing data has led to debates about how molecules<br />

enter,” Sandro says. “There has been quite a bit of discussion about the<br />

peptide penetratin, which manages to get through the barrier.<br />

Penetratin is a small module of another protein that has to move into<br />

the cell during the development of the fruit fly. Penetratin is like a key,<br />

and if you attach it to another molecule, it sometimes will escort that<br />

into the cell as well. What was unclear was whether penetratin somehow<br />

entered the cell on its own, or whether it came wrapped in a vesicle<br />

– sort of a small membrane bubble. Cells often swallow up vesicles<br />

in a process called endocytosis. The process by which they absorb single<br />

molecules is different, and it’s important to know which one is<br />

happening.”<br />

In 2006 Sandro and collaborators at the FMP (Michael Bienert and<br />

Margitta Dathe) and at the Martin Luther University Halle-Wittenberg<br />

helped to demonstrate that the latter was the case, using a method<br />

called isothermal titration calorimetry. The technique monitors and<br />

compares the temperature of two cells: one a control, and one which<br />

is subjected to experiments. In this case the researchers injected pen-<br />

197<br />

etratin into the experimental cell. They monitored its interactions with<br />

lipid molecules by measuring the heat released upon adding lipid vesicles<br />

into the cell. The instrument is so precise that it can distinguish<br />

between different types of chemical reactions based on the heat<br />

released or consumed in the reaction cell. The study showed that penetratin<br />

was not entering on its own – suggesting that endocytosis<br />

plays a key role in the cellular uptake of the peptide.<br />

“Calorimetry is very good for tracking the movement of substances<br />

into the cell,” Sandro says, “but it is not easy to use, the results are difficult<br />

to analyze – and most researchers simply don’t have access to it.”<br />

Searching for a simpler method, Sandro and his colleagues began<br />

using fluorescence spectrometry to watch the passage of proteins<br />

through membranes. “This approach monitors lipid vesicles with ultraviolet<br />

radiation and measures light instead of heat,” he says. “It can be<br />

used with any molecule whose fluorescence properties change when<br />

it comes in contact with the cell membrane.”


Natalie Bordag<br />

The method uses ultrasound (sound that is inaudible to the human<br />

ear) to force the compound of interest into the interior of lipid vesicles.<br />

The vesicles are then diluted to “loosen up” the interactions between<br />

the compound and the lipid molecules. If the compound can cross the<br />

lipid membrane and leave the vesicles, its fluorescence properties<br />

change since it is no longer in contact with the membrane. If, however,<br />

the compound is not able to get out of the vesicles, it will remain<br />

bound to the membrane, and its fluorescence properties won’t<br />

change. The technique passed a proof of principle test by confirming<br />

the results of the earlier calorimetry studies of penetratin, and the laboratory<br />

is now using it to study other molecules.<br />

��<br />

Anew delivery vehicle for drugs will need a key – a cell-penetrating<br />

peptide like penetratin – ideally, one giving access to specific,<br />

well-defined types of cells. Whether a vesicle can dock onto a cell and<br />

deliver its cargo depends on the proteins and lipids on its surface. But<br />

having the right ingredients isn’t always enough; they also need to be<br />

put together the right way.<br />

“One of the simplest sorts of transport vehicles that people use to try<br />

to introduce substances into the cell is a small fat droplet called a<br />

micelle,” Sandro says. “This is a small greasy drop that self-assembles<br />

because of its chemistry. It’s the same phenomenon that you see<br />

when you put oil into water – it forms a drop. That’s because each oil<br />

Part Three: Frontiers and ferrymen<br />

molecule has a hydrophobic tail that wants to avoid contact with<br />

water and a hydrophilic side; the oil molecules cluster together and<br />

turn the hydrophobic tails inward, away from the water, toward the<br />

inside of the droplet.”<br />

Even when accompanied by a cell-penetrating protein like penetratin,<br />

micelles don’t always work well as transporters, he says. Micelles<br />

assemble as a single layer, and the cell membrane is a double layer of<br />

lipids. A membrane protein in a single-layer micelle probably won’t<br />

fold or behave the way it would in its normal environment of a doublemembrane<br />

layer.<br />

Some proteins dock onto lipids and are able to assemble membranes<br />

around themselves. The best-case scenario would be to have a delivery<br />

protein that did this, as well. If it could collect a high number of lipids,<br />

they might form a double layer. After a search, Margitta Dathe and colleagues<br />

at the FMP decided to work with a small molecule called A2.<br />

They discovered that attaching this peptide to a vesicle caused it to be<br />

absorbed by capillary cells in the brains of rats. Sandro wondered<br />

whether A2 could be prompted to assemble lipids around itself.<br />

That’s where the bribe came in; something could be added to A2 to<br />

make it more likely to build vesicles. “We knew from the literature of a<br />

way to do this that might work,” he says. “If you create an artificial<br />

molecule by combining a small protein and fatty acids, it tends to<br />

accumulate a lot more lipid molecules – maybe enough to make a<br />

double-layered membrane. So we modified A2, giving it binding sites<br />

198


that lipids could dock onto, and another element that should serve as<br />

a binding site for glycoproteins. These proteins are found in membranes<br />

everywhere and when they bind to their partner molecules,<br />

they cause them to be taken up into cells.”<br />

In the test tube, the new hybrid A2 molecules bound to each other in<br />

clusters. Isothermal titration calorimetry revealed that the groups contained<br />

a “record-setting” number of lipids, compared to other hybrid<br />

molecules. A positive sign, Sandro says, because it made it more likely<br />

that delivery vehicles with double-layered membranes would form.<br />

Next it was time to see how cells would respond to the peptide. The<br />

scientists “doped” vesicles with A2 and watched what happened using<br />

the laser scanning microscope and other techniques.<br />

They discovered that vesicles with A2 readily docked onto cells and<br />

were carried through the membrane very efficiently; once inside, they<br />

dissolved slowly. This hints that in addition to being better at entering<br />

cells, vesicles with A2 might protect a drug long enough for it to be<br />

effective.<br />

All in all, the new version of A2 seems to work well as a “bribe” to get<br />

cells to take up drugs and other substances. “The A2 vesicles and<br />

micelles seem to be a significant improvement over some other types<br />

of delivery vehicles,” Sandro says. “It suggests that you might be able<br />

to add A2 to other transporters and improve their characteristics.<br />

We’re trying that now. We also hope to use the vesicles to solve other<br />

kinds of problems – for example, to insert complex probes that can<br />

monitor cellular processes.”<br />

��<br />

The River Styx and its ferryman are images from the dark corners<br />

of Western culture; they figure prominently in Dante’s Inferno<br />

and centuries of retellings of the myth of Orpheus. Michelangelo<br />

depicted Charon and his boat among the writhing figures of the Last<br />

Judgment. Despite their gloominess, the literary references appeal to<br />

Sandro, who considered doing many other things before committing<br />

himself to science.<br />

Berlin doesn’t have the Styx, but there are places along the Spree<br />

downtown where at night the water turns black and takes on a bleak,<br />

industrial sort of feeling; the mythological river might look that way if<br />

you transported it into the 21st century. Maj Britt, this book’s photographer,<br />

had found a good place along the river to take Sandro’s portrait,<br />

and he liked the idea. On the way into the city we talk about literature.<br />

At the moment he is rereading Candide, by Voltaire, and is<br />

working his way through the books of Sartre and Camus.<br />

He considered going into literature or linguistics, he says. He grew up<br />

in Grisons in Southeastern Switzerland – a breeding ground for polyglots,<br />

with three official languages. His father was a chef in a great<br />

199<br />

Mistic is a small membrane protein from bacteria which is built of four small<br />

helices that cross the membrane. Because the protein folds correctly in water<br />

and membranes, Sandro and his lab can use it to compare the process of<br />

folding in different environments.<br />

restaurant, Sandro says, who comes from the German-speaking part of<br />

Switzerland; his mother is a native of Austria, a kindergarten teacher,<br />

and a practitioner of extreme sports. “We spoke German at home and<br />

as a result I was the only kid who couldn’t speak Romansch when I<br />

went to kindergarten,” he says. “That changed quickly; by the fourth<br />

grade I had completely lost my German. We had it as a foreign language<br />

one hour a week, and I remember I couldn’t spell anything.”<br />

He took the “language track” in school, where he says no subject was<br />

boring. He couldn’t imagine focusing on only one thing. But he was<br />

eager to get out in the world, so he left school early, studied hard at<br />

home for a year and a half, and passed his school-leaving exams at the<br />

age of 18. After working as a meteorologist during his military service,<br />

he entered the university and committed himself to biophysical chemistry.<br />

That led to a PhD on the biophysics of membranes, and now his<br />

own group. Which now has nine members. More people means that<br />

the lab gets more things done, but it also costs time.<br />

“Coaching is important,” he says. “You try not to push, not to be too<br />

directive, but to be supportive.” He talks about a student who was<br />

struggling after six or seven months in the group. “He was getting<br />

frustrated because his project wasn’t working out. You have to sit<br />

down and help them learn to be patient. Not to lose the big picture.<br />

And to keep their eyes open. Sometimes when an experiment doesn’t<br />

work out, it’s still giving you an answer – just not to the question you<br />

asked.”<br />

Part Three: Frontiers and ferrymen


The electrician’s toolbox<br />

The earliest recorded cure for a migrane involves standing on a<br />

fish. It sounds like the beginning of a bad joke, but it’s true. The<br />

Torpedo, or electric ray, is a round, flat fish that lives on the sandy bottom<br />

of the ocean. It stuns its prey by delivering a massive electrical<br />

shock. This fact has been known since the ancient Greeks, although the<br />

first reference to the animal comes from philosophy rather than medicine.<br />

In Plato’s dialog on the nature of virtue, Meno says to Socrates,<br />

“You seem to me both in your appearance and in your power over others<br />

to be very like the flat torpedo fish, who torpifies those who come<br />

near him and touch him, as you have now torpified me, I think. For my<br />

soul and my tongue are really torpid, and I do not know how to answer<br />

you.”<br />

The first medical use of the fish comes from Scribonius Largus, court<br />

physician to the Roman emperor Claudius. Although he was not the<br />

emperor’s personal doctor (that post was held by Xenophon, who later<br />

poisoned Claudius with a feather), Scribonius was regarded highly<br />

enough to accompany Claudius on a campaign to Britain in 43 ce. Later<br />

he was asked to record his medical knowledge and wrote a book called<br />

the Compositiones, which contains 271 prescriptions. Its preface contains


Thomas Jentsch<br />

the earliest recorded reference to the<br />

Hippocratic Oath, in which doctors are<br />

advised to do no harm.<br />

One of the worst things a doctor can do,<br />

Largus writes, is not to give a drug when he<br />

knows it works. But one must be highly educated<br />

and critical, and Largus denounces<br />

some of the remedies proposed by other<br />

physicians. He points out that drinking the<br />

blood of a freshly killed gladiator, for example,<br />

will not cure epilepsy. Standing on a Torpedo,<br />

on the other hand, can cure the pain of gout<br />

in the feet (which become numb from the<br />

shocks). For migranes, a patient can immerse<br />

his hands in a basin of water containing one<br />

of the creatures.<br />

Some of Largus’ prescriptions may have<br />

worked well. For swollen tonsils he described<br />

a complex mixture containing:<br />

2 drachmas each of costus, celery seeds,<br />

anise seeds, oil of camel grass, and<br />

cinnamon-cassia, one-half drachma of<br />

cardamom, 2 drachmas of the wild rue<br />

(two-thirds of which is the seed), onehalf<br />

ounce of fissile alum, 5 mediumsized<br />

ground-up oak galls, 2 drachmas<br />

of saffron, one-half drachma of the<br />

refined residue of the oil of saffron,<br />

one-half drachma of myrrh, 4 drachmas<br />

of Cretan birthwort, 3 drachmas of<br />

cinnamon, one ounce of the ashes of<br />

a young wild swallow, and one-half<br />

drachma of spikenard


202<br />

Several of the ingredients contain biologically<br />

active compounds; the pharmacological<br />

screening library at the FMP (described in the<br />

next story) likely includes extracts from some<br />

of them. (Well, maybe not the ashes of a<br />

young wild swallow...)<br />

��<br />

The FMP doesn’t have a Torpedo, either,<br />

although a researcher who splits his<br />

time between the FMP and <strong>MDC</strong> is interested<br />

in the animal. For nearly 20 years Thomas<br />

Jentsch has been conducting a systematic<br />

study of an ancient and now widespread family<br />

of molecules called chloride channels.<br />

When they first evolved in one-celled organisms,<br />

their function was to sense changes in<br />

the environment and help the cell survive by<br />

adjusting its chemistry. With the arrival of<br />

multicellular life, they allowed cells to communicate<br />

with each other and coordinate the<br />

activities of the body. Dozens of types of chloride<br />

channels are known. Thomas’ work has<br />

focused on an important subset of these molecules,<br />

called CLCs, and other ion channels and<br />

transporters.<br />

“The human genome contains nine CLC chloride<br />

channels and transporters, all copies of a<br />

single gene that evolved long ago,” Thomas<br />

says. His laboratory found the first example<br />

over a decade ago and has systematically<br />

uncovered many more. “The extra copies arose<br />

because of errors that were made as DNA was<br />

replicated, and over time they have evolved<br />

different functions.”<br />

Many of these molecules float in the plasma<br />

membrane on the fringes of the cell, where<br />

they act as doormen. Their job is to control the<br />

flow of negatively-charged chloride atoms<br />

(ions) through the membrane. Sometimes<br />

this traffic of ions balances the overall charge<br />

of a cell with that of its environment. At other<br />

times an imbalance builds up, creating a positive<br />

charge on one side and a negative charge<br />

on the other. In between is the thin cell membrane,<br />

which suddenly becomes a superb conductor<br />

of electricity. In neurons, this causes an


Different ways that proteins of the CLC family assemble to form channels in membranes.<br />

electric charge to race huge distances down<br />

the surface to the far end of the cell, where it<br />

opens and closes channels. That causes a<br />

release of molecules which drift over to the<br />

next cell, open its channels, and start the<br />

process all over again.<br />

The movement of ions across the cell membrane<br />

is crucial to a wide range of cellular<br />

processes, for example moving salt and water<br />

into and out of the cell, or coordinating the<br />

activity of muscles and neurons. Some CLCs<br />

are lodged in membranes inside the cell. One<br />

of their jobs is to recycle cellular waste. But<br />

scientists still don’t have a full view of the<br />

functions of channels and transporters in the<br />

body’s tissues.<br />

The main method of finding out has been to<br />

delete genes one-by-one in model organisms<br />

like the mouse and then to closely watch the<br />

animals’ development and behavior. “Another<br />

approach has been to guess that a CLC gene<br />

might be mutated in certain human disorders,”<br />

Thomas says. “If we find CLC mutations<br />

in certain types of patients, then diseases can<br />

also provide powerful clues toward the normal<br />

function of the protein.”<br />

Over the last two decades Thomas has taken<br />

on the CLCs one-by-one to uncover their func-<br />

tions. Some of these projects have yielded<br />

insights into disease and have promising<br />

therapeutic implications. Thomas’ experiments<br />

have shown that removing one channel<br />

leads to deafness; when another is deleted,<br />

kidney stones form; yet another is needed<br />

to maintain the normal density of bones, and<br />

others seem to be connected to conditions<br />

like epilepsy.<br />

To understand why a defect in a chloride<br />

channel or transporter has these effects,<br />

Thomas says, you need to understand how it<br />

works in healthy cells, but that’s not easy. It<br />

requires a sophisticated array of techniques –<br />

from structural investigations of the building<br />

plans of the molecules, to their roles in mouse<br />

cells and human diseases. Thomas’ new home<br />

at the Timoféeff-Ressovsky building on the<br />

Berlin-Buch campus has been an ideal place<br />

to put together an interdisciplinary team.<br />

Some methods have been imported from his<br />

last post at the Center for Molecular<br />

Neurobiology at the University of Hamburg.<br />

Others are available at the FMP. He can also<br />

draw on resources from the <strong>MDC</strong>, his second<br />

employer.<br />

��<br />

203<br />

Part Three: Frontiers and ferrymen


Lena Wartosch<br />

Like many of Buch’s scientists, Thomas’<br />

interests extend to art and culture. His<br />

office is adorned with cult objects he<br />

obtained in Papua New Guinea: carvings of<br />

the heads of crocodiles and other creatures.<br />

They all have a triangular shape, which<br />

reflects their source: originally, each was an<br />

ornament carved into the prow of a boat.<br />

When a boat falls into disuse, Thomas says,<br />

the carvings are removed. Then, at least in<br />

some cases, they are sold to scientists from<br />

Berlin.<br />

They look over his shoulder with fierce expressions<br />

as he talks about his 20 years of<br />

research into membrane proteins.<br />

“Our cells contain an astounding variety of<br />

channels for the transport of ions,” he says.<br />

“There are over 60 different genes that<br />

encode channels to transport potassium ions.<br />

Why should there be so many molecules that<br />

have about the same function? The diverse<br />

roles of some of them have been discovered<br />

through studies of knockout animals, or<br />

through examination of tissues of patients<br />

suffering from diseases, but a great number<br />

remain puzzles. I believe that as they are<br />

investigated, we will uncover many more connections<br />

to disease.”<br />

In the late 1980s, when Thomas began studying<br />

the molecules, this work was in its infancy.<br />

Scientists knew that chloride channels had to<br />

exist but had not yet identified any genes<br />

encoding them. A good place to look, Thomas<br />

thought, was the Torpedo.<br />

His interest stemmed from the fact that<br />

Torpedo cells require an unusually high number<br />

of chloride channels to produce the massive<br />

electric discharge, which can reach levels<br />

Part Three: Frontiers and ferrymen<br />

204<br />

of 100 volts, with currents up to one ampere.<br />

Everyone knew the channels had to exist, but<br />

attempts to isolate a CLC from the fish’s cells<br />

had failed.<br />

“To find a protein you needed a molecule that<br />

would bind to it quite specifically, and we<br />

hadn’t found one for the chloride channels,”<br />

Thomas says. “So we took another approach;<br />

we decided to clone Torpedo RNAs that might<br />

contain the channel and put them into<br />

Xenopus egg cells – a frog which is widely<br />

used as a model organism.<br />

The genes of these eggs are inactive, but all<br />

the machinery is there to transform RNAs into<br />

proteins. The cell used the Torpedo RNAs to<br />

make proteins. By watching its electrical<br />

behavior, we could tell whether we had<br />

imported a molecule that functioned as a<br />

channel.”


This could be checked using a voltage clamp<br />

technique. In this method an electric current<br />

is “injected” into the egg and scientists measure<br />

the output – changes of voltage over the<br />

egg’s membrane. If the cell has extra chloride<br />

channels – because it has made them from<br />

Torpedo molecules – the charge will leak out<br />

through them, and more current needs to be<br />

injected into the cell to reach the same voltage.<br />

The measurements allowed Thomas and his<br />

colleagues in Hamburg to identify the first<br />

CLC and pin down specific regions of the protein<br />

that were responsible for opening the<br />

channel and moving ions. Over the next<br />

decade, Thomas and his laboratory identified<br />

several new CLCs and began removing them<br />

one-by-one from mice to study their functions.<br />

The work has revealed numerous connections<br />

between defects in the channels and<br />

disease. In several cases, the scientists first<br />

discovered patients who carried mutations in<br />

the genes, and then later created mouse models<br />

of the diseases.<br />

For example, a lack of ClC-1 leads to a condition<br />

known as inherited myotonia. A person’s<br />

muscles don’t relax after voluntary activity<br />

like shaking someone’s hand or squinting in<br />

bright sunlight. The condition was first<br />

described in 1876 by Danish physician Julius<br />

Thomsen, who suffered from it himself, as did<br />

several members of his family. It took more<br />

than a century, but Thomas and his colleagues<br />

finally traced the disease to the chloride channel<br />

gene and even pinpointed the specific<br />

mutation that caused the disease in Dr.<br />

Thomsen himself.<br />

“We think that by allowing chloride ions to<br />

flow into the cell, ClC-1 helps muscles relax<br />

205 Part Three: Frontiers and ferrymen<br />

and ‘reset’ so that they are ready for the next<br />

stimulation,” he says. “Without the molecule,<br />

external potassium levels build up and cells<br />

begin to stimulate themselves. This leads to a<br />

sort of seizure in which muscles no longer<br />

relax.”


It wasn’t hard to imagine something similar<br />

happening in other tissues. Thomas was<br />

intrigued by a report from a clinical group<br />

which had discovered mutations in ClC-2, a<br />

version of the channel found in the brain and<br />

other organs. The family that had the problem<br />

suffered from epilepsy. Was the loss of channel<br />

proteins causing an overstimulation of<br />

neurons, leading to brain seizures? He was<br />

skeptical. A couple of years earlier, his lab had<br />

developed a strain of mouse without ClC-2.<br />

Although mice can suffer from epilepsy, the<br />

knock-out animals didn’t. Instead, they experienced<br />

degeneration of the testes and the<br />

retina.<br />

Judith Blanz, a student in Thomas’ group at<br />

Hamburg, decided to conduct a more careful<br />

examination of the nervous system of mice<br />

lacking this chloride channel. She discovered<br />

that while individual neurons still had their<br />

normal forms, the animals’ brains and spines<br />

developed severe defects. “They took on a<br />

sponge-like appearance because of holes that<br />

formed in the white matter of the brain,”<br />

Thomas says. “The only real behavioral difference<br />

was related to blindness, caused by the<br />

retinal degeneration we had noticed in the<br />

previous study. And the mice responded<br />

somewhat more slowly to stimuli, caused by<br />

holes in the white matter. Signals in the brain<br />

were not being transmitted as quickly.”<br />

The sponge-like gaps in the brain were familiar;<br />

they also appear in a human disease<br />

called leukodystrophy. Researchers had<br />

Part Three: Frontiers and ferrymen<br />

206<br />

already linked several genes to different forms<br />

of human leukodystrophy, but some families<br />

suffering from the disease didn’t have mutations<br />

in any of these genes. So other, unknown<br />

genes might be involved, including ClC-2. The<br />

scientists obtained DNA from a group of 150<br />

leukodystrophy patients and began looking<br />

for a pattern of mutations. “We didn’t find<br />

them,” Thomas says. “That was disappointing;<br />

on the other hand, the damage in the brains<br />

of humans who suffer from this condition is<br />

mostly more severe than what we observe in<br />

the mice. We may have picked the wrong subset<br />

of patients; we don’t know of human<br />

patients where the symptoms are combined<br />

with degeneration of the retina. Such a case<br />

would more closely resemble what we see in


the mouse. We are currently collaborating<br />

with groups in human genetics to try to find<br />

such patients.”<br />

��<br />

Amutation’s effects on health are often<br />

strange and unpredictable. While looking<br />

at a potassium channel called KCNQ2, the<br />

scientists found mutations which cause<br />

epilepsy in human newborns. The discovery<br />

suggested that KCNQ proteins would make a<br />

good target for anti-epileptic drugs. Later it<br />

was found that a known substance, currently<br />

in phase III clinical trials for use in treatments<br />

of epilepsy, had its effects by docking onto<br />

KCNQ potassium channels. Pharmaceutical<br />

companies are now trying to develop<br />

207


more specific anti-epileptic drugs by targeting<br />

the channels. It’s another example of the<br />

beneficial side-effects of basic research,<br />

Thomas says: finding a gene responsible for a<br />

very rare genetic condition can sometimes<br />

lead to therapies for much more common<br />

diseases.<br />

Another serendipitous discovery arose from<br />

studies of mice that lacked ClC-7. “Removing<br />

this gene leads to unusually thick and brittle<br />

bones in mice,” Thomas says. “So we had a<br />

look at human bone diseases. We found a<br />

defective form of the gene in some people<br />

who suffer from osteopetrosis, a rare bone<br />

disease also characterized by thick bones. One<br />

aspect of this study is that it may give us hints<br />

about how to treat a much more common<br />

disease – osteoporosis, in which bones<br />

become too thin, particularly in elderly<br />

women.”<br />

What’s the connection between thick bones in<br />

mice and thin ones in aging humans? “ClC-7 is<br />

active in osteoclasts, cells which degrade<br />

bones,” he explains. “You wouldn’t want that<br />

degradation to happen in elderly people, especially<br />

women. A drug that partially inhibited<br />

ClC-7 might help preserve bone density. And<br />

that could make a treatment for osteoporosis.”<br />

��<br />

Recently the lab has extended its attention<br />

to a family of molecules called KCCs<br />

that transport both chloride and potassium in<br />

a coupled way. “We call these molecules cotransporters,<br />

and they were first discovered in<br />

studies of red blood cells,” Thomas says. “As<br />

these cells are squeezed through tiny capillaries,<br />

and have to go through cycles of oxygenation<br />

which are associated with fluxes of salt<br />

and water, they have to precisely regulate<br />

their volume. At least two types of KCCs help<br />

carry out this job in the membrane of red<br />

blood cells.”<br />

A channel is relatively passive – allowing ions<br />

to pass “downstream”, for example by shipping<br />

negatively charged particles into an area<br />

where the charge is positive. But transporters<br />

like KCCs are able to push chloride ions<br />

“uphill” – by moving them at the same time as<br />

potassium ions.<br />

In general, Thomas says, the KCCs are mainly<br />

responsible for the outward passage of chloride<br />

and fluid from the cell. Seth Alper of the<br />

Harvard Medical School, who has been collaborating<br />

with Thomas on the KCC project,<br />

showed that there is too much transport of<br />

potassium and chloride out of the cell during<br />

sickle cell anemia. In this disease, red blood<br />

cells lose too much volume and take on a<br />

deflated, sickled shape. The cells then become<br />

wedged in tiny capillaries, obstructing the<br />

flow of blood and causing irreversible damage<br />

to organs. A mutation in sickle cell anemia<br />

causes hemoglobin proteins to form clumps<br />

that do not dissolve inside the cell when<br />

things get too crowded – and that happens<br />

when red blood cells can’t control their volume<br />

properly.<br />

Seth has done a great deal of work on the<br />

relationship between hemoglobin, cell vol-<br />

Part Three: Frontiers and ferrymen<br />

208<br />

ume, and the symptoms of the disease.<br />

“Clumps form when there is a high concentration<br />

of mutant hemoglobin in the cell,” Seth<br />

says. “This somehow causes cells to take on<br />

their odd, sickle shape. If we could prevent the<br />

aggregations, we could stop their harmful<br />

effects. That means keeping the concentration<br />

of hemoglobin low, and having enough<br />

water in the cell is an important part of the<br />

process. If KCC channels are too active, the cell<br />

loses ions and water. If that didn’t happen, we<br />

think the course of the disease would be<br />

greatly slowed.”<br />

It sounded good, but proving the hypothesis<br />

would require studying KCCs in the blood<br />

more carefully. Thomas and his colleagues<br />

from Hamburg developed strains of mice<br />

lacking KCC1 and KCC3, the two types of<br />

potassium-chloride transporters found in<br />

red blood cells. Removing KCC1 had no obvious<br />

effects on the mice. But taking away<br />

KCC3, or both molecules, reduces the transport<br />

of chloride and potassium ions out of the<br />

cell.


Would this offer the animals protection from<br />

sickle-cell disease? The scientists crossed the<br />

mice with another strain which has defects in<br />

hemoglobin and develops a condition very<br />

much like human sicke-cell anemia. “On the<br />

whole,” Thomas says, “the blood cells of these<br />

animals had more volume, and there was less<br />

ion transport. But it didn’t help the cells that<br />

were most dehydrated and deformed. One<br />

hypothesis is that this subpopulation of cells<br />

lose their water a different way, earlier in the<br />

process by which red blood differentiates.<br />

Maybe a combination of inhibiting KCCs and<br />

other ion transporters like potassium channels<br />

could help.”<br />

��<br />

Currently Thomas is most excited about<br />

his work with chloride transporters at<br />

work inside the cell, including other members<br />

of the CLC family. “These molecules are found<br />

in membranes along the endocytotic pathway,”<br />

Thomas says. “That’s a route used by the<br />

cell to absorb molecules from the outside, to<br />

capture toxins and other foreign substances,<br />

Ioana Neagoe<br />

and to recycle proteins to the outer membrane<br />

or to degrade them.”<br />

Substances are wrapped in membranes so<br />

that they can be processed. These membrane<br />

bubbles are absorbed by internal cellular compartments<br />

that contain high levels of acid and<br />

enzymes that break things down. Raising the<br />

acid level means bringing high numbers of<br />

protons into the compartments. Vesicles are<br />

equipped with proton pump molecules that<br />

do this, but they can’t do the whole job – once<br />

the acidity reaches a certain level, the pump<br />

can no longer push protons “upstream” into<br />

the compartment. “The solution is to neutralize<br />

the positive charge going in by importing<br />

an equal negative charge,” Thomas says. “That<br />

happens through the CLC transporter, and it<br />

allows the vesicles to continue to increase<br />

their acidity.<br />

“A whole branch of the CLC family – five molecules<br />

in humans – are found mostly in internal<br />

membranes, and we didn’t know what<br />

they were doing,” he says. “In 2005, we<br />

showed that ClC-4 and ClC-5 were also acting<br />

209 Part Three: Frontiers and ferrymen<br />

as coupled transporters, and not as channels –<br />

inside the cell.”<br />

Here, too, there were links to disease: ten<br />

years ago, Raj Thakker in London and Thomas<br />

had found that mutations that inactivate ClC-<br />

5 cause Dent’s disease, an inherited disorder<br />

that causes people to develop kidney stones.<br />

Nils Piwon, a graduate student in Thomas’ lab,<br />

then generated a mouse without ClC-5 and<br />

found that the animals’ kidneys were severely<br />

impaired in their ability to take up proteins,<br />

which then entered the urine. The scientists<br />

developed a hypothesis to explain why this<br />

defect in protein uptake causes kidney stones:<br />

without the ClC-5 chloride transporter, kidney<br />

cells are unable to absorb another small protein,<br />

a hormone called PTH. It therefore accumulates<br />

at too high levels in the primary<br />

urine, triggering abnormal signals in the cell<br />

that limit the reabsorption of phosphate, .<br />

Through a complex series of signaling events,<br />

this also increases vitamin D levels. This molecule,<br />

along with PTH, is the main hormone in<br />

regulating body calcium. Patients with Dent’s<br />

disease take in too much calcium from food,<br />

which the body has to get rid of through the<br />

kidney. The combination of high calcium and<br />

high phosphate in the urine then leads to precipitation<br />

of calcium crystals that finally build<br />

up to the terribly painful kidney stones.<br />

Thomas lists other diseases linked to internal<br />

CLCs: neurodegeneration associated with ClC-<br />

3, lysosomal storage disease (ClC-6) and now<br />

osteopetrosis (ClC-7). The genome encodes 60<br />

potassium channels alone, I remember. Are all<br />

of them linked to disease?<br />

You can tell that Thomas isn’t finished, that<br />

he’s somewhere in the middle of the research<br />

roadmap that he designed 20 years ago. It’s<br />

still a good plan and he talks about it eagerly.<br />

Behind him the lab is filling with new methods,<br />

postdocs, students, and technicians.<br />

There is still an immense amount to do with<br />

ion channels. He’s building a laboratory for<br />

the next 20 years.


Hartmut Oschkinat<br />

A kiss, interrupted<br />

It’s probably dangerous to compare a pair of proteins to a sculpture,<br />

especially one that has been called a “sensual masterpiece,” a “mixture<br />

of idealism and eroticism” – Auguste Rodin’s kissing lovers sit on a<br />

rock that is only half-hewn from the marble block; above, their glossy<br />

bodies entwine to bring their lips together. But having this image in<br />

mind might be helpful in imagining the encounters between molecules<br />

that drive processes within cells. Proteins also kiss, sometimes in brief<br />

encounters that last for picoseconds. They enjoy a fleeting contact that<br />

changes one or both molecules, transmitting information, loading each<br />

other with energy, or carrying out other types of transformations.<br />

Sometimes there is only a small peck on the cheek, precisely and neatly<br />

delivered. In other cases, the molecules fold and wrap around each other.<br />

Many of the encounters in the cell are the pecking type; they take place<br />

between a tiny partner such as a hormone and a much larger protein.<br />

Others involve whole-body embraces between two proteins, huge surfaces<br />

that scientists have had trouble understanding and manipulating.<br />

Achieving this understanding is important, says Hartmut<br />

Oschkinat of the FMP, if researchers want to find drugs that control<br />

interactions between proteins.<br />

211 Part Three: Frontiers and ferrymen


Peter Schmieder and Barth Rossum<br />

The active components of drugs are often small molecules that have<br />

been extracted from much larger substances and then purified, leaving<br />

only the pair of lips – only what is needed to deliver a kiss or block<br />

someone else’s attempt to do so. Often they snap into a pocket or a<br />

groove in one of the cell’s proteins, taking the place of another molecule<br />

that is somehow linked to the symptoms of a disease. If<br />

researchers find such a pocket, it may not be hard to find or design<br />

something else that fits. That doesn’t mean it will be easy to turn their<br />

invention into a drug, for a variety of reasons. Even so, it’s easier than<br />

blocking contacts between large surfaces.<br />

“Such interactions are crucial in most cellular processes,” Hartmut<br />

says. “Finding ways to inhibit them is usually very challenging. In the<br />

first place, these surfaces are large and flat, and it’s hard for a small<br />

Part Three: Frontiers and ferrymen<br />

molecule to get a grip – the way that you can’t free-climb on a surface<br />

that’s too smooth. And there are very few natural small molecules that<br />

bind to these surfaces, which would also give you a place to start in<br />

designing new drugs.”<br />

That’s not to say that it can’t be done; some existing drugs work on<br />

larger surfaces. Taxol, which is widely used to fight lung and breast<br />

cancer, was isolated from the bark of Pacific yew trees growing near<br />

the Mount St Helens volcano in Washington. It latches onto a protein<br />

called tubulin and prevents this molecule from binding to its partners,<br />

thus triggering a cellular self-destruct program called apoptosis.<br />

Another drug, cyclosporine A, was derived from a fungus that thrives<br />

in the soil in Norway. It has been widely used for over 20 years as an<br />

immune system suppressor to reduce inflammations and prevent<br />

212


organ rejections. It works by blocking interactions between proteins in<br />

T cells, reducing their ability to properly stimulate an immune<br />

response.<br />

“These examples show that in principle, protein-protein interactions<br />

can be inhibited, and the molecules that do so might make powerful<br />

drugs,” Hartmut says. “But we lack a lot of basic knowledge about the<br />

interactions. I thought that if we could find the right pair of proteins,<br />

it might give us a good model system to use to get a handle on the<br />

problem.”<br />

��<br />

The publication of the human DNA sequence and other complete<br />

genomes has given scientists a full catalogue of proteins and<br />

other molecules that can be produced by our cells. A lot of them are<br />

cousins; long ago they started off as one gene in one of our ancestors.<br />

Then duplicates arose through errors that were made as DNA was<br />

copied. Some of them were preserved through evolution, gradually<br />

undergoing mutations and taking on different functions. One structure,<br />

called a PDZ domain, has now been found in 250 proteins.<br />

“This is one of the most important modules in cells that govern interactions<br />

between proteins,” Hartmut says. “It helps organize networks<br />

of signaling proteins; it participates in the delivery of parts to ‘molecular<br />

machines,’ and one of its most important jobs is to control events<br />

at the cell membrane. We began looking at this module because of its<br />

importance, and also because of a feature found on its surface. PDZ<br />

domains contain a shallow groove – not a real pocket, but something<br />

that an inhibitor might get a grip on.”<br />

Mangesh Joshi, a PhD student in the group, took on the project and<br />

focused on a protein called AF6, which contains a PDZ module. One<br />

reason for the choice was the role of AF6 in the development of a disease<br />

called chronic myeloid leukemia. This type of cancer arises<br />

through a flaw in another protein, called BCR, which binds to the PDZ<br />

domain in AF6. Recently Yunyu Shi’s laboratory at the University of<br />

Science and Technology in Hefei, China, discovered that BCR docks<br />

onto just the surface groove that Mangesh intended to look at.<br />

Each PDZ domain docks onto a specific module in one other protein.<br />

Ten years ago, Hartmut’s group helped figure out how PDZs recognized<br />

the right targets: they latch onto a loose, tail region of a domain<br />

in the partner molecule. Because the amino acid recipe of each tail is<br />

slightly different, it has different binding properties. Evolution has tailored<br />

a specific tail to fit the groove of a particular PDZ.<br />

“Because of what we know about how these molecules bind, it<br />

seemed likely that plugging the groove with another molecule might<br />

block access of the normal partner,” Mangesh says. “But to show that<br />

to be the case, we had to find an inhibitor.”<br />

��<br />

Since 2002 the FMP has been creating a “library” of molecules –<br />

ranging from natural substances to small artificial peptides – that<br />

can be used in projects like the one planned by Hartmut and Mangesh.<br />

The library is housed in a high-tech screening facility in the newly constructed<br />

Genomics center, a sleek, black building on the east edge of<br />

the campus. The building was opened in 2006 by the FMP and their<br />

campus neighbor, the Max-Delbrück Center for Molecular Medicine.<br />

Jens Peter von Kries heads the facility. His office has large windows<br />

that look over the campus, which he knows well; from 1995-2000 he<br />

was a scientist in Walter Birchmeier’s group at the <strong>MDC</strong>. He spent the<br />

next two years of his career as the scientific head of the screening unit<br />

213 Part Three: Frontiers and ferrymen


for a Berlin pharmaceutical company before<br />

being recruited by the FMP to set up a similar<br />

facility. Today his foot is in a temporary cast<br />

from a football accident that happened while<br />

he played with his son. He places it on his<br />

desk like a paperweight as he talks about the<br />

mission of the facility.<br />

“Our aim is to help scientists search for<br />

inhibitors for specific molecules and processes,”<br />

Jens says. “The point may be to find a lead<br />

compound that can be developed into a drug<br />

– or the goal may be to develop a new tool to<br />

study a cellular process.”<br />

Sometimes, as in the case of Mangesh’s project,<br />

the role of the facility is simply to open its<br />

library for scientists to use. Mangesh had<br />

decided to carry out the screen itself in<br />

another part of the institute, using the FMP’s<br />

NMR machines. Hartmut’s group has helped<br />

turn NMR into the institute’s major tool used<br />

in studying protein structures, and it also has<br />

other applications such as screening.<br />

“There were two reasons to do the screen<br />

with NMR,” Mangesh says. “First, because of<br />

the nature of the PDZ binding region, we<br />

anticipated that our first ‘hits’ might bind<br />

quite weakly. NMR is good at detecting weak<br />

interactions. Secondly, it would give us a<br />

detailed look at the structures of the molecules<br />

as they interacted. That would be<br />

important for the second stage of the project:<br />

taking a substance and rebuilding it so<br />

that it would bind better.”<br />

At other times the screening facility takes a<br />

much more active role in projects, offering a<br />

full range of screening services and assistance<br />

in chemistry for in-house researchers<br />

and visitors. Compounds from the library are<br />

incubated with samples, either purified proteins<br />

or live cells, and then examined to see if<br />

there have been effects. All of this work used<br />

to be done by hand, but the entire process<br />

has been automated. The day of our tour,<br />

final tests were being run on a new automated<br />

screening machine, so massive that it<br />

takes up half of a lab. Packed inside are robots


that handle samples in a “hotel”, driving their containers to a central<br />

corridor where they encounter substances from a large library of<br />

chemical compounds.<br />

The FMP’s library currently includes about 50,000 substances, mostly<br />

purchased from other collections. “Some major pharmaceutical companies<br />

have collections of millions of substances,” Jens says.<br />

“Obviously the FMP couldn’t afford to build a library of that size, so we<br />

had to be selective. That was an intensive process – making decisions<br />

on the best representatives of different classes of molecules, so that<br />

we would have a good spectrum of substances to work with.”<br />

FMP Director Walter Rosenthal says the screening unit is necessary to<br />

help the institute and the wider research community carry out better<br />

research as well as fill in some of the gaps in the drug discovery<br />

process. “The point is to take a tool that is commonly used in the pharmaceutical<br />

industry and give academic scientists access to it,” he says.<br />

“We’re not trying to compete with pharmaceutical companies – they<br />

are much better equipped to ‘try everything’ in the search for new<br />

drugs. The point is that academic researchers need access to a screening<br />

platform, the way they have access to other mass technologies.<br />

Here we’ve built a platform – unique to Germany and Europe – that<br />

they can use; we’re making it as open as possible. Your everyday scientist<br />

hasn’t had that kind of access because the price of having a screen<br />

done may be prohibitive, or a major pharmaceutical company may<br />

simply not be interested in a particular project. Academic researchers<br />

may need an inhibitor simply to study a process that they’re interested<br />

in, which is a much too basic problem for most companies.”<br />

If a screen turns up a hit, researchers can turn to the FMP’s chemical<br />

groups for help in improving it. New groups such as that of Jörg<br />

Rademann have been brought in to strengthen its chemistry program.<br />

“The farther we make it along the process of trying, testing, and<br />

improving an inhibitor that works in a disease process,” Jens says, “the<br />

more likely it is that a company will take interest and move it into<br />

development and clinical trials.<br />

There have already been some interesting success stories. One project<br />

of the unit involves Mycobacterium tuberculosis (Mtb), the organism<br />

that causes the dreaded lung disease, and is a good example of how<br />

modern molecular science is being applied to infectious diseases. The<br />

incidence of tuberculosis is on the rise – epidemiologists estimate that<br />

one-third of the world’s population is infected, although only a small<br />

percentage show symptoms – and about 20 percent of the bacteria<br />

have developed resistance to the most common anti-tuberculosis<br />

drugs. So national and international research campaigns have<br />

branched out, searching for new ways to fight the disease.<br />

Tuberculosis projects across the world have been looking for good<br />

drug targets among the molecules in the tuberculosis bacterium. The<br />

ideal target would be a protein with unique characteristics, which<br />

might allow it to be blocked without interfering with one of the<br />

healthy proteins in the cell. Recently Jens and Larissa Podust, now at<br />

the University of California, San Francisco, carried out a screen to find<br />

inhibitors for a protein called CYP51. This molecule is found in all kinds<br />

of cells, ranging from bacteria to humans, where one of its jobs is to<br />

synthesize cholesterol. That is vital to creating membranes, and without<br />

it the bacteria cannot survive. The version of CYP51 found in bacteria<br />

is different than the human form, with unique features that might<br />

make a good drug target.<br />

“Another advantage is that when CYP51 binds to another molecule, it<br />

undergoes a change,” Jens says. “You can observe it with a spectrometer,<br />

so it gave us a good method to observe the results of the screen.”<br />

215 Part Three: Frontiers and ferrymen


A region of the AF6 protein called the PDZ domain changes when a small<br />

binding partner attaches itself. This suggests that the site can be used as<br />

a target for the development of small molecules or substances that<br />

inhibit the activity of the many proteins that contain PDZ domains.<br />

Part Three: Frontiers and ferrymen<br />

High-throughput experiments turned up three substances that<br />

bound to CYP51, including a known antifungal drug called EPBA. It<br />

locked onto the bacterial protein strongly and very specifically – not<br />

disturbing other molecules. EPBA is also used widely as a preservative<br />

in cosmetics, which means that its effects on humans have been carefully<br />

studied. If EPBA does turn out to be effective against tuberculosis,<br />

that will save time when it has to be tested on patients.<br />

The scientists discovered that EPBA strongly resembled the structure<br />

of another molecule, called BSPPA, which they also tested; the second<br />

compound behaved exactly the same way.<br />

“Larissa Podust went to the synchrotron at the Argonne National<br />

Laboratory in Illinois, in the US,” Jens says. “The purpose of the trip was<br />

to use X-rays to get detailed structural pictures of how both of these<br />

molecules dock onto CYP51. The images show us how the compounds<br />

might be adapted to target other types of CYP51s. We’re hopeful that<br />

EPBA can be used as a structural scaffold to spin off specific new substances<br />

to do that. The fact that the molecule has already been<br />

through clinical trials in another context – as an antifungal drug –<br />

could dramatically speed up its development into an anti-tuberculosis<br />

therapy.”<br />

The collaboration with a lab in the United States shows that the facility<br />

is ready to take on long-distance partnerships in interesting projects.<br />

Jens is already working closely with a network in Scandinavia<br />

called ChemBioNet, and other collaborations are underway with<br />

Oxford University.<br />

One ongoing project involves a group at the University of Oslo,<br />

Norway, that works on specialized types of stem cells. Since the screen<br />

involves live cells, they had to be shipped to Berlin from Scandinavia.<br />

It’s not as far as Illinois, but in this case, distance proved to be a<br />

problem.<br />

216


“Jo Waaler tried to send the cells three times using express shipping<br />

services,” Jens says. “Each time they got hung up somewhere at the<br />

border, and the cells died. Finally Jo gave up, packed the cells in a rental<br />

car, and drove them down himself. All the way from Oslo.”<br />

This time the cells – and the project – survived.<br />

��<br />

When Mangesh began his screen on the PDZ module of AF6, no<br />

synthetic binding partners were known. He began working<br />

with a subset of the FMP’s collection, 5,000 small molecules particularly<br />

suited for screening with NMR. This included representatives of<br />

all the basic kinds of molecules in the collection, including the types<br />

most frequently found in drugs.<br />

“There are three critical stages for the design of inhibitors in drug discovery,”<br />

Mangesh says. “First you try to find lead compounds – these<br />

are substances that show moderate activity toward the target. Then<br />

you go to the test tube and try to optimize the compound. Here is<br />

where it pays off to know how the molecule binds to its target, to have<br />

this structural pictures. That points out regions that are crucial to the<br />

interaction. Finally, the system has to be put to work in cells and organisms.”<br />

The screen turned up three different types of compounds that could<br />

bind to the PDZ domain of AF6. The scientists chose the one with the<br />

strongest effect, 2-thioxo-4-thazolidinone, for further work. Carolyn<br />

Vargas, another student in Hartmut’s group, began systematically synthesizing<br />

and purifying new spin-off compounds from it. To do so she<br />

had help from Volker Hagen, head of the Synthetic Organic Chemistry<br />

group at the FMP.<br />

The structural pictures from the NMR screen revealed specific points<br />

that might help strengthen the bonds between the compound and<br />

AF6. After some work, Carolyn and Volker had come up with a new sub-<br />

stance with the awe-inspiring name (2R,5R)-2-sulfanyl-5-[4-(trifluoromethyl)benzyl]-1,3-thiazol-4-one.<br />

It’s not the kind of thing you would<br />

want to have to say in the lab every day, or spell for someone on the<br />

telephone, so the compound was nicknamed 7i. Further NMR experiments<br />

showed that 7i was about as good at binding to the PDZ<br />

domain as AF6’s natural partners in the cell.<br />

“Although the AF6 PDZ itself is highly similar to other PDZ domains,”<br />

Carolyn says, “the way that it bound to the new compound was surprising.<br />

Somehow 7i had docked itself into a subpocket of the PDZ<br />

domain that had never been seen before, in any structural picture of a<br />

PDZ. Maybe the pocket wasn’t even there before. We think this means<br />

that as 7i and the PDZ come into contact, the domain undergoes<br />

rearrangements that create the new pocket.”<br />

You’ll never see a statue do that – not even one as sensual as Rodin’s<br />

carving of a kiss – but proteins are anything but static objects. Their<br />

highly dynamic nature sometimes makes it hard to design a new drug,<br />

or predict how a substance will affect a protein. But if AF6 really<br />

undergoes a structural change, it might simplify things.<br />

“When we originally thought about finding a way to make drugs to<br />

disturb interactions between proteins,” Hartmut says, “we chose the<br />

PDZ domain because of its shallow groove. Now it appears that a class<br />

of molecules like 7i can create new pockets that allow them to get a<br />

better grip. This makes PDZs even more promising candidates for<br />

manipulations by drugs – especially if it turns out that these structural<br />

rearrangements are a general characteristic of PDZ domains, that<br />

they also happen in modules of other proteins. That’s likely to be the<br />

case; if so, it may give us a new strategy to design drugs that can influence<br />

a huge range of important cellular molecules.”<br />

217 Part Three: Frontiers and ferrymen


Friedrich Luft


Interlude:<br />

The case of the<br />

short-fingered musketeer<br />

This story could easily be made into a book of its own. It has all the<br />

elements of a good adventure: a scientific mystery that dates<br />

back three decades and then some, travels to remote regions on four<br />

continents, a strange mutation that causes disease in humans, an<br />

attack by wild boars, and some gunfire. And a physician became a<br />

guinea pig on his own operating table. There was almost a jailbreak,<br />

but it was made unnecessary by a last-minute reprieve.<br />

A word of warning from the outset: most of the stories in this book<br />

have a conclusion, but this particular case has not yet been solved. A<br />

solution could be just around the corner thanks to the recent surpris-<br />

ing discovery of a new type of code written into the human genome.<br />

On the other hand, the new code may be a false lead, and the real cul-<br />

prit may be a biological process that has not yet been identified. In the<br />

end the case may stay open in the files, waiting for a new generation to<br />

come along with a fresh perspective and new technology.<br />

��<br />

219 Interlude: The case of the short-fingered musketeer


Franz Volhard<br />

Nearly every day until the winter of 2007<br />

you could see Friedrich Luft on the<br />

Lindenberger Weg, the main road that connects<br />

the scientific campus to the former hospital<br />

complex that included the clinics of the<br />

Charité on the other side of Berlin-Buch. He<br />

was hard to miss – you looked for the only person<br />

with salt-and-pepper hair, dressed in<br />

white doctor’s garb, riding a bicycle. In 2007<br />

he still had an office on both sides of town.<br />

The one on the <strong>MDC</strong> campus is at the top of<br />

the Helmholtz house, with large windows<br />

overlooking the campus. He always takes the<br />

stairs. If he sees students waiting for the elevator,<br />

he’ll tease them: “What, are the stairs<br />

broken again?”<br />

Up on the fourth floor he’s planning the new<br />

Experimental and Clinical Research Center<br />

(ECRC), a project which has taken up a lot of<br />

his time over the past few years. So far it’s a<br />

virtual center, of which he is director. If things<br />

go as planned, the ECRC will soon take on<br />

Interlude: The case of the short-fingered musketeer<br />

physical form as a new building on campus.<br />

More on that topic later.<br />

For our first meeting Friedrich invited me to<br />

his office on the other side of Berlin-Buch, the<br />

hospital grounds known as Area 1, before the<br />

the Franz Volhard Clinic made its move to the<br />

Helios complex. His role there was head clinician,<br />

devoted to nephrology (the study and<br />

treatment of kidney diseases) and cardiovascular<br />

disease that arises from hypertension.<br />

As I walked over from the bus stop, he zoomed<br />

by. “Get a bicycle,” he yelled.<br />

The corridor to Friedrich’s office in the clinic<br />

bore portraits of many of the personalities<br />

who had worked there, and one who was<br />

present solely in spirit: Franz Volhard as a<br />

young man, holding a violin. The photograph<br />

was taken in 1893 while Volhard was a student<br />

at the University of Halle. He maintained<br />

a life-long love of music; throughout his<br />

career, as he moved from clinic to clinic, he<br />

encouraged his staff to brush up their musical<br />

220<br />

skills and perform together. The tradition was<br />

carried on in the clinic that bore his name;<br />

until the day of the move, the Volhard Clinic<br />

had a Bechstein piano in the lobby.<br />

“Franz Volhard was the preeminent German<br />

cardiovascular clinician of the early 20th century,”<br />

Friedrich says. “His many contributions<br />

concerned the kidney and the heart, oftentimes<br />

both. Nephrologists and cardiologists<br />

can equally lay claim to him as a role model<br />

and therefore the Franz Volhard Clinic was<br />

named after him.”<br />

Early in his career Volhard became interested<br />

in the fact that kidney diseases were almost<br />

always accompanied by hypertension – chronically<br />

elevated blood pressure. He was particularly<br />

intrigued by a condition called Bright’s<br />

disease, named after Richard Bright, the<br />

physician who discovered chronic kidney disease<br />

at Guys Hospital in London in 1827. The<br />

illness was puzzling because it affected several<br />

bodily systems: the kidneys became dysfunctional<br />

and shrank; blood pressure<br />

increased; the heart became enlarged;<br />

patients suffered brain damage and inflammations<br />

in the retinas of their eyes. Volhard<br />

believed that all of the symptoms might be<br />

due to an more basic problem related to<br />

hypertension, and he set off on a search for<br />

the causes of high blood pressure. While he<br />

didn’t succeed, Volhard and his students managed<br />

to convince the medical world that<br />

hypertension was a systemic problem linking<br />

some types of kidney and cardiac disease.<br />

Researchers at the Volhard Clinic – including<br />

Friedrich Luft – continue to pursue the underlying<br />

causes. Friedrich is particularly interested<br />

in essential hypertension, and I ask him<br />

what this means.<br />

“Essential hypertension is what everybody<br />

has – it occurs in 25 percent of the entire population.<br />

The name comes from a hypothesis of<br />

fifty years ago. At that time, it was believed<br />

that if blood vessels became constricted<br />

somewhere in the body, likely the kidneys, it<br />

would be ‘essential’ to raise pressure so that


lood could still be squirted through the<br />

arteries. It was a natural conclusion. But after<br />

all these years of research, we still haven’t<br />

pinned down the true causes of the kind of<br />

high blood pressure that affects individual<br />

patients and the general population. Multiple<br />

factors are probably responsible.”<br />

In time, visionaries such as Franz Volhard recognized<br />

that hypertension was not “essential”<br />

– but rather a major health menace. The<br />

name, however, stuck.<br />

A few forms of hypertension are simple inherited<br />

conditions that run through families, due<br />

to defects in individual genes that are passed<br />

from parent to child. This type of hypertension<br />

is inherited according to the patterns<br />

first described by Gregor Mendel in his studies<br />

of peas. In these families, a single genetic<br />

mechanism is responsible for a trait – in this<br />

case high blood pressure. Studying those conditions<br />

and making models of them in mice<br />

have given insights into each particular dis-<br />

ease. But most of these situations are different<br />

from essential hypertension, where many<br />

genes – rather than just one – are likely to be<br />

responsible.<br />

“The genes in these cases are involved in the<br />

reclamation or reabsorption of sodium and<br />

chloride in the kidneys,” Friedrich says.<br />

“They’re all related to people’s intake of salt –<br />

you see some pretty drastic changes in blood<br />

pressure among these patients, depending on<br />

the salt in their diet. However, these mechanisms<br />

don’t seem be responsible for essential<br />

hypertension. The majority of people who suffer<br />

from it do experience an increase in blood<br />

pressure if they eat more salt, but it’s not<br />

much.”<br />

��<br />

William Bateson, a classical geneticist of<br />

the early 20th century, advised his disciples<br />

to “Treasure your exceptions.” Bateson<br />

was referring to unusual discoveries in plants,<br />

221 Interlude: The case of the short-fingered musketeer<br />

but the same is true of exceptional families<br />

harboring mutations that lead to disease. If<br />

“classical” disease hypertension genes didn’t<br />

say much about blood pressure in the general<br />

population, Friedrich thought there might be<br />

other rare Mendelian genetic syndromes that<br />

did. “When I first arrived here in Berlin-Buch, I<br />

was supposed to be working on genetics and<br />

cardiovascular disease – but I could barely<br />

spell genetics. Thomas Wienker, a real geneticist<br />

who knew all about William Bateson,<br />

showed me an interesting paper that had<br />

been published in 1973 in the Journal of<br />

Medical Genetics. The topic was an extended<br />

family living in Turkey, in the region around<br />

the Eastern Black Sea coast. Members of the<br />

family suffered from an unusual genetic condition<br />

with two main phenotypes – brachydactyly,<br />

or shortened fingers and toes and<br />

overall stature, and severe hypertension. If the<br />

hypertension went untreated, the affected<br />

family members all died of strokes before they<br />

reached the age of 50. What particularly inter-


The region of Turkey where Nihat Bilganturan found an extended family with<br />

brachydachtyly and hypertension.<br />

ested us was the fact that their hypertension<br />

wasn’t particularly dependent on salt. This<br />

might mean that the mechanisms that<br />

caused it were closer to essential hypertension<br />

than other known genetic conditions. So<br />

we thought it would be worthwhile to identify<br />

the gene that caused it.”<br />

One of Friedrich’s students, Hakan Toka, had<br />

some background in medical genetics. He also<br />

came from a Turkish family who had settled in<br />

Germany a few decades ago, when the father<br />

had been hired to work in a BMW plant.<br />

Hakan had been born in Germany but spoke<br />

Turkish. He volunteered to call his uncle, a<br />

general in the Turkish army, to see if the<br />

researcher who conducted the original study<br />

was still alive. “Nihat Bilginturan was 68 years<br />

old, and had retired from his position as professor<br />

of pediatrics and endocrinology at the<br />

University of Ankara,” Friedrich says. “He was<br />

still living in Ankara, so several of us flew over<br />

to see him. We met in a café. He was a pleasant,<br />

rather formal fellow. Initially he was hesitant<br />

about participating – afraid the people<br />

would be commercialized or compromised<br />

somehow. But we managed to convince him<br />

that our motives were pure.”<br />

Bilginturan introduced the clinician-scientists<br />

to the families. They were spread among<br />

three clans that didn’t get along with each<br />

other, but the researchers managed to convince<br />

them to undergo some initial tests.<br />

Friedrich and his team stayed for a week conducting<br />

interviews, carrying out over 50 physical<br />

exams, measuring blood pressure, drawing<br />

Interlude: The case of the short-fingered musketeer<br />

blood to obtain DNA samples, and photographing<br />

fingers. By the end of the stay the<br />

family had agreed to participate in a yearlong<br />

clinical study of anti-hypertensive drugs. There<br />

were five German participants in the first<br />

Turkish adventure. Hakan Toka and Okan Toka,<br />

brothers and medical students who were fluent<br />

in Turkish, Thomas Wienker, human geneticist<br />

from the <strong>MDC</strong>, Herbert Schuster, internist<br />

and medical geneticist from the Franz Volhard<br />

Clinic and an associate of Friedrich’s. Not to<br />

forget Friedrich himself, who was responsible<br />

for anything that went wrong.<br />

“We had no funding for any of this – so far<br />

everything was being funded by my VISA<br />

card,” Friedrich says. “Then I heard about a<br />

NATO grant program from a captain I knew in<br />

the air force – I had also been in the U.S. military.<br />

They gave us 20,000 dollars, which was<br />

used to pay the VISA card debt, and that’s how<br />

the project started.”<br />

��<br />

Back in Germany, the group faced two<br />

challenges. The first was to conduct additional<br />

phenotyping, namely to get at the clinical<br />

mechanism underlying the patients’<br />

hypertension. The second was to find out<br />

where the gene might be located. So a few<br />

months after the visit, several members of the<br />

family with the disease – and a few without –<br />

were flown to Germany. They were checked<br />

into the Franz Volhard Clinic, where they<br />

underwent a week of extensive clinical testing.<br />

At the time the process was unusual. “The<br />

222<br />

Turkish patients were obviously not insured<br />

and could lay no claim to any health care in<br />

Germany,” Friedrich says. “So who paid for<br />

these studies and the patients’ time in the<br />

hospital? That issue, which was addressed in<br />

the U.S. more than 50 years ago, is one that<br />

has occupied quite a bit of my time over the<br />

past 15 years.”<br />

Numerous tests were necessary to find out<br />

whether the family’s problems resembled<br />

essential hypertension. “What you do is measure<br />

how the body responds to certain things<br />

that are normally signals for an elevation or<br />

drop in blood pressure,” Friedrich says. “If the<br />

volume of the blood increases – if the system<br />

is carrying more liquid – then pressure needs<br />

to drop, and this is managed by changes in<br />

the levels of certain hormones. If there’s more<br />

salt in the diet, it tells the system to absorb<br />

more water. We also checked other typical<br />

kinds of problems that accompany hypertension<br />

– we checked arteries in the kidneys and<br />

eyes to make sure they were okay, and looked<br />

at levels of things like calcium and lipids in<br />

the diet. What we found was pretty normal,<br />

which meant that this condition had a lot in<br />

common with essential hypertension.”<br />

The major challenge was to find the culprit<br />

gene. The task fell mainly to Herbert Schuster<br />

and Silvia Bähring. Herbert was a clinician-scientist<br />

that Friedrich had recruited to assist in<br />

genetic research; Silvia had a PhD in molecular<br />

biology and substantial technical experience.<br />

The two began a linkage analysis, a<br />

method to identify the region of DNA that


carried a particular gene responsible for a disease.<br />

Linkage is the road to answering the<br />

question “Where is it?” – namely, what part of<br />

the genome is responsible for a trait that scientists<br />

want to study. Thomas Wienker was in<br />

charge of the data analysis, making mathematical<br />

models of the data to prove the location<br />

of the responsible gene.<br />

The strategy goes back to the early days of the<br />

20th century, when the great geneticist<br />

Thomas Hunt Morgan launched a new science<br />

now known as “classical genetics.”<br />

Morgan traced the locations of hundreds of<br />

genes in the four chromosomes of the fruit fly<br />

Drosophila melanogaster – in a day when no<br />

one knew what genes were and there were no<br />

molecular techniques to investigate them. His<br />

students spread throughout the world, starting<br />

new laboratories and leaving a trail of<br />

Nobel prizes in their wakes. One member of<br />

his laboratory, Herman Muller, spent some<br />

time in Berlin-Buch. Muller, a Nobel prize winner<br />

from Indiana University in Bloomington,<br />

was a mentor to Nikolai Wladimirowitsch<br />

Timofejew-Ressowski, who in turn worked<br />

with Max Delbrück to put Berlin-Buch on the<br />

scientific map. But that is another story.<br />

Friedrich is undoubtedly one of the most<br />

widely read people on campus, and it shows<br />

when I mention Morgan’s name. “His grandfather<br />

was a famous horse thief and confederate<br />

general named John Morgan from<br />

Kentucky. If you’re interested in the history of<br />

the U.S. Civil War, he shows up as a bandit – in<br />

the South he was called a general.” This<br />

Bilganturan and<br />

Thomas Wenker<br />

discussing the<br />

Turkish pedigree<br />

Thomas Morgan<br />

vignette isn’t the only colorful aspect of the<br />

lineage of Morgan, who was also a direct<br />

descendant of governors, diplomats, and<br />

Francis Scott Key, the composer of the U.S.<br />

national anthem.<br />

Morgan started his career at a pivotal<br />

moment in the history of science. Three<br />

researchers in different European countries,<br />

working independently, were about to rediscover<br />

the principles of heredity that had been<br />

found by the monk Johann Gregor Mendel<br />

and then lost for 40 years. One of the rediscoverers<br />

was Hugo de Vries, who coined the<br />

term “gene”. In 1900 Thomas Morgan dropped<br />

in on de Vries in Amsterdam and wrote, “No<br />

one can see his experimental garden, as I have<br />

had the opportunity of doing, without being<br />

greatly impressed.”<br />

The visit was the beginning of a life-long<br />

friendship between the two men and for<br />

Morgan, an introduction to genetics. His initial<br />

interest was the connection between<br />

heredity (concerned with how existing genes<br />

were shuffled about and combined to give<br />

organisms their physical features) and evolution<br />

(which focused on the origins of new<br />

223 Interlude: The case of the short-fingered musketeer<br />

species). For the first decades of the 20th century,<br />

a fierce debate raged about whether or<br />

not the two ways of thinking were at all compatible.<br />

Genetics could explain two of the<br />

basic concepts of evolution – why individuals<br />

within a species were different from each<br />

other, and how they passed down characteristics<br />

to their offspring. But evolution obviously<br />

involved more than simply shuffling around<br />

existing genes; new traits had to come from<br />

somewhere.<br />

De Vries came up with a proposal: organisms’<br />

genes might undergo changes in a process<br />

called mutation. He hoped that his studies of<br />

plants might catch evolution in the act – if he<br />

watched enough pea plants give rise to seeds<br />

that were round or wrinkled, green or yellow,<br />

a puffy purple one might suddenly appear.<br />

Back in the U.S., Morgan decided to try the<br />

same thing with animals. Finding mutations<br />

would require a lot of animals and a lot of<br />

generations, so he needed an organism that<br />

reproduced quickly, had a lot of offspring, and<br />

was easy to care for (his students would need<br />

to learn to handle them). A colleague gave<br />

him some fruit flies, which could be kept alive<br />

on a diet of mashed bananas, and the stage


was set for over three decades of pioneering<br />

research. Within two years Morgan began<br />

finding mutations in flies. Over the course of<br />

his career he found hundreds. He didn’t learn<br />

much about evolution – his flies didn’t suddenly<br />

develop into a new species – but he discovered<br />

a great deal about the nature of<br />

genes.<br />

Mendel’s laws stated that genes were inherited<br />

independently of each other, but Morgan<br />

and his colleagues quickly began to find<br />

exceptions. Genes on different chromosomes<br />

behaved independently, but two genes on the<br />

same chromosome were usually inherited<br />

together. Even that principle had its exceptions,<br />

which were revealed by statistical studies<br />

of tens of thousands of flies over generations.<br />

Within the same chromosome, genes<br />

that were near each other were inherited<br />

together more often than those farther apart.<br />

The reason is that pieces of chromosomes<br />

occasionally break as they are copied. The<br />

pieces can be glued back into the chromosome,<br />

but sometimes they are reinserted in<br />

the wrong one. This separates genes that<br />

were formerly on the same chromosome.<br />

Genes that are close together are more likely<br />

to be in the same fragment when it moves.<br />

This concept, called linkage, allowed Calvin<br />

Sturtevant and other members of Morgan’s<br />

lab to use simple math and statistics to map<br />

the locations of genes on chromosomes.<br />

Today the same general strategy allows scientists<br />

to discover genes responsible for human<br />

diseases. They no longer do linkage studies<br />

Interlude: The case of the short-fingered musketeer<br />

quite the way Morgan did – which would<br />

require huge, long-term studies with an enormous<br />

number of participants. In the 1980s a<br />

British researcher named Alec Jeffries discovered<br />

a shortcut. He found small regions of<br />

DNA called microsatellites, identical units<br />

spread throughout the genome of every<br />

human. Satellite DNA consists of “repeat”<br />

sequences which – like at least 97 percent of<br />

the genome – do not encode proteins. The<br />

function of most of this non-coding DNA is<br />

unknown, leading some to call it “junk” (a designation<br />

which is “dangerous and premature,”<br />

Friedrich says).<br />

Microsatellites are very small pieces of satellite<br />

DNA. They do not encode genes but are<br />

inherited in much the same way that genes<br />

are. Their locations is monotonously similar<br />

from person and family to family. Within<br />

months of Jeffries’ discovery, a huge number<br />

of uses had been discovered for microsatellites,<br />

including “DNA fingerprinting,” which<br />

can match DNA from a crime scene to a suspect,<br />

and paternity testing.<br />

��<br />

Everyone on Earth (as well as anyone currently<br />

on the space station, Friedrich<br />

quips) carries the same microsatellites at the<br />

same locations in their genomes. They are<br />

made of long strings of repeated sequences,<br />

spelled from the four letters of the genetic<br />

code (C, A, T, and G). A common microsatellite<br />

repeat is CACACACACACA... (it goes on like<br />

that for a while). What makes microsatellites<br />

224<br />

useful, Friedrich says, is that they come in different<br />

lengths. Over time, some microsatellites<br />

collect extra repeats through random<br />

errors that happen as DNA is copied. They are<br />

then passed along from parent to child, and<br />

can be followed like footprints down a family<br />

tree, over time.<br />

Each person inherits a double copy of the<br />

“DNA library” – 23 paired chromosomes, one<br />

from each parent – and thus two copies of<br />

each microsatellite. So there is one version of<br />

a particular microsatellite on the “short arm”<br />

of chromosome 12, and another version of the<br />

same microsatellite (inherited from the other<br />

parent) at the same location on the second<br />

copy of chromosome 12.<br />

“The idea is to identify specific versions of<br />

microsatellites that are always inherited with<br />

a particular disease, and never in non-affected<br />

family members,” Friedrich says. “When we<br />

find such sequences and know precisely<br />

where they are located, we can conclude that<br />

the disease gene must be very close to the<br />

one or more microsatellites that have turned<br />

up.”<br />

Herbert and Silvia were going to use<br />

microsatellites to home in the location of the<br />

disease that caused the Turkish family’s<br />

hypertension and brachydactyly. “We were<br />

beginners at this,” Friedrich says. “And it was<br />

expensive – at the time you needed radioactively<br />

labeled microsatellite markers. Who<br />

was going to pay for a bunch of beginners to<br />

do a study like this? Our bosses were pes-


simistic and suggested we should turn over<br />

the project to “experts”. I didn’t see things<br />

that way – we were good enough to do the<br />

dirty work, gathering the samples and working<br />

with the patients, and now we were going<br />

to let professionals get the glory? We said, ‘To<br />

hell with them; we are doing this ourselves.’<br />

Smart money was bet against us but the jury<br />

was still out.”<br />

To the rescue came Applied Biosystems Inc.<br />

(ABI), a biotech company based in California.<br />

“ABI had just developed non-radioactive fluorescent<br />

microsatellite markers that covered<br />

the entire genome,” Friedrich says. “‘We’ll give<br />

you the markers and show you how to use<br />

them if you acknowledge that effort in your<br />

paper,’ they said, and our response was, ‘Done<br />

deal.’”<br />

Herbert traveled to ABI headquarters in<br />

California and began working up through the<br />

chromosomes, starting with chromosome 1.<br />

Silvia stayed in Berlin and started with chromosome<br />

22, working her way backwards. “It<br />

would have been nice if the gene locus had<br />

been at one of the ends; as things turned out,<br />

it lay on chromosome 12, almost exactly in the<br />

middle,” Friedrich says. “So there were no<br />

shortcuts. Thomas Wienker shouted “Eureka”<br />

somewhere on the short arm of chromosome<br />

12. Even so, the whole process was accomplished<br />

in six months – not bad for a bunch of<br />

beginners.”<br />

��<br />

The original team that flew to Turkey.<br />

Ithink of a trip my family took to the great<br />

San Andreas Fault in California, source of<br />

the terrible earthquakes that regularly rock<br />

the coast, not far from ABI headquarters. The<br />

fault passes through a park with a hiking trail.<br />

You don’t see a great crack in the earth, but at<br />

one point there is a picket fence that is no<br />

longer straight – the earthquake of 1906 shifted<br />

half of it five or six meters. If you were to<br />

study the damage done to fences, houses, and<br />

other landmarks, you could trace the line of<br />

the fault on the map.<br />

The Turkish family has undergone a sort of<br />

genomic earthquake; Silvia and Herbert’s<br />

study pinpointed its site of origin: a specific<br />

region on the “short arm” of the twelfth<br />

human chromosome. The next step was to<br />

study the area and search for a potential gene<br />

which had undergone mutations. That was<br />

much easier said than done. No genes were<br />

known to exist there. And the Human<br />

Genome Project was not yet completed,<br />

which meant that Silvia had to try to find a<br />

gene the old-fashioned way, through a laborious<br />

process called positional cloning. This<br />

required selecting pieces of human DNA<br />

packed in yeast or bacteria and sorting them<br />

along the linkage interval. It allowed her to<br />

identify a number of genes within the region<br />

and sequenced them, but none had mutations.<br />

The genes were identical in family<br />

members affected by the disease and those<br />

without it.<br />

This outcome isn’t so unusual; a large number<br />

of human proteins and the genes that encode<br />

225 Interlude: The case of the short-fingered musketeer<br />

them remain to be found. “It wasn’t necessarily<br />

a dead end,” Friedrich says. “Now that the<br />

genome has been completed, you can scan<br />

regions like this by computer and look for bits<br />

of code that are hallmarks of genes. Well, we<br />

found sequences that had some of these<br />

characteristics, but other features were missing.<br />

So we were stuck.”<br />

Members of the lab were keeping their eyes<br />

open for other reports in the medical literature<br />

that might shed light on the case. Silvia<br />

found a report on a Japanese child with<br />

brachydactyly whose condition had also been<br />

pinned down to a region in the twelfth chromosome.<br />

The Japanese child had a deletion<br />

syndrome – namely it was missing an entire<br />

portion of the chromosome 12 short arm.<br />

Silvia asked the lead researcher for samples of<br />

the child’s DNA and an X-ray picture of the<br />

child’s hands.<br />

There are many types of short fingers, probably<br />

caused by defects in different parts of the<br />

genome. The Turkish family had a particular<br />

form termed “Type E brachydactyly.” Infor -<br />

mation from the Japanese child would only be<br />

helpful if it had the same type.<br />

The scientists were ecstatic to find that the<br />

fingers matched. Now Silvia used the same<br />

microsatellites to study the child’s DNA. She<br />

discovered that the abnormalities weren’t identical<br />

to those of the Turkish family, but the portion<br />

of the child’s genome that had been lost<br />

partially overlapped with the linkage region<br />

that had been identified from the family.


“It seemed very unlikely to us that there<br />

would be two separate genes so close to each<br />

other that caused brachydactyly,” Friedrich<br />

says. “That assumption allowed us to narrow<br />

down the region where the responsible gene<br />

must lie by half.” It was a big step, but it didn’t<br />

solve the major problem – there was still no<br />

gene.<br />

��<br />

The next stage of the study would require<br />

someone to go to Turkey for an extended<br />

period, and the team found a candidate –<br />

Hakan Toka’s younger brother Okan, a medical<br />

student at the University of Munich, one of<br />

the members of the team during the first visit<br />

to Turkey. Okan needed a thesis topic and was<br />

excited at the prospect of doing a clinical<br />

study on hypertension.<br />

“When you do this type of thing, you also have<br />

a responsibility to take care of the patients,”<br />

Friedrich says. “Okan was going to have a big<br />

job on his hands – alongside taking regular<br />

blood pressure measurements of 50 people<br />

scattered across a large region, prescribing<br />

medications, and managing the rest of the<br />

clinical part of the study, he was responsible<br />

for their basic health care and played social<br />

worker. We got him an apartment but he<br />

Interlude: The case of the short-fingered musketeer<br />

needed a car. You can’t write a grant for a car,<br />

but we received a little money from a drug<br />

company to help us do the study. We bought<br />

him a brand new car for 2,000 dollars – at the<br />

end of the study we gave it to the family. We<br />

talked to some local physicians who promised<br />

to keep an eye on Okan, people he could call if<br />

he needed any help. And then we left him<br />

there. We talked regularly by cell phone, but I<br />

don’t recall any panic calls from him. He managed<br />

to solve his problems by himself.<br />

For Okan, the story began in 1994. “I was a<br />

young third-year medical student,” he says. “I<br />

was looking for a challenging doctoral thesis<br />

when Professor Luft told me he was looking<br />

for someone to spend about a year in the<br />

mountain villages of the Black Sea Cost in<br />

Turkey to perform a quite sophisticated crossover<br />

drug study with first-line-antihypertensive<br />

drugs. I would get all the equipment and<br />

resources needed. I remember he stressed<br />

that the job would require a lot of dedication<br />

because it wouldn’t be ‘routine research.’ I had<br />

to do the clinical research but would also have<br />

to serve as the family’s primary care physician<br />

and social worker, too.”<br />

The villages of the eastern coast of the Black<br />

Sea are perched on the slopes of lush, forested<br />

mountains that begin at the water’s edge.<br />

226<br />

Okan says that most have a poor infrastructure.<br />

“There is no steady supply of warm<br />

water, heat or power. People there are usually<br />

farmers and live from plantations where they<br />

grow black tea and hazelnuts. I had to visit<br />

each of my patients every two weeks to measure<br />

blood pressure over 24 hours, to take<br />

blood and urine samples, and to try to detect<br />

the effects – possibly adverse effects – of the<br />

new medications that were being tried in the<br />

study. Most of the patients lived in this coastal<br />

region. However, others were scattered all<br />

over Turkey, in Istanbul, Bursa, and Ankara. It<br />

took me about a week to see all the patients<br />

in the mountain villages, driving off-road from<br />

one to the next in my vehicle. Afterwards I<br />

usually flew to Istanbul, then on to Bursa and<br />

ended my rounds in Ankara, before flying back<br />

to the coast. One round took me about 12<br />

days. Then I had two days to do all the analyses<br />

before starting the next round. I flew<br />

about 60 times and drove 30,000 kilometers<br />

– mostly off-road – during that year.”<br />

Sometimes it was a rough trip. One visit to a<br />

patient required a drive into the mountains,<br />

taking Okan to an altitude of about 1,500<br />

meters. He finished late and the family told<br />

him to stay overnight – descent in the dark<br />

would be dangerous due to snow-covered<br />

roads and wild boars. But he decided to leave<br />

anyway. Halfway down, after midnight, in the<br />

middle of nowhere, his car had a flat tire.<br />

“I had to change the tire while one of the family<br />

members who had escorted me held a flashlight<br />

in one hand and a gun in the other. Yes, we<br />

got some visitors – a pack of wild pigs. My escort<br />

was afraid of an attack. I was mostly concerned<br />

about the precious blood sample that I had just<br />

obtained. He fired a shot and scared the pigs<br />

away, and I finally got the tire changed and we<br />

could continue on down the mountain.”<br />

Okan’s duties often extended beyond medical<br />

care. All hands were needed – including some<br />

of his patients – when harvest time came<br />

around. “I didn’t think it was a good idea to<br />

send somebody out to do hard field work


while I was recording 24 hours of blood pressure<br />

values to evaluate the benefits of a new<br />

antihypertensive drug,” he says. What was a<br />

doctor to do? He spent most of the month of<br />

August harvesting hazelnuts, taking his<br />

patients’ shifts in the field.<br />

��<br />

The second gunshot of the story had more<br />

serious consequences.<br />

“A lot of funny but also awkward things happened<br />

during this time. I remember one particular<br />

incident where one older member of<br />

the family was put in jail. He was the head of<br />

the clan we were investigating. He is a very<br />

friendly, sweet old man; he is about 5 feet tall<br />

and he wears huge glasses. The problem<br />

started when I got an excited call from his<br />

son. The father had gotten into some trouble<br />

with his neighbors and when things got<br />

threatening, he fired his gun in the air to scare<br />

them away. They ran off, all right, but then<br />

they called the police. That wouldn’t have<br />

been a problem in itself. He wasn’t afraid of<br />

the police because he had the right to defend<br />

his land and scare them away. The problem<br />

was the gun. He had fired an old musket he<br />

had inherited from his grandfather and he<br />

didn’t have a license for it. When he saw the<br />

police coming he looked for a place to hide it<br />

and in the heat of the moment, he simply<br />

threw it out the window. It didn’t take them<br />

long to find it, and when they did, they arrested<br />

him.<br />

“The son wanted me, as the family doctor, to<br />

help his father out of prison – and that’s what<br />

I tried. I visited the scared old man. He told me<br />

the whole story all over again. I asked the<br />

police what the accusations were and what<br />

sort of punishment he should expect. The officer<br />

told me that things looked bad – in the<br />

Black Sea Coast area, using an unlicensed gun<br />

was a serious offence. Judges usually imposed<br />

harsh penalties. In any case, the old man<br />

would probably be in jail for two or three<br />

months until the trial. The only thing I could<br />

do, the officer suggested, was talk to the dis-<br />

Hakan Toka, Okan Toka<br />

trict attorney. I might be able to obtain his<br />

release until the date of the trial.<br />

“So I went to the district attorney. I told him<br />

that the old man was an honest person and<br />

promised that he wouldn’t run away and that<br />

he would appear for the trial. More importantly,<br />

I said, he had a congenital disease with<br />

severe high blood pressure. He had already<br />

had one heart attack and from a medical<br />

point of view, the stress of being in jail could<br />

kill him. The attorney said he couldn’t be of<br />

any help, but the judge might be able to do<br />

something.<br />

“I went to the judge and told him the whole<br />

story again. At first he didn’t show any inclination<br />

to be helpful. I was getting desperate so I<br />

promised I would vouch for the old man, and I<br />

would also write a medical report to his superiors<br />

in Ankara, testifying that the old man<br />

was too ill to be in prison. I dropped a lot of<br />

names, trying to impress him. The study, I said,<br />

was a joint collaboration between the famous<br />

Hacettepe University Hospital in Ankara and<br />

the famous Max-Delbrück Center for<br />

Molecular Medicine in Berlin and was receiving<br />

support from the Department of Health in<br />

227 Interlude: The case of the short-fingered musketeer<br />

Ankara. I told him that the old man was a key<br />

subject in our study, and that we hoped we<br />

would learn about the genetic mechanisms<br />

behind primary arterial hypertension, a common<br />

disease worldwide.<br />

“The name-dropping didn’t seem to impress<br />

him very much. What did get his attention<br />

was therapy for hypertension – it had been<br />

the cause of a heart attack that his mother-inlaw<br />

had suffered. He asked, ‘What is all of this<br />

about high blood pressure? Don’t you have a<br />

cure for it?’ I started to tell him about primary<br />

and secondary hypertension and about<br />

‘exceptional families’ who can help to evaluate<br />

the genetic causes of primary hypertension.<br />

He ordered some black tea from his secretary.<br />

We sat and talked at least for an hour.<br />

At the end, he ensured me that I would have<br />

his full support for the study and entrusted<br />

the old man in my care.<br />

“The old man was excited and very grateful to<br />

leave the prison. He told everybody in the clan<br />

that ‘the German Doctor’ had secured his<br />

release from jail. After that I wasn’t the foreign<br />

investigator anymore – I was adopted, and<br />

regarded as a respected member of the clan.”


The year was full of stories, Okan says. “But<br />

the main thing is that this year was an exceptional<br />

period of my life. I learned how truly different<br />

scientific fieldwork is from lab experiments.<br />

It requires a great deal more spontaneity<br />

and improvisation and certainly lots of<br />

dedication to overcome difficulties in the ‘protocol.’<br />

And I’m very grateful to Professor Luft,<br />

who is an extraordinary personality as well as<br />

a brilliant scientist and physician, for giving<br />

me this opportunity.”<br />

One important aspect of the year for Okan<br />

was to establish contact with his “roots”. It is<br />

one thing for a Turkish-Bavarian to have contact<br />

with his heritage through his parental<br />

home. It is quite another thing to experience<br />

that heritage directly, alone, as a primary care<br />

physician with immense responsibilities for<br />

strangers, in a rural setting.<br />

Okan obviously wasn’t the only one to have<br />

profited. His work ensured that the family<br />

received dedicated medical care, and one<br />

result of the year was the discovery that their<br />

most serious medical problems could be kept<br />

under control with anti-hypertensive drugs. It<br />

means a temporary reprieve for people who<br />

previously had a life expectancy of about 50<br />

years. Since the study began, Friedrich says,<br />

none of the family members have died.<br />

��<br />

Finding a mutation in a gene would be a<br />

start – but probably only a start – in<br />

understanding why people develop brachydactyly<br />

and hypertension. The symptoms of<br />

Interlude: The case of the short-fingered musketeer<br />

the Japanese and Turkish patients show that<br />

wherever the genetic defect lies, it alters the<br />

way the body develops. One effect is a change<br />

in the growth of fingers. Hypertension might<br />

be the result of another anatomical change,<br />

somewhere else.<br />

Some cases of high blood pressure could be<br />

linked to the structure of arteries in the brain.<br />

Ramin Naraghi, a neurosurgeon at the<br />

University of Erlangen-Nürnberg, had used<br />

magnetic resonance tomography (MRT) to<br />

chart the course of blood vessels in patients<br />

with essential hypertension. In 20 of 24<br />

patients with essential hypertension – and only<br />

three of his control patients – Ramin found that<br />

blood vessels were squeezing the left side of a<br />

small structure in the brain called the rostral<br />

ventrolateral medulla. This tissue helps regulate<br />

autonomous body functions like the rate of the<br />

heartbeat and blood pressure.<br />

“What happens is a bit like putting a kink in a<br />

garden hose,” Friedrich says. “The twist in the<br />

artery causes a build-up of pressure, and that<br />

puts stress on the medulla. The compression<br />

might make the tissue send more signals out<br />

through nerves, which could explain high<br />

blood pressure. Ramin discovered this right<br />

about the time we had started our study, and<br />

we hadn’t done imaging in the brain. So we<br />

thought we’d better have a look at our<br />

patients.”<br />

The facilities were available in Turkey, so<br />

Ramin, the Toka brothers, and Silvia – who<br />

needed a break from all that lab work – took<br />

228<br />

off for Trabson. Trabson is a bustling city on<br />

the eastern Black Sea coast with a university<br />

hospital and a magnetic resonance facility. In<br />

all of the 15 family members with hypertension<br />

that were examined – and none of those<br />

without it – Ramin discovered a loop in the<br />

brain’s blood vessels which would put pressure<br />

on the medulla.<br />

As a physiological explanation for the patients’<br />

high blood pressure, it made sense. Mutations<br />

often lead to changes in the body – in this<br />

case, shortened fingers and an abnormal routing<br />

of a brain artery. It’s the kind of thing that<br />

is seen all the time in animal models in the lab:<br />

remove a gene and change the body.<br />

All that would have been fine – even the<br />

geneticists would have been happy – if the lab<br />

could have pointed the finger at a mutated<br />

gene. It might have permitted them to reproduce<br />

the condition in mice and then begin<br />

looking for possible treatments.<br />

But so far, they still hadn’t found one. And<br />

that was likely to be crucial to finding a treatment<br />

for the condition.<br />

“With serious developmental defects, even if<br />

you’ve identified a problem such as the route<br />

of a blood vessel, the solution for the patient<br />

isn’t usually going to be to go in with surgery<br />

and try to rebuild the body,” Friedrich says.<br />

“Instead we need to pin down the mechanisms.<br />

But without a gene, we were<br />

stumped.”<br />

Friedrich says that the hypothesis regarding<br />

the kinked artery is technically termed a “neu-


ovascular compression syndrome.” This<br />

mechanism is known to be involved in some<br />

types of brain-stem diseases, and it can be<br />

partially dealt with through surgery.<br />

“However, operating on people’s heads to<br />

lower their blood pressure is not a generally<br />

accepted strategy to cope with hypertension<br />

– although several neurosurgical groups volunteered<br />

to do so. It wouldn’t occur to me to<br />

operate on the family’s fingers to lower their<br />

blood pressure, so why would I want to operate<br />

on their heads?” asked Friedrich. “You certainly<br />

wouldn’t go that way without a lot<br />

more clinical evidence.”<br />

So the family members were once again invited<br />

to Berlin. By this time, a more sophisticated<br />

Clinical Research Center (CRC) had been<br />

established at the Franz Volhard Clinic. A CRC<br />

is a unit dedicated to research in human subjects<br />

and its mission is different from a hospital,<br />

whose efforts are focused on treatments<br />

rather than research.<br />

The CRC is headed by Jens Jordan, who was<br />

trained in medicine and clinical pharmacology<br />

in Berlin-Buch, and cardiovascular regulatory<br />

physiology at Vanderbilt University in<br />

Nashville, Tennessee. Jens reasoned that the<br />

neurovascular compression might increase<br />

the activity of the sympathetic nervous system<br />

in the hypertensive family members. To<br />

test that notion, he needed a “readout” from<br />

the sympathetic nervous system. This<br />

involved implanting a tiny tungsten needle in<br />

the peroneal nerve, which serves the leg<br />

below the knee joint.<br />

“The needle was inserted in sympathetic<br />

nerve fibers that can then be monitored with<br />

a recorder,” Friedrich says. “Now we had to<br />

repeat our tests of the cardiovascular system,<br />

challenging it with drugs. As we artificially<br />

induced changes in blood pressure, we were<br />

now watching how the nervous system<br />

responded. One of the final tests involved<br />

temporarily paralyzing the entire autonomic<br />

nervous system with a drug. This procedure<br />

was a major intervention for the patients and<br />

the university ethics committee reviewed the<br />

study protocol with immense care.”<br />

When the procedure was approved, the team<br />

found that family members did not develop<br />

increased activity of the sympathetic nervous<br />

system. Instead, the team learned that family<br />

members with the condition had almost no<br />

ability to cope with events that changed<br />

blood pressure. In other words, their bodies<br />

weren’t adapting and responding to changes.<br />

Mechanisms that step in to lower blood pressure<br />

in healthy people weren’t working.<br />

The finding is an important clue, and it shows<br />

that pressure on the medulla is connected to<br />

the problem of hypertension. “But we won’t<br />

offer any operations until we understand the<br />

molecular mechanisms that underlie the<br />

problem,” Friedrich says. “There may be more<br />

going on, and until we have pinpointed the<br />

defect in the genome, we can’t use animal<br />

models such as mice to analyze and study the<br />

disease.”<br />

��<br />

229<br />

The study was only one of many other<br />

things going on in the doctors’ lives.<br />

Hakan Toka, Okan’s brother, went off to<br />

Harvard for a clinical rotation in pediatric<br />

nephrology. The rotation was arranged with<br />

Friedrich’s help: Harvard students pay huge<br />

tuition fees and to arrange a senior student<br />

rotation there tuition-free is not a part of the<br />

standard German medical education. While<br />

there, Hakan was asked to give a talk on his<br />

research, and he chose this hypertension<br />

study as his topic.<br />

Interlude: The case of the short-fingered musketeer<br />

When he finished a man in the back raised his<br />

hand. It was James Melby, an endocrinologist<br />

from Boston, who has known Friedrich for<br />

over 30 years. “I’m treating a person with<br />

brachydactyly and hypertension, in my outpatient<br />

clinic,” he said. “Maybe we should talk.”<br />

The family of the patient that Melby was<br />

treating fit the pattern perfectly. Four members<br />

had the symptoms, four had died of<br />

strokes, and three had normal stature and<br />

blood pressure values. The same cause had to<br />

underlie both the Turkish and American cases,<br />

Hakan reasoned; DNA samples from Melby’s


Group picture in the hospital in Trabzon. The “muskateer” of the story is on the left.<br />

patients might help narrow the search for a<br />

defective gene.<br />

Another adventure ensued when Herbert,<br />

Silvia, Hakan, and Friedrich traveled to<br />

Johannesburg, South Africa. They had been<br />

given a lead on yet another group there with<br />

brachydactyly and hypertension. When they<br />

examined the family, they discovered the<br />

same fingers, namely type E brachydactly.<br />

However, not all affected persons were hypertensive.<br />

“Essential hypertension is common,” Friedrich<br />

reminds me. “Short fingers and hypertension<br />

could coincide by chance alone. But when I<br />

was asked to present a talk on the syndrome,<br />

once again a hand went up in the back. The<br />

hand belonged to a staff pediatrician at the<br />

university hospital. She had a patient, a small<br />

child with type E brachydactyly and hypertension.<br />

Silvia went back to the lab and proved<br />

that the child had our syndrome. This patient<br />

is the only isolated case – an individual where<br />

the syndrome doesn’t run in the family – that<br />

we have identified so far. Such cases can happen<br />

through spontaneous mutations in a single<br />

person, if they affect a region of genome<br />

that is known to cause disease.”<br />

Then a third family turned up, in Canada. They<br />

had been discovered by David Chitayat, a<br />

physician and genetic counselor who is now<br />

Interlude: The case of the short-fingered musketeer<br />

head of the Prenatal Diagnosis and Medical<br />

Genetics Program at Mount Sinai Hospital in<br />

Toronto, Canada. At the time he was at the<br />

Hospital for Sick Children in Toronto. Chitayat<br />

got involved when a seven-year-old child was<br />

admitted with persistent headaches that<br />

were followed a few days later by seizures. The<br />

child was short for his age, with the shortened<br />

fingers and toes typical of brachydactyly<br />

– and high blood pressure. A family history<br />

revealed that his brother, father, uncle, cousin,<br />

and grandmother (all on the father’s side)<br />

were also affected. They were all being treated<br />

for hypertension.<br />

Chitayat had read the 1973 study by Nihat<br />

Bilginturan (in his own paper he called the<br />

Canadian family’s condition “Bilginturan<br />

Syndrome”) and had also read the article in<br />

which Friedrich’s group pinned the source of<br />

the problem to a region of chromosome 12. In<br />

1998 Hakan, Chitayat, Melby, and Friedrich’s<br />

lab put data from the three families together.<br />

Microsatellite studies of the Canadian and<br />

American patients showed a genetic change<br />

within the same region of the chromosome.<br />

Because the problems did not completely<br />

overlap with the Turkish data, it allowed the<br />

group to narrow down the search even more.<br />

There were still no clear genes within the target<br />

region, but there was one nearby: a<br />

sequence encoding the molecule L-SOX5. At<br />

230<br />

the time, little was known about the functions<br />

of human L-SOX5. But Veronique Lefebre,<br />

a French scientist working in Cincinatti, Ohio,<br />

was investigating the version of L-SOX5 found<br />

in mice. She sent an email to Silvia telling her<br />

that she had discovered one of the functions<br />

of the molecule. L-SOX5 protein encoded a<br />

gene-activating protein called a transcription<br />

factor. In mice it activated a molecule called<br />

collagen type II. “That was interesting,”<br />

Friedrich says, “because L-SOX5 is produced in<br />

the tips of the toes, the long bones, and vertebra<br />

of developing mice. Collagen type II is a<br />

tough, fibrous protein that is the main component<br />

of bone and cartilage. This made L-<br />

SOX5 a great candidate gene in terms of<br />

explaining the skeletal anomalies in patients.<br />

So we thought we’d better sequence the<br />

entire human gene.”<br />

In the days before the completion of the<br />

human genome, this often required a huge<br />

effort – a bit like searching for a small bit of<br />

text in the Encyclopedia Britannica and then<br />

reconstructing the pages that come before<br />

and after it, letter by letter. The gene was likely<br />

to be spread across a large area of chromosome<br />

12, probably broken into pieces, and<br />

interrupted by bits of garbled text called<br />

introns. Far from being irrelevant “junk”, these<br />

bits of nonsense within a gene can disrupt a<br />

gene’s functions and make the protein that it<br />

encodes useless.<br />

“It took Silvia a year to sequence the entire<br />

human L-SOX5 gene,” Friedrich says. “It turned<br />

out to be a monster, over 500,000 base pairs<br />

long. The cause of our syndrome might be a<br />

change in any one of those letters – we’d only<br />

know if we found the same flaw in every<br />

member of the family with the disease, and in<br />

none of those who didn’t have it. In such a<br />

huge gene you find all kinds of variations. So<br />

every time Silvia found a single letter that had<br />

changed, she checked the pattern of inheritance<br />

in the family. There wasn’t a consistent<br />

pattern. It allowed her to rule out L-SOX5.<br />

There were lots of tears in the beer after this<br />

setback! Our best candidate was gone.”


Silvia Bähring


Friedrich Luft plays guinea pig.<br />

L-SOX5 lay just outside the target region on<br />

chromosome 12. “But mutations weren’t the<br />

only possibility,” Friedrich says. “We knew there<br />

had to be something going on in that region<br />

and started thinking about other kinds of<br />

changes that DNA can undergo. By that time,<br />

Silvia had already sequenced most of the coding<br />

genes in the region. If nothing had become<br />

misspelled, or none of the letters of the code<br />

had been lost, then maybe the order of things<br />

has just been rearranged.” (Some of the ways<br />

that DNA sequences become rearranged are<br />

described in the chapter “Waking a Sleeping<br />

Beauty and other tales of ancient genes.”)<br />

To discover whether this had happened, Silvia<br />

used bits of human DNA grown in bacteria<br />

(called BACs) to make fluorescent probes of<br />

various colors. Silvia attached red, green, and<br />

yellow fluorescent markers to specific locations<br />

in the DNA from the BACs – like planting<br />

flags in the ground to track the course of an<br />

earthquake. If there had been no rearrangements,<br />

all the flags would always appear in<br />

the same order, in every person’s DNA. That<br />

was the case for all the healthy members of<br />

the family. But those that had inherited<br />

brachydactyly had also inherited a jumbled<br />

Interlude: The case of the short-fingered musketeer<br />

region on one copy of chromosome 12 A piece<br />

of DNA had been cut out, flipped around, and<br />

reinserted into a new place. This had changed<br />

the order of the flags.<br />

“The real test was to look at the same<br />

sequences in the Canadian and American<br />

families, as well as the child from South<br />

Africa” Friedrich says. “There, too, we found<br />

rearrangements. Interestingly, each family<br />

and the child had shuffled things around a bit<br />

differently. But in each case, the same region<br />

was disturbed.”<br />

Even though the changes didn’t directly<br />

involve a gene, they still might affect a gene<br />

someplace else. For example, DNA sequences<br />

near a gene often contain instructions telling<br />

the cell when and where in the body it should<br />

be switched on and off. In this case there<br />

might be instructions important to the construction<br />

of fingers, toes, or blood vessels.<br />

Scrambling that information could well lead<br />

to fingers that were too short, or to the<br />

rerouting of an artery.<br />

Such instructions are most likely to affect<br />

nearby genes, and in this case there were four.<br />

“I don’t need to go into detail,” Friedrich says,<br />

232<br />

“but from what we knew, any of them might<br />

be involved in regulating blood pressure. We<br />

had four pretty hot candidate genes We reasoned<br />

that if we had tissue from affected and<br />

nonaffected family members, we could test<br />

these genes. That meant taking a probe of a<br />

likely tissue – a blood vessel.”<br />

Obtaining samples would mean another trip<br />

to Germany for several members of the<br />

Turkish family. A small section of blood vessel<br />

would be removed in a simple procedure.<br />

Once again the team relied on the CRC, run by<br />

Jens Jordan at the Franz Volhard Clinic.<br />

“We could obtain a sample of blood vessel<br />

through a relatively simple procedure to remove<br />

one from the patient’s buttock,” Friedrich says.<br />

“Still, it was surgery – and elective surgery, not<br />

directly aimed at improving the person’s health.<br />

You never want to do that unless the whole procedure<br />

has been perfectly ironed out. We needed<br />

a guinea pig. So don’t ever say I haven’t put<br />

my ass on the line for this study.”<br />

One of the candidate genes included an ion<br />

channel (see the story called “The electrician’s<br />

toolbox”) and a second protein that controlled<br />

its behavior. Ion channels sit in cell membranes<br />

where they open and close, allowing<br />

charged particles to pass into and out of the<br />

cell. This changes the cell’s charge and creates<br />

electrical impulses that travel along nerves –<br />

which could well be involved in the kinds of<br />

nervous system problems that had been<br />

detected in the patients. Maik Gollasch,<br />

internist, nephrologist, and electrophysiologist<br />

– with credentials ranging from Moscow<br />

to Vermont, but nonetheless a clinical investigator<br />

from Friedrich’s department at the<br />

Franz Volhard Clinic – relied on a sophisticated<br />

set of experiments using a technique called<br />

patch clamp, which can detect whether or not<br />

ion channels are opening and closing properly.<br />

(See “The electrician’s toolbox” for a closer<br />

description of the patch clamp technique and<br />

more of its uses.)<br />

Unfortunately the results were negative. The<br />

researchers couldn’t detect any differences in


the way the four genes were being used in the<br />

tissues of patients and their healthy relatives.<br />

The candidate genes were innocent.<br />

More tears in more beers. Back to the drawing<br />

board.<br />

��<br />

At some point or another, Norbert Hübner,<br />

a basic <strong>MDC</strong> scientist whose specialty is<br />

complex genetics, has been involved in several<br />

of the stories in this book. Besides carrying<br />

out several of their own research projects, his<br />

group has been running the <strong>MDC</strong>’s DNA<br />

microarray facility and helping other groups<br />

look for diseases caused by combinations of<br />

genes. That’s a huge task involving vast quantities<br />

of data which can only be grasped using<br />

sophisticated statistics and biocomputing<br />

methods. Friedrich had been talking to<br />

Norbert about the project for a while, and in<br />

2003 he jumped in to make a major contribution<br />

to the project.<br />

“Up until that time there was still the possibility<br />

that we were deluding ourselves about<br />

the connection between this genetic<br />

condition and ‘normal’ essential hypertension,”<br />

Friedrich says. “It was time to try to<br />

bridge the gap, and the way we did it was to<br />

involve yet another continent in the project –<br />

China.”<br />

The idea was to try to repeat the type of study<br />

that had pinpointed chromosome 12 with the<br />

Turkish family, but this time with a “normal”<br />

population suffering from “normal” essential<br />

hypertension. “Such studies have been carried<br />

out before, but generally the results have not<br />

been clear,” Norbert says. “The problem is that<br />

in any typical population you find a huge<br />

range of genetic variation. People migrate and<br />

genes from all over the world get mingled. It’s<br />

hard to sift through and find single things<br />

that contribute to disease. If there are multiple<br />

factors involved, which is the case in<br />

essential hypertension, the job gets much<br />

harder. You’re not looking for a needle in a<br />

haystack, you’re looking at a whole field of<br />

haystacks, each of which has lots of needles in<br />

it. You’re searching for the haystack with two<br />

identical needles.”<br />

The best solution is to find a large group of<br />

people who live in a fairly isolated region,<br />

where family relationships are clear so that<br />

diseases can be tracked accurately through<br />

233<br />

Interlude: The case of the short-fingered musketeer<br />

many generations. Norbert’s lab had extensive<br />

collaborations in China, involving families<br />

that had participated in clinical studies<br />

before. This meant that an enormous amount<br />

of preparatory work had already been done.<br />

The project was carried out under the leadership<br />

of Norbert and Maolian Gong, at the


Sino-German Laboratory of FuWai Hospital in<br />

Peking. The group specializes in molecular<br />

genetics. An active collaboration has developed<br />

between the Fu Wai hospital and the<br />

<strong>MDC</strong>. Maolian is currently “on loan” in Berlin,<br />

Friedrich says. “Her endless enthusiasm and<br />

capacity for hard work has made her a valued<br />

partner for the research teams of Norbert and<br />

myself.”<br />

“We started with a group of 94 people from a<br />

large family group,” Norbert says. “The analysis<br />

showed a strong link between hypertension<br />

and alterations in chromosome 12. Then<br />

we added 32 more families – a total of 174 parents<br />

and their children – from the Shijingshan<br />

district. There was a high rate of hypertension<br />

in this group. When we looked at the distribution<br />

of this larger group, there was an even<br />

stronger trend. There was a nearly perfect<br />

Interlude: The case of the short-fingered musketeer<br />

overlap between the region found in this<br />

study and the one identified by Friedrich in his<br />

Turkish patients. These groups live thousands<br />

of kilometers from each other in different cultures.<br />

This is very strong evidence that we’ve<br />

found a part of the genome that – somehow<br />

– plays an important role in regulating essential<br />

hypertension.”<br />

��<br />

The best way to get an impression of<br />

Friedrich Luft’s sense of humor, and his<br />

finely-tuned ironic world view, is to sit next to<br />

him at a meeting. He doesn’t hesitate to voice<br />

his opinions when he thinks he’s hearing nonsense.<br />

(“Never suffer fools gladly,” he grumbles.)<br />

If you never have that pleasure, you can get a<br />

sense of the man by scanning the titles of the<br />

review articles he writes on a regular basis for<br />

234<br />

the Journal of Molecular Medicine, whose editorial<br />

offices are located on the Berlin-Buch<br />

campus. “A fat attack occurred in fat city,” one<br />

title reads. “Rocking around the clock, while<br />

time is relative.” “Is imprinting in printing or in<br />

press?” “Phosphate’s fate made easier.” “Blue<br />

acid blues,” and “Escalator-driven research.”<br />

And that’s just going back to 2004. (All right, I<br />

have to mention a classic from 2003: “Baa,<br />

baa, black sheep, are your kidneys full?” which<br />

discusses a link between poor diet in mother<br />

sheep, diabetes, cardiac disease and hypertension.)<br />

Alongside writing, caring for patients, and<br />

heading research projects, Friedrich has had<br />

another hobby over the past few years: has<br />

been fighting for the construction of a new<br />

type of institute on the Berlin-Buch campus.<br />

“With ongoing changes in the health care sys-


tem and the way hospitals are administered,<br />

it is becoming more difficult to carry out basic<br />

research projects that involve patients,” he<br />

says. “Today if we needed to check a family<br />

into a hospital for a week, in pursuit of some<br />

important but basic question, there’s no clear<br />

structure to fund that, and the hosptials<br />

aren’t set up for that kind of arrangement.”<br />

And Friedrich knows first-hand how difficult it<br />

is to bridge the gap between clinical research<br />

and the laboratory. New structures are needed<br />

as points of contact for collaborative projects,<br />

to inspire the kind of interdisciplinary<br />

thinking that is needed to carry them out, and<br />

to give young scientists training from both<br />

perspectives.<br />

“There aren’t many good models for doing<br />

this in Europe, but the United States have<br />

been successful at it for many years,” Friedrich<br />

says. “Exactly these issues led them to set up<br />

clinical research centers, or CRCs. The contribution<br />

that the Franz Volhard Clinic CRC has<br />

made to the Turkish project is self-evident.<br />

What we’re now planning in Buch is called the<br />

ECRC, the Experimental and Clinical Research<br />

Center. If things go as we hope, this will<br />

include a significant amount of laboratory<br />

space, support, and infrastructure for common<br />

projects, as well as a more elaborate inpatient<br />

CRC on the Charité side. That will be fully<br />

equipped with examination rooms and diagnostic<br />

tools to carry out a wide range of tests<br />

with people on an outpatient basis. Patients<br />

that need to be admitted can be cared for in<br />

so-called ‘scatter beds’ in the hospital.”<br />

A CRC must be multidisciplinary, Friedrich<br />

says. “When my oncology friends finally devel-<br />

235 Interlude: The case of the short-fingered musketeer<br />

op that gene-therapy or T-lymphocyte cure,<br />

we need to have the beds, laminar flow<br />

rooms, etc. ready for them almost without<br />

notice. The health care system cannot take<br />

the responsibility for research; its not their job<br />

and they cannot do it anyway.”<br />

The new structure will have access to some of<br />

the most sophisticated new instruments in<br />

the world for studying the whole human<br />

body. “Under the ECRC umbrella, with funding<br />

from the BMBF, the <strong>MDC</strong>, and the Charité, we<br />

have just made two major purchases,”<br />

Friedrich says. “One magnetic resonance<br />

imaging scanner that can carry out imaging<br />

of the entire human body, and a second MRI<br />

machine for use with small animals. The<br />

buildings to house them are being constructed<br />

as we speak. These instruments are important<br />

because to extend what we are learning


to patients, we need to extend what we have<br />

learned about cellular processes to the whole<br />

organism. And we have to connect what we<br />

learn from model organisms to humans – our<br />

patients.”<br />

Interlude: The case of the short-fingered musketeer<br />

Accomplishing all of this will require a new<br />

type of scientist, so one of the main activities<br />

of the ECRC will be training clinician scientists.<br />

The investment requires as much as four<br />

years’ bench research that should culminate<br />

236<br />

A comparison between blood vessels from<br />

patients with the hypertension condition<br />

(right) and those without. An expert can<br />

tell the difference!<br />

in a PhD degree for these clinicians, obtained<br />

with the same rigor as the PhD earned by the<br />

basic scientist candidates.<br />

��<br />

Until the late 1990s, the failure to find a<br />

gene in the key region of chromosome<br />

12 would have probably been the end of the<br />

road in the search for the causes of hereditary<br />

brachydactyly and hypertension. Today things<br />

have changed through the discovery that<br />

genomes encode more than just genes. For<br />

example, a great deal of the sequence is used<br />

to make small molecules called microRNAs.<br />

Unlike other RNA molecules, they aren’t used<br />

to make proteins. Instead, they influence<br />

whether other RNAs are allowed to do so. If a<br />

cell produces a microRNA whose chemistry<br />

allows it to bind to another RNA, the two molecules<br />

dock onto each other. This prevents the<br />

other RNA from being used to make proteins.<br />

(The process is described in more detail in the<br />

chapter “A very quiet cure.”)<br />

The discovery of this “code within the code”<br />

presents new mechanisms by which cells regulate<br />

their affairs and diseases might arise.<br />

Silvia, Friedrich and Norbert think the problem<br />

arising from chromosome 12 could well<br />

have to do with a microRNA. Some of these<br />

small molecules have been shown to have<br />

important functions. By preventing the production<br />

of a protein, they can change how a<br />

cell behaves and how it develops. “It’s entirely<br />

reasonable to think that the result could be a<br />

change in the building plan of fingers or<br />

blood vessels in the brain,” Norbert says.<br />

Rearrangements that have been found in the<br />

chromosome may stop the cell from producing<br />

an important microRNA or scramble the<br />

sequence of another molecule that it needs to


ind to. Proving that this is the case will<br />

require discovering the microRNAs encoded in<br />

the region and identifying the target molecules<br />

they bind to. These small bits of code are<br />

hard to find; Norbert and his colleagues are<br />

working on the problem.<br />

“More needles and more haystacks,” he says.<br />

“This time the needles are a lot smaller. But<br />

we have some leads.” Silvia’s work to date is<br />

consistent with a mutated microRNA, a regulator<br />

of regulators as responsible. To bring in<br />

the evidence to convince the scientific community<br />

is a major challenge.<br />

Fortunately, the team has a new strong partner.<br />

Nikolaus Rajwesky recently joined the<br />

<strong>MDC</strong> from the Rockefeller University in New<br />

York. Nikolaus is a world’s authority in<br />

microRNAs. (His work is described in “Playing<br />

the piano of planaria.”) He is also bringing socalled<br />

“deep sequencing” to the <strong>MDC</strong> – a strategy<br />

to discover what the genome is doing<br />

beyond simply turning out proteins based on<br />

its genes. With all the amassed expertise,<br />

Friedrich says, there is real reason to believe<br />

this “rearrangement syndrome” can be solved.<br />

“Had we found the thing quickly, we would<br />

have written one fancy paper and gone on to<br />

other things,” Friedrich muses. “There are so<br />

many blessings from this project that are less<br />

tangible.” In the laboratory, Atakan Aydin<br />

moved from technician to full-fledged scientist<br />

with a “summa cum laude” PhD degree.<br />

Yvette Neuenfeld attended the university on<br />

the side and obtained a Masters degree. In the<br />

clinic, Hakan Toka is a nephrologist currently<br />

working at the Massachusetts General<br />

Hospital at Harvard University. Okan Toka is a<br />

pediatric cardiologist at the University of<br />

Erlangen. Herbert Schuster became professor<br />

of medicine and has moved on to other challenges.<br />

Thomas Wienker is professor of<br />

human genetics at the University of Bonn.<br />

Jens Jordan has just assumed the chairmanship<br />

of the Department of Clinical<br />

Pharmacology at the University of Hannover.<br />

Silvia and Friedrich remain embroiled in the<br />

The campus guest house provided lodgings when the families came to Berlin.<br />

project. “We will never give up,” they say – and<br />

then, why should they?<br />

The culprit has been chased for more than<br />

three decades, across four continents.<br />

Although its identity remains a mystery, the<br />

culprit has been surrounded. Getting this far<br />

has required plodding detective work as well<br />

as the use of the most sophisticated technology<br />

science has to offer.<br />

That’s where the scientific story ends, for the<br />

moment. But stay tuned. A definitive solution<br />

to the mystery may be close at hand. Or it may<br />

require the development of new concepts and<br />

methods. Ten years ago no one would have<br />

guessed that a microRNA might be responsible<br />

for disease. Who knows what the genome<br />

will be whispering to us a decade from now?<br />

237 Interlude: The case of the short-fingered musketeer


Part four: From the protein village<br />

to the cosmopolitan cell


A hub in the city of the cell<br />

By February the city had grown weary of the strikes. For two weeks<br />

there were no S-Bahns, the main link between Buch and central<br />

Berlin. That was an inconvenience, but with creative use of streetcars<br />

and buses and some quick reconstructive surgery on an old bicycle,<br />

most of us without cars could still get to work. When the trains began<br />

rolling again, the bus drivers and U-Bahn staff took their turn. Suddenly<br />

many parts of the city, including the campus, could only be reached by<br />

taxi or on foot or that decrepit bike. Service briefly resumed for Easter,<br />

but the first thing we heard after the holidays was that more strikes<br />

were planned.<br />

At the height of the strike, long lines of disgruntled passengers formed<br />

at the service desk at the Friedrichstrasse, a central hub in Berlin’s transportation<br />

network. They began to share strike stories. Did you hear<br />

about the fistfight at the airport? Four people fighting over a taxi? At<br />

the front of the line a man raised his voice. No reaction from the man<br />

and woman behind the counter – they had heard it all before. The<br />

phone rang and the woman picked it up. “Where are you coming from,<br />

sir?” she asked. “Where do you need to go?” She frowned at her


Claus Scheidereit<br />

computer. Even the most familiar routes had<br />

to be replanned. Finally we got tired of waiting<br />

and went to check a map of the transportation<br />

system – maybe we had overlooked<br />

a connection somewhere that would get us<br />

downtown.<br />

The map reminded me of a frustrating discussion<br />

I’d had that afternoon with Claus<br />

Scheidereit, a scientist at the <strong>MDC</strong>. The frustration<br />

was all mine. Claus’ work concerns<br />

details of another type of network: signaling<br />

pathways in the cell. These are the information<br />

routes by which signals from the environment<br />

are passed to genes, a theme that<br />

appears over and over in this book because of<br />

the crucial role pathways play in most biological<br />

processes.<br />

Sometimes they resemble subway maps – but<br />

they are far more complex. When a biochemist<br />

begins talking about signaling it’s<br />

easy to get lost in the details. To help me keep<br />

track, Claus drew a map – it seemed endless. I<br />

had to remember that nearly every feature on<br />

this chemical map represented years of work<br />

on the part of someone’s laboratory. For an<br />

important purpose: nearly everything that<br />

goes right in a cell – and nearly everything<br />

that goes wrong – can be attributed to how<br />

information travels along the chemical S-<br />

Bahn of signaling pathways.<br />

Although it takes thousands of molecules to<br />

pass the signals that control the cell’s business,<br />

some of the routes are more crucial than<br />

others, and occasionally they cross at<br />

Friedrichstrasse-like hubs. Claus’ laboratory<br />

focuses on one of them, a sort of service desk<br />

that receives information from all over the cell<br />

and routes it to the right genes. Most of his<br />

work concerns a complex of proteins called<br />

NF-κB, spoken “N-F-kappa-B.”<br />

These molecules have dominated his entire<br />

career, partly because of their important functions.<br />

And there’s another reason: if you want<br />

to get an overview of Berlin’s transportation<br />

system, you should go to the Friedrichstrasse.


All the transportation systems meet up here.<br />

People from all corners of the city pass<br />

through on their way to all other parts of the<br />

city. Likewise, if you want to get a deep look at<br />

the basics of how cells manage signals, NF-ΚB<br />

is a good place to start.<br />

��<br />

In 1988, while working as a postdoc at<br />

Rockefeller University in New York, Claus and<br />

his colleagues purified a complex of proteins<br />

containing NF-ΚB from human cancer cells. In<br />

the test tube the factor could activate genes<br />

brought into cells by HIV, the AIDS virus. The<br />

tumor cells originally came from a patient suffering<br />

from Burkitt’s lymphoma (described in<br />

the story “Where slow rivers meet”).<br />

“NF-ΚB had been discovered two years earlier<br />

in the lab of David Baltimore at MIT,” Claus<br />

says. “At the time it was only known to be<br />

Part four: From the protein village to the cosmopolitan city<br />

present in immune system cells. NF-ΚB helps<br />

program the cell to release molecules called<br />

cytokines, which alert the body to the presence<br />

of infectious agents. Another function is<br />

to help switch on the most complex genes in<br />

the body – the sequences used to make antibodies.”<br />

These genes are spread out over a huge<br />

region of the genome. Making them requires<br />

slicing out great chunks of that information,<br />

throwing it away, and pasting the broken<br />

ends of DNA back together. The process<br />

begins when proteins bind to DNA. But until<br />

the late 1980s, very few of the molecules<br />

involved in making antibodies were known.<br />

That changed with the discovery of NF-ΚB.<br />

Baltimore and his colleagues discovered that<br />

it could bind to regions of DNA called Kappa<br />

sequences found in antibody gene sequences<br />

and many other genes. If NF-ΚB didn’t bind<br />

242<br />

there, crucial parts of some antibodies couldn’t<br />

be made. B cells failed to develop properly.<br />

“Once we knew that NF-ΚB bound to Kappa<br />

sequences, we started finding such sites all<br />

over,” Claus says. “They occurred near many<br />

other genes and in the genomes of several<br />

viruses.” HIV had such a site. Another was<br />

found in the monkey virus SV40 (introduced<br />

in the story “A pact with the devil.” ) Both of<br />

these infectious agents attack immune system<br />

cells and insert their genomes into that<br />

of the host. That information has to be turned<br />

into RNA and proteins in order for the viruses<br />

to reproduce, and they get the cell to do it for<br />

them. By including a binding site for NF-ΚB,<br />

they take advantage of a powerful gene-activating<br />

molecule.<br />

What was missing, Claus says, was a better<br />

overview of what genes NF-ΚB might be activating.<br />

And what kept it from switching on all


of those genes all the time? He moved to<br />

Germany and started a group at the Max-<br />

Planck Institute for Molecular Genetics in<br />

Berlin. Six years later he started his lab at the<br />

<strong>MDC</strong>. NF-ΚB has accompanied him on each<br />

move.<br />

After two decades and 70 research articles on<br />

the topic, he’s still at it.<br />

��<br />

Part of the lab’s work has been devoted to<br />

taking the NF-ΚB machine apart to see<br />

how it works. NF-ΚB proteins are made in the<br />

main compartment of the cell, the cytoplasm.<br />

To activate genes they have to move into the<br />

nucleus. This is a critical step, Claus says, and<br />

understanding it has required learning how<br />

NF-ΚB is built and how it interacts with other<br />

molecules.<br />

“The more we have worked on NF-ΚB, the<br />

more functions we’ve found,” Claus says. “I<br />

know you’ve written about DNA chips – the<br />

technology we use to get a look at the entire<br />

gene activity of a cell. A few years ago Daniel<br />

Krappmann, a postdoc, used the method to<br />

compare cells with active NF-ΚB to very similar<br />

cells in which the pathway was quiet. He<br />

found a ‘signature’ of about 70 target genes<br />

that are affected by the signal. Many of those<br />

genes encode proteins that go on to activate<br />

other genes.”<br />

NF-ΚB can set off cascades of events that lead<br />

to the division, self-destruction, or specialization<br />

of cells. Or they cause disease. “So it would<br />

be a dangerous thing to have the signal<br />

switched on for very long,” Claus says. “The status<br />

quo in the cell is for NF-ΚB to be switched<br />

off. It’s kept in check by keeping it away from<br />

genes, keeping it out of the nucleus.”<br />

This is accomplished by a sophisticated protein<br />

“lock”. By finding the molecules that NF-<br />

ΚB binds to – taking apart the machine –<br />

Claus’ lab contributed to identifying the parts<br />

of the lock and determining what kind of key<br />

the cell uses to open it.<br />

NF-κB can activate genes in different ways, depending on how it is activated itself. The pathway should not be<br />

switched on all the time, but that happens in Hodgkin lymphoma disease. Claus hopes to understand why by<br />

untangling the very complex interactions in the pathway.<br />

“The lock is built of proteins called IKBs,” Claus<br />

says. “They fasten themselves directly onto<br />

NF-ΚB and have to be released before it can<br />

activate any genes. Releasing the brake is the<br />

job of a complex of proteins called IKK. So the<br />

main signaling pathway that activates NF-ΚB<br />

does so via IKK, which then unlocks IΚB.”<br />

The lock isn’t just removed, he says; it gets<br />

destroyed. An incoming signal changes the<br />

chemistry of IΚB proteins. This brings it to the<br />

attention of enzymes that come along and<br />

break it down. Once that happens, NF-ΚB is<br />

free to go.<br />

IKK, NF-ΚB and the lock have to work to establish<br />

the right pattern of active and silent<br />

genes in cells. In 2007 PhD student Meike<br />

Broemer and Michael Hinz, a postdoc in Claus’<br />

group, discovered that a protein complex<br />

called Hsp90-Cdc37 helped to give IKK the<br />

shape it needs so that the signal – the key –<br />

fits. Hsp90 and Cdc37 are chaperones, molecules<br />

that help fold proteins into the right<br />

structure. If they don’t do that for IKK, the<br />

complex can no longer release the lock on<br />

NF-ΚB.<br />

Closing the Friedrichstrasse station for a day<br />

would inconvenience a lot of people moving<br />

through Berlin, but wouldn’t be a disaster.<br />

Disrupting a major hub in a signaling pathway,<br />

on the other hand, usually has devastating<br />

consequences for the cell. As the functions<br />

of NF-ΚB became clearer, Claus expected<br />

to find diseases caused by defects in the<br />

mechanisms that keep it under control.<br />

��<br />

Alois Alzheimer’s chance encounter with a<br />

patient led brought him in contact with<br />

the disease that now bears his name. A coincidence<br />

brought Bernd Dörken, a clinical<br />

243 Part four: From the protein village to the cosmopolitan city


Bernd Dörken<br />

researcher at the <strong>MDC</strong> and Charité, in contact<br />

with Hodgkin’s disease research.<br />

Nearly two decades ago, while working in the<br />

clinic at the University of Heidelberg, Bernd<br />

was treating a terminally ill patient. “He had<br />

enormous lymph nodes and a swollen<br />

abdomen where fluid had collected,” Bernd<br />

says. “While we were still trying to find a way<br />

to help, we drained some of that fluid and put<br />

it in cell cultures, planning to save it and look<br />

at it later. The cells began growing wildly.”<br />

A look through the microscope revealed that<br />

the sample contained some large, very odd<br />

cells. They contained two bulging nuclei,<br />

which made them look like owls’ eyes. They<br />

are known as Hodgkin/Reed-Sternberg (HRS)<br />

Part four: From the protein village to the cosmopolitan city<br />

cells and are a characteristic of Hodgkin’s disease,<br />

a cancer of the lymph system.<br />

In 1991 Bernd moved to Berlin and brought<br />

the cells along, hoping they would yield<br />

insights into the causes of the disease. “At the<br />

time labs only had seven lines of cells derived<br />

from Hodgkin’s patients,” he says. “An eighth<br />

would be useful. And it turned out these cells<br />

had some unusual characteristics. They didn’t<br />

produce a protein called CD30, which had<br />

been found in every other Hodgkin’s cell line.”<br />

That fact led to a small scientific skirmish<br />

when the group published their findings in<br />

the Journal of Experimental Medicine in 1993.<br />

A scientist at another institute in Berlin<br />

refused to believe the results. He warned<br />

Bernd to retract the paper – if not, he would<br />

challenge it in print. But further work confirmed<br />

the findings.<br />

“One characteristic of HRS cells is that they<br />

produce huge quantities of cytokines – small<br />

signaling molecules that exert a powerful<br />

attraction on other cells,” Bernd says. “It<br />

explains the swelling of the lymph nodes during<br />

the disease; so many blood cells collect<br />

there. Very few of these cells – often less than<br />

one percent – are actually dangerous tumor<br />

cells. The rest are healthy cells that have been<br />

attracted there. HRS are the only tumor cells<br />

known to release cytokines this way. We<br />

thought that we should look at a transcrip-<br />

244<br />

tion factor that caused the release of the<br />

cytokines, and the logical candidate was NF-<br />

ΚB. That got us talking to Claus Scheidereit<br />

here at the <strong>MDC</strong>.”<br />

It was the beginning of a collaboration that<br />

has now been going on for more than a<br />

decade. One step in finding out what causes<br />

Hodgkin’s disease, Bernd says, means figuring<br />

out what makes blood cells become so curious<br />

that they don’t fit in any normal blood lineage.<br />

That was difficult because the HRS cells<br />

had characteristics of both T and B cells (introduced<br />

in part two, “Identity crisis”).<br />

In the late 1990s Klaus Rajewsky’s lab in<br />

Cologne managed to isolate single cancer<br />

cells from the tumors. When Klaus looked at<br />

the cells’ antibody genes, part of the story<br />

suddenly became clear. He discovered that the<br />

genes had undergone two types of changes<br />

that only happen in B cells. First, the DNA had<br />

been rearranged in a complex cut-and-paste<br />

operation that creates unique sequences. As<br />

new cells are created, each inherits one of the<br />

types.<br />

Bernd compares them to the many types of<br />

key forms found in a locksmith’s shop. “This<br />

already makes a huge number of types of<br />

cells, but things aren’t finished,” he says.<br />

“There’s still fine cutting to do, to make random<br />

changes here and there that really give<br />

each cell its own identity – like taking one of


the blanks and cutting it to fit a specific door.<br />

That’s a process called somatic hypermutation,<br />

and it happens during the last stages of<br />

the specialization of B cells.<br />

“Klaus Rajewsky found that HRS cells had<br />

undergone both types of changes,” he says.<br />

“That means they must have started out as<br />

mature B cells. But rather than finishing their<br />

specialization, they now start off on a strange<br />

developmental path. They start to produce<br />

molecules found in other types of blood cells,<br />

in T cells and macrophages.”<br />

The findings change how researchers have<br />

been thinking about both cell differentiation<br />

and cancer. “A common feature of these blood<br />

tumors is that cells reproduce at a high rate,”<br />

Bernd says. “Since that’s a feature of stem<br />

cells, a common assumption has been that<br />

the process of cancer starts high in the tree of<br />

development, in the early stem cell stages.<br />

Then mutations or other problems push the<br />

cells down a aberrant path. That’s the conclusion<br />

you come to if you assume that specialization<br />

is a one-way road, that once a cell has<br />

taken a step, there’s no going back.<br />

“But look at what we’ve found here. There is<br />

plasticity! HRS cells start at the end of B cell<br />

specialization, but they turn around and come<br />

part of the way back. Somewhere along the<br />

route they become pregenitor cells.”<br />

The implications go beyond cancer, he says. If<br />

a specialized cell can undo part of its programming<br />

and revert to a more generic state<br />

in cancer, could healthy cells be taught to do<br />

the same thing? “Think of what that would<br />

mean for regenerative medicine,” Bernd says.<br />

“We have assumed that in order for the body<br />

to repair or renew its tissues, it needs a stock<br />

of stem cells. But with age, those stocks get<br />

used up or die out, and some types of tissues<br />

can’t be rebuilt. The only solution would be to<br />

replace them. Unless you could find a way to<br />

take existing specialized cells, make them<br />

more generic again, and program them to go<br />

down a new pathway.” Understanding this<br />

escape from “terminal differentiation” would<br />

require a deeper look at the genetic programs<br />

at work in HRS cells.<br />

In 2001 Michael Hinz of Claus’ group used<br />

DNA chips to look at their total gene activity.<br />

Of the many thousands of genes that are<br />

active in the tumor cells, he identified a pattern<br />

of about 50 molecules which were under<br />

control of NF-ΚB. This “NF-ΚB signature” has<br />

been a big help as the researchers try to figure<br />

out the effects of the pathway in cancer cells.<br />

“Overactive NF-ΚB signaling has several<br />

effects that lead to cancer,” Claus says. “One<br />

characteristic of tumors is that cells aren’t<br />

able to stop dividing. The cell cycle is controlled<br />

by signals, some of which are passed<br />

along by NF-ΚB, as Michael Hinz found out<br />

earlier. Those signals should only switch on at<br />

the right times, but if something breaks<br />

down, the pathway can get stuck in ‘transmitting<br />

mode,’ and it keeps sending the message<br />

245 Part four: From the protein village to the cosmopolitan city<br />

Michael Hinz


to divide. With Stephan Mathas, an MD from<br />

Bernd’s group, we then searched for further<br />

signal-regulated factors that are out of control<br />

in the tumor cells. This led to the discovery<br />

of another growth promoting and gene-activating<br />

protein, called AP-1, which was always<br />

switched on in Hodgkin tumor cells. That molecule<br />

has been linked to control of cell division<br />

and it is known to be modulated by NF-ΚB.”<br />

Loss of this control is one of the worst things<br />

that can happen to cells, so they have a backup<br />

system to protect themselves. Tumors and<br />

other serious malfunctions often trigger a cellular<br />

self-destruct program called apoptosis.<br />

“That, too, is managed by signals,” Claus says.<br />

“As you can probably guess, some of them<br />

pass through NF-ΚB. So defects in this network<br />

can have two different effects that lead<br />

to the same end. They push the cell to divide;<br />

simultaneously they rob it of its ability to stop<br />

the process when it gets out of control.”<br />

A look at patient tissues by Bernd’s group<br />

revealed cases in which IkB proteins, the NF-<br />

ΚB brake, had undergone mutations. This<br />

echoed what Claus and his lab had found in<br />

the test tube: without the lock, cells lose control<br />

of NF-ΚB and become cancerous. Next,<br />

human tumor cells were transplanted to mice<br />

and Claus’ lab could prove that NF-ΚB was one<br />

player in triggering the development of<br />

tumors. Uwe Kordes, an MD who worked in<br />

Claus’ lab and now a specialist in children’s<br />

oncology, discovered the same basic problem<br />

with NF-ΚB in a frequent form of leukemia,<br />

and in 2005 Stephan found it in yet another<br />

type of lymph node tumor.<br />

NF-ΚB’s status as a signaling hub, Claus says,<br />

makes it attractive when trying to figure out<br />

how to cope with diseases such as these.<br />

“Most things that go wrong involve the braking<br />

system. That makes IKK proteins an interesting<br />

drug target. If you can keep them from<br />

receiving signals, they won’t release NF-ΚB to<br />

enter the nucleus. Arsenic is poisonous<br />

because it interferes with IKK, so we looked at<br />

what effect it would have on Hodgkin’s cells.


We also tried drugs that block the activity of<br />

Hsp90, the chaperone that helps shape IKK<br />

proteins so that they can receive signals. Both<br />

of these strategies prevented the activation of<br />

NF-ΚB. As a result a lot of the cancerous cells<br />

self-destructed.”<br />

Until Claus and Bernd turned their attention<br />

to the problem, the signals and pathways that<br />

changed gene activity in Hodgkin’s disease<br />

were unknown. Treatments for the disease<br />

have been based on radiation and chemotherapy.<br />

Uncovering the role of NF-ΚB and finding<br />

mutations linked to its activation were huge<br />

steps forward that suggested a clear strategy<br />

for new therapies involving IKK. In recognition,<br />

Claus and Bernd were awarded the 2005<br />

German Cancer Prize.<br />

Their findings suggest a more effective method<br />

of treating Hodgkin’s disease and other NF-ΚB<br />

associated cancers. That’s the goal, Claus says.<br />

That’s what we are all aiming for.<br />

��<br />

It has now been 18 years since Claus moved<br />

to Berlin, and when he talks about the city<br />

it sounds like he’s at home. He lives in<br />

Charlottenburg, an area of restaurants and<br />

cafés that he enjoys when he has the time.<br />

His other pasttime is sailing. Usually this<br />

means renting a boat on the Wannsee, but he<br />

has ventured much farther than that.<br />

Claus grew up in Schleswig, an old coastal<br />

town between the seas in Northern Germany.<br />

In college he joined a rowing club, and then<br />

didn’t have much time for water sports for a<br />

few years. When he moved to Berlin a colleague<br />

at the university, an MD, reminded him<br />

that the world consisted of more than just science,<br />

and they sneaked out for an afternoon<br />

every week to take sailing lessons. Eventually<br />

Claus received a license to sail on the high<br />

seas, and that led to trips to the Canary<br />

Islands and other parts of the Atlantic.<br />

Now he tries to help his own students keep<br />

perspective. “It’s a tough stage in their<br />

careers,” he muses. “I see young people here,<br />

spending 12 hours a day in the laboratory.<br />

Hurrying from one building to the next without<br />

noticing what a nice day it is. They have to<br />

work hard right now, but you still somehow<br />

have to keep a balance. Nobody teaches them<br />

how to do that. One of the jobs of the group<br />

leader is to help them through the tough<br />

times and to keep up their excitement. ”<br />

Does it take a certain kind of person to<br />

become a biochemist? If so, Claus admits he<br />

may be the archetype. His interest in the topic<br />

goes back a long way. “A few years ago I went<br />

to a school class reunion,” he says. “Somebody<br />

there reminded me that in high school I had<br />

given a talk on signaling. On synapses.”<br />

At the university he considered medicine but<br />

decided on chemistry. “It was about the most<br />

complicated thing I could find,” he laughs. “I<br />

wanted to do something hard. I got interested<br />

in molecules and their structures. I liked the<br />

sophistication of the methods.”<br />

He seems like the type of person who would<br />

learn the timetables of trains by heart, who<br />

would enjoy intricate puzzles. You couldn’t ask<br />

for a more complicated puzzle than unraveling<br />

the NF-ΚB network, and to work on it he<br />

has called in a wide range of expertise.<br />

Alongside biochemists, cell biologists, and<br />

chemists, the group now regularly works with<br />

experts in bioinformatics and mathematics.<br />

One reason is simply to keep track of all the<br />

components of the network and their interactions.<br />

“As long as we were dealing with a relatively<br />

small number of molecules, we found ways to<br />

represent their interactions,” he says. “But the<br />

hub-like nature of NF-ΚB and the fact that it<br />

can activate so many genes require much<br />

more complex models – it’s like going from a<br />

multiplication table to advanced calculus. Our<br />

theoretical partner in this is Jana Wolf, who<br />

has a small group that focuses on mathematical<br />

modeling of cellular processes.” Jana’s<br />

group has recently received funding under an<br />

initiative launched by the <strong>MDC</strong>’s parent<br />

organization, the Helmholtz Association, to<br />

support projects in “systems biology.” As well<br />

as collaborating with Claus, Jana is studying<br />

the Wnt/β-catenin signaling system that<br />

Walter Birchmeier’s group has worked on for<br />

so many years.<br />

Claus has known Jana for some years, and<br />

when they first met she told him they had a<br />

mutual acquaintance. “When she was in high<br />

school, she won a big math competition,” he<br />

says. “The first prize was a chance to meet this<br />

famous person at the Humboldt University. It<br />

was Jens Reich – now an emeritus of the<br />

<strong>MDC</strong>.”<br />

A factor that makes building a model of the<br />

NF-ΚB network so hard is the pace of new discoveries.<br />

“I used to be able to read every paper<br />

on NF-ΚB,” he says. “Now if you look at the scientific<br />

literature, you find over 20,000 papers<br />

on the subject – there are more than 2,000<br />

just on IKK alone. So it’s a major task to keep<br />

track of the data and to integrate it. To try to<br />

integrate it.”<br />

��<br />

One rainy afternoon in February, during a<br />

short respite from the strikes, Claus and<br />

I ride together into town on the S-Bahn. We<br />

talk about sailing for a while; then the talk<br />

turns inevitably to NF-ΚB. I’m getting a better<br />

idea of the molecule’s place in the cell signaling<br />

network, and its role in cancer. It’s such a<br />

central node that it is probably involved in<br />

guiding the development of embryos. I ask<br />

him about it.<br />

“Hair,” he laughs.<br />

Hair?<br />

247 Part four: From the protein village to the cosmopolitan city


Part four: From the protein village to the cosmopolitan city<br />

“At first it seemed trivial. Ruth Schmidt-Ulrich is our mouse specialist. She<br />

developed a strain in which NF-ΚB is generally supressed. We call it a<br />

‘super-repressor’ mouse. It has a version of IkB that can’t be broken down,<br />

so the lock is never taken off NF-ΚB. This mouse is interesting because it<br />

can be used to test hypotheses about NF-ΚB’s normal physiological activities.<br />

It is also useful in studying what happens when NF-ΚB is blocked –<br />

that’s necessary to discover harmful effects that might appear during<br />

treatments, so it serves as a model for therapeutic inhibition.”<br />

As Ruth studied the animals, she found predictable problems – defects<br />

in the immune system because of NF-ΚB’s role in immune cell regulation,<br />

and a high rate of cell death in some parts of the liver. Then she<br />

began noticing problems with the hair.<br />

“We had been hoping to find major developmental defects,” Claus says.<br />

“We already expected that NF-ΚB was needed for the development of<br />

lymph nodes and other structures formed by immune cells which<br />

makes sense because many of their biological functions depend on NF-<br />

ΚB. And there is a strong effect if NF-ΚB signaling breaks down. But otherwise<br />

there was nothing seriously wrong with the animals’ major<br />

body structures. A big disappointment.<br />

“Then Ruth brought me this mouse that had slanted eyes and sort of<br />

‘punk’ hair. My first reaction was, ‘We are now getting into hair<br />

research?’ I told her, ‘We’re not a molecular barber shop.’ But it was well<br />

worth taking a closer look. There were other problems with structures<br />

in the skin. Something was wrong with the follicles; it meant that the<br />

mice only produced one type of hair. And sweat glands didn’t develop<br />

properly. There were also problems with the fine structure of teeth –<br />

the structure of cusps of the molars.”<br />

A search of the medical literature revealed a similar combination of<br />

symptoms in humans called ectodermal dysplasia. The disease had<br />

been known for a long time – Charles Darwin was even familiar with it.<br />

One of his many correspondants, a Scottish civil servant working in<br />

India, had called it to his attention. In 1875 Darwin wrote:<br />

“I may give an analogous case, communicated to me by Mr. W.<br />

Wedderburn, of a Hindoo family in Scinde, in which ten men, in the<br />

course of four generations, were furnished, in both jaws taken together,<br />

with only four small and weak incisor teeth and with eight posterior<br />

molars. The men thus affected have very little hair on the body, and<br />

become bald early in life. They also suffer much during hot weather<br />

from excessive dryness of the skin. It is remarkable that no instance has<br />

occurred of a daughter being affected...though the daughters in the<br />

above family are never affected, they transmit the tendency to their<br />

sons: and no case has occurred of a son transmitting it to his sons. The<br />

affection thus appears only in alternate generations, or after long intervals.“<br />

248


Without knowing it, Darwin was describing a<br />

rare mutation of a gene on the X chromosome,<br />

which explained the pattern by which<br />

it was inherited. But it would be nearly 40<br />

years before fly geneticist Thomas Morgan<br />

found sex-linked traits and figured out why<br />

they occurred.<br />

Ruth’s discovery meant that NF-ΚB played a<br />

role in the development of hair and gland<br />

structures – as well as ectodermal dysplasia –<br />

in mice. And as the scientists talked about it,<br />

they realized that the condition might offer a<br />

new way to explore NF-ΚB’s functions.<br />

“If you think about it, you have in this tiny<br />

space in the skin a mini-organ with a complex<br />

structure involving stem cells that specialize<br />

to create multiple types and structures such<br />

as the hair follicle and glands,” Claus says.<br />

“That happens because of signals. We could<br />

use our tools to manipulate the system and<br />

watch the effects on the skin.”<br />

Further studies by Ruth and the rest of the<br />

team have shown that NF-ΚB and other components<br />

of the machine work together in subtly<br />

different ways to create different parts of<br />

skin structure. In the early embryo NF-ΚB is<br />

activated to produce placodes – thickened tissue<br />

that will later form structures in the skin.<br />

It is also needed later, as hair follicles form, to<br />

stimulate cells to divide and push downward<br />

through other layers of skin.<br />

“It has been very hard to determine NF-ΚB’s<br />

role in any developmental process in mammals,”<br />

Claus says. “This is one of the very few<br />

cases in which that could be done in detail. By<br />

studying this mini-organ, Ruth and the lab<br />

have identified the signals upstream that<br />

guide this process and what part of the<br />

machine they act on. For example, we know<br />

that the downward movement of the placode,<br />

which creates the follicle, is controlled<br />

by signals that come through β-catenin.<br />

That’s an important topic of Walter<br />

Birchmeier’s group, so we’ve been able to pursue<br />

it with him. At the same time, close by in<br />

other parts of the skin, IKK proteins direct<br />

some processes that don’t involve NF-ΚB.”<br />

As the S-Bahn rolls toward the city center, he<br />

talks about his lab’s future plans. Working out<br />

the functions of the NF-ΚB pathway will take<br />

many more years. Most of the work will<br />

involve hammering out intricate details of cell<br />

signaling. Sometimes it’s hard to see the big<br />

picture, he admits. But if you want to deeply<br />

understand how cells manage the huge variety<br />

of signals that cross paths on their way to<br />

genes, NF-ΚB is an excellent place to start.<br />

One reason is the pathway’s involvement in<br />

Hodgkin’s disease and other forms of cancer. If<br />

you want to reroute the signals that are<br />

responsible, without disrupting something<br />

that healthy cells need, you’ll have to know<br />

the details.<br />

My stop comes first and we say goodbye.<br />

From the platform I watch the train slide on to<br />

the city. Claus will take the S2 to Friedrich -<br />

strasse, then switch to a westward line for<br />

Charlottenburg. That will work today, anyway.<br />

If the strike begins again tomorrow, he’ll have<br />

to find another route. On my way out of the<br />

station I take another look at the transportation<br />

map, with its crossing S-Bahn and U-<br />

Bahn lines. Today, for some reason, it doesn’t<br />

seem so complicated.<br />

249 Part four: From the protein village to the cosmopolitan city


The pleasures and<br />

powers of green tea<br />

In 1973, a year after receiving his law degree, businessman Philippe<br />

Neeser left his home in Geneva to study in Osaka, Japan. A few years<br />

later he experienced a life-changing moment when he was introduced<br />

to the traditional Japanese tea ceremony. Over the next three decades,<br />

Neeser became an internationally recognized master of this tradition<br />

which is part art, part philosophy, part religion.<br />

History says that green tea had been known in Japan for several centuries<br />

but remained a drink of aristocrats until 1191. In that year the<br />

priest Myoan Eisai returned from a trip to China, bringing with him the<br />

Zen Buddhist religion and tea seeds. He used the brew as a stimulant to<br />

keep monks awake during their long hours of meditation. Three hundred<br />

years later Marata Shuko developed the custom of serving tea as a<br />

means of establishing spiritual harmony between a host and a guest.<br />

Sen-no Rikyu, a disciple of one of Shuko’s disciples, developed the practice<br />

into a highly ritualized ceremony. Rikyu wrote poems about tea in<br />

the deceptively simple language of Zen:<br />

Part four: From the protein village to the cosmopolitan city<br />

250


Eric Wanker


Tea is not but this:<br />

First you make the water boil,<br />

Then infuse the tea.<br />

Then you drink it properly.<br />

That is all you need to know.<br />

Thus the history of tea in Japan is inseparable from religious tradition.<br />

It is equally intertwined with medicine. Eisai wrote a two-volume book<br />

whose title translates as “How to stay healthy by drinking tea.” The<br />

first sentence reads, “Tea is the ultimate mental and medical remedy<br />

and has the ability to make one’s life full and complete.” Among the<br />

curative powers he cites: improving the functions of the brain, bladder<br />

and heart, easing thirst, and curing hangovers.<br />

Part four: From the protein village to the cosmopolitan city<br />

If anything, the ancient Japanese may have underestimated the medical<br />

value of green tea. At the beginning of the 21st century, scientists<br />

like Erich Wanker of the <strong>MDC</strong> are learning that its active substances<br />

may have many more beneficial effects on health, and they are beginning<br />

to understand why.<br />

252<br />

��<br />

That a poet and priest should also be a healer is no contradiction<br />

in Buddhist tradition; it is much more unusual in the West, in<br />

modern times. Yet a few people like Philippe Neeser manage to fuse<br />

radically different facets of modern existence. When he is not serving<br />

tea, he is head of legal affairs for Ciba Chemical Specialties, a Swiss<br />

company with offices in Japan. In 2005 he guided Japanese Emperor<br />

Akihito and his wife through the high-tech Swiss pavilion of the World<br />

Exposition, which was being held in Japan. Three years earlier Neeser<br />

had received another tremendous honor when he was selected to<br />

carry out a ceremony reserved for the greatest tea masters. He served<br />

tea to the great Buddha of Nara during celebrations of the statue’s<br />

1450th anniversary. The Buddha weighs 500 tons and has fingers the<br />

size of a human being. Building the statue and its temple required the<br />

efforts of over two million people, wrecked the Japanese economy, and<br />

used up most of the country’s copper.<br />

Erich Wanker’s path to the <strong>MDC</strong> is not quite that unusual, but it has<br />

still involved a big cultural leap for the young Austrian researcher. “I<br />

come from a small village in the Alps,” he says. “I had virtually no contact<br />

with real science as I grew up. Of course I had biology, physics and<br />

chemistry in high school. But there were no role models – scientific<br />

research was happening ‘out there’ in some higher, remote sphere. In<br />

spite of that, alongside sports, I always had an interest in nature.”<br />

He says that at the age of 16 or 17, that interest blossomed into an<br />

interest in “hard-core” sciences like engineering and chemistry, accompanied<br />

by a desire to make an impact on the world. At the university<br />

he started off with chemical engineering and then became attracted<br />

to biochemistry and molecular genetics.<br />

“I was fascinated the first time I heard of enzymes,” he says. “There<br />

were people in the world who were making real discoveries, really finding<br />

out how life works. That was for me – I had to do that. But it wasn’t<br />

easy to explain this to my family and friends. There had never been<br />

a doctor or a professor in our family, or even in the village. I had no idea<br />

whether I was doing the right thing. But at some point there was no<br />

turning back.”<br />

He laughs as he remembers a conversation with his parents. “This was<br />

after I had had some successful publications and it was clear I was<br />

going to make it as a scientist,” he says. “They were worried, the way<br />

parents always worry about their kids. They shook their heads and<br />

asked, ‘Are you ever going to be able to make a living doing this?’”


He ticks off the stages of his career. As he speaks his hands move rapidly<br />

and his blue eyes flash. It’s not hard to imagine that a small<br />

Austrian village would have trouble containing all that energy.<br />

He received his PhD in Biochemistry from the Technical University of<br />

Graz in 1992, then it was off to the University of California in Los<br />

Angeles to do postdoctoral work in the lab of David Meyer. One theme<br />

of the lab concerned the processing of proteins that cells secrete or<br />

put in their membranes. These molecules are synthesized from RNAs<br />

at the surface of a structure called the endoplasmic reticulum, or ER.<br />

Getting there requires transport molecules that recognize and dock<br />

onto a receptor protein. Erich and his colleagues helped to identify the<br />

receptor and show how it docked onto the protein-synthesizing<br />

machinery. It was an important finding, but he had his sights set on<br />

bigger things.<br />

“I had decided to move out of the basic cell biology I had done at the<br />

Meyer lab and into disease research because I felt a need to work in an<br />

area that had a direct impact on more urgent human questions,” he<br />

says. “Coincidentally, the gene that causes Huntington’s disease had<br />

just been discovered. In late 1994 I started looking for a group leader<br />

position, and heard that Hans Lerach – one of the people who had<br />

found the gene – was looking for a group leader in Huntington<br />

Disease research.”<br />

Huntington’s is a brain disease that develops because of the premature<br />

death of particular types of neurons. Symptoms include jerky<br />

movements and a general loss of control of the body, accompanied by<br />

changes in mental abilities and behavior. The problems usually appear<br />

in middle age and grow progressively worse over time. As in<br />

Alzheimer’s disease, the problem seems to stem from protein fragments;<br />

they form clumpy fibrils that do not dissolve. Whereas in<br />

Alzheimer’s the fibrils form in the cell cytoplasm and the space<br />

between cells, in Huntington they mainly appear in the cell nucleus,<br />

with some clusters in the cytoplasm.<br />

Only a small percentage of Alzheimer’s cases are clearly hereditary,<br />

linked to variants of genes which are described in the story “A Swiss<br />

253 Part four: From the protein village to the cosmopolitan city


Alexandra Redel<br />

Army knife of the cell.” But Huntington’s follows a strict hereditary pattern.<br />

It is caused by a dominant mutation in a gene, which means that<br />

a person needs to inherit only one copy of the molecule to develop the<br />

disease. For a long time scientists had been trying to identify the<br />

culprit.<br />

Hans Lehrach’s group at the Imperial Cancer Research Center Fund in<br />

London and other laboratories had steadily been closing in on a region<br />

of the fourth human chromosome. The gene was finally identified in<br />

1993 by the Huntington’s Disease Collaborative Research Group, an<br />

international collaboration between six laboratories, including that of<br />

Lehrach. They named it huntingtin, with an “i” – a common ending for<br />

proteins – ensuring decades of spelling problems for biology students.<br />

In 1994 Lehrach moved to the Max Planck Institute for Molecular<br />

Genetics in Berlin to head a new department and began hiring new<br />

groups. Erich applied and was initially given a one-year position, which<br />

was then extended. “At the beginning it was just me and one student,”<br />

he says. “That’s about the smallest group you can get.”<br />

As Erich began setting up his new lab, scientists were discovering<br />

some of the differences between the dangerous and healthy forms of<br />

the huntingtin protein. Like all proteins, huntingtin is spelled of chemical<br />

subunits called amino acids. At the “head” of the molecule is a long<br />

stretch consisting of just one letter, glutamine, repeated over and over,<br />

as if a key on the computer has become stuck. In healthy people this<br />

string is between 11 and 23 letters long, but if the molecule has 40<br />

repeats, a person is highly likely to develop the disease. Because the<br />

chemical abbreviation for glutamine is “Q”, researchers call the region<br />

a polyQ sequence.<br />

While the gene responsible for Huntington’s disease has been known<br />

since the early 1990s, no one knew why it caused problems until 1997.<br />

Part four: From the protein village to the cosmopolitan city<br />

Pablo Porrasmillan<br />

That year Erich and his lab were the first to discover that huntingtin<br />

accumulated in clumpy aggregates. “That discovery fundamentally<br />

changed the way HD is classified,” Erich says. “From then on, it could be<br />

grouped with amyloid diseases like Alzheimer’s. And it spawned an<br />

enormous amount of research into the aggregation of proteins with<br />

polyQ motifs, leading to the discovery that nine further diseases that<br />

could be classified as polyQ diseases.”<br />

The discovery of the aggregates was very intriguing, but it didn’t say<br />

anything about why one form of the molecule led to problems for cells<br />

or the brain. And it said nothing about why that problem starts to<br />

become serious when a person reaches middle age – until then, the<br />

body somehow copes.”<br />

There were many possibilities, he says. The defective form of the molecule<br />

might have an altered architecture, allowing many copies of<br />

huntingtin to cluster in a way that couldn’t be dissolved. Or different<br />

proteins might dock onto the two types of huntingtin, folding one of<br />

them into a dangerous form. Scientists weren’t even sure why the protein<br />

was toxic. The questions were similar to those posed by Bernd Reif<br />

in the story “Bad origami:” the protein fragments that cause<br />

Alzheimer’s and Huntington’s disease can accumulate in different<br />

ways, forming clusters of twos, threes, fours, or much larger complexes<br />

that go on to form super-complexes. The most dangerous form<br />

might be a grouping that appeared somewhere along the way and<br />

then adopted another form.<br />

“The fact that symptoms normally arise between the ages of 40 and<br />

50 means that something has changed at that point in life,” Erich says.<br />

“With age, just as with some kinds of stress, the cell begins losing its<br />

ability to cope with the continuous challenge of dealing with a misfolded<br />

protein. This suggested we ought to look at other players – the<br />

proteins that interacted with huntingtin – to discover the mecha-<br />

254


nisms of misfolding and toxicity. But 15 years ago we didn’t have great<br />

tools to discover what proteins interacted with each other. We were<br />

going to have to spend time developing that technology.”<br />

Erich’s personal history has taken him from a small village in the Alps<br />

to a major world capital. Now science was about to take him from a<br />

single molecule to the big city of cell networks.<br />

��<br />

During the 1990s improvements in DNA sequencing technology<br />

had suddenly given researchers a huge list of new genes in<br />

humans and other organisms. Most of these molecules had never<br />

been seen in experiments, and no one knew what jobs they performed<br />

in cells. “We were all looking for shortcuts to determining their functions,”<br />

Erich says. “Figuring out what molecules they interacted with<br />

would go a long way toward that goal.”<br />

In 1989 Stanley Fields and Ok-Kyu Song of the State University of New<br />

York invented a method called a two-hybrid system that used yeast<br />

cells to “fish” for proteins that can dock onto each other. They began by<br />

removing a gene that the cells needed to survive. They replaced it with<br />

a new version of the gene that could be triggered by an artificial protein.<br />

The protein had two modules: one that docked onto the gene and<br />

activated it; another that acted as a bait. The trigger would only be<br />

pulled if another protein (the prey) took the bait, and only then would<br />

the yeast cell activate the gene and survive.<br />

Fields’ and Songs’ clever strategy was to give one strain of yeast the<br />

combination of bait and trigger, and another strain the prey. Then they<br />

allowed the two strains to mate. If this brought together two proteins<br />

that could bind, the yeast colony survived, and the results could be<br />

seen with the naked eye.<br />

By doing this across the entire genome, the method would theoretically<br />

give you a map of all the possible interactions between proteins. But<br />

creating tens of thousands of yeast clones containing baits and others<br />

with the prey would be a huge amount of work – unless the procedure<br />

could be automated. So Erich and his group began building the robots<br />

and tools needed to carry out yeast two-hybrids on a massive scale.<br />

This was crucial for the creation of the first map of protein interactions<br />

across the whole human genome. Today his group has become so proficient<br />

at the methods that when they have the time, they offer them<br />

as services to other labs.<br />

The resulting map is an important first step in charting interactions<br />

between proteins. “Whether a pair actually binds in cells is another<br />

question,” Erich says. “Even if their chemistry makes them a perfect fit,<br />

two molecules may never get the chance to interact because they<br />

occupy different regions of the cell – or are never made together in the<br />

same type of cell. So at the end of a two-hybrid experiment you have<br />

a lot of data that needs to be verified through other types of experiments.”<br />

The mid-1990s saw the development of another technology that<br />

could help out, mass spectrometry. It can be used to identify the proteins<br />

in small samples taken from cells. While part of Erich’s growing<br />

lab developed the two-hybrid system, he put others to work on mass<br />

spectrometry.<br />

The two-hybrid system provided a list of candidates that could bind to<br />

huntingtin, but it didn’t answer other questions: how huntingtin<br />

assembled into fibrils and what made them toxic. Other methods<br />

developed in the lab were beginning to provide insights into these<br />

questions. A filter method helped them trap large fibers that didn’t<br />

dissolve; Erich hoped that these large clumps would contain other<br />

molecules that played a role in their formation.<br />

255 Part four: From the protein village to the cosmopolitan city<br />

Anne Wagner


And what was it in huntingtin itself that allowed it to snap together<br />

in groups, rows, and fibers? At a meeting he heard a talk from Max<br />

Perutz, the great pioneer of protein structures. The clustering of huntingtin<br />

proteins posed interesting questions that had attracted Perutz’s<br />

attention.<br />

While working in Cambridge in the late 1940s and 1950s, Perutz had<br />

found a way to determine the shapes and architecture of proteins<br />

loading purified proteins with atoms of metal and exposing them to<br />

X-rays. The method allowed Perutz and his student John Kendrew to<br />

obtain structures of the first two proteins and earned the men a Nobel<br />

Prize in 1962. It was the same year that their close colleagues in<br />

Cambridge, Francis Crick and James Watson, received the award for<br />

deducing the double-helix structure of DNA. The two discoveries<br />

launched modern molecular biology.<br />

Perutz tells the story of his colorful life in a collection of autobiographical<br />

essays called I Wish You’d Made Me Angrier Sooner. A native of<br />

Austria, like Erich, his family had fled the country during the Nazi<br />

takeover in 1938. During the war, he was briefly interred in a Canadian<br />

prison camp until a group of important scientists intervened on his<br />

behalf. Upon his release the British government put him to work on<br />

the war effort. Because Perutz had once carried out a study of the<br />

behavior of ice crystals in glaciers, he was enlisted in a crazy project to<br />

Part four: From the protein village to the cosmopolitan city<br />

build artificial icebergs to be used as massive floating aircraft carriers.<br />

It didn’t work, Perutz said later, because the combination of wood pulp<br />

and ice used to make the icebergs... melted.<br />

“When I met him, Max Perutz was in his eighties but was still going<br />

strong working on proteins,” Erich says. “At the conference he gave a<br />

talk about huntingtin and figuring out what disease mutations did to<br />

its structure. He had the idea that if two molecules lay side by side but<br />

inverted, so that the head of one protein lay next to the tail of the<br />

other, there would form a kind of zipper between the glutamine<br />

repeats – the polyQ sequence. That’s why these big clusters would form.”<br />

Erich approached Perutz and said that some of the data his group had<br />

been collecting might support the hypothesis. That encounter led to a<br />

long-term collaboration that continued until Perutz’s death in 2002.<br />

Erich is a co-author on some of the scientist’s last papers. One of them<br />

expanded the huntingtin findings to other molecules.<br />

“There are many proteins that accumulate in and between cells and<br />

cause disease,” Erich says. “Several of them contain repeats of amino<br />

acids. Even though the recipe of each of these proteins is different,<br />

Max wondered whether a single underlying principle might be<br />

responsible for making them behave the same way. For example, a protein<br />

in yeast – similar to the prion molecule that causes mad cow dis-<br />

256


The current map of the “Human Interactome,” as depicted by a database that Erich’s<br />

group has developed. Each dot represents a protein. Lines connect the molecules that<br />

directly interact with each other.<br />

ease – contains repeats of the amino acid asparagine. Prion diseases<br />

also involve the accumulation of molecules between cells. Max<br />

showed that adjacent asparagines can zip up to each other in much<br />

the same way as glutamines.”<br />

The paper also showed how regions of the β-amyloid protein fragment<br />

– which collects between cells in Alzheimer’s disease – can link<br />

up to other proteins into β-sheet structures that lead to fibers.<br />

The work shows how sensitive some proteins are to changes.<br />

Increasing the concentration of a molecule in the space between cells,<br />

or a switch in one letter of its recipe, can have huge effects on its<br />

behavior. “A protein folds in a certain way because the amino acid<br />

string is loaded with energy,” Erich says. “Chemical interactions inside<br />

the molecule stabilize it. But it achieves that stability because of its<br />

complex equilibrium with everything else in the environment. If something<br />

changes, you may quickly see a sort of chain reaction of misfolding<br />

and aggregation – at least that’s how we imagine the process.”<br />

��<br />

So ultimately, understanding the behavior of a protein requires<br />

understanding its place in the whole, a network of protein interactions<br />

that change all the time as cells carry out their functions and<br />

produce new molecules. The connections on the map are redrawn as<br />

cells grow and develop, when they come under stress from the environment,<br />

and as a natural result of growth and aging. The protein<br />

interaction maps of healthy cells are different than those affected by<br />

a disease. If you knew how the network normally behaved, you could<br />

monitor the impact of things like mutations or drugs.<br />

“That’s important because if you want to fix a problem with the drug,<br />

you have to fix the network and not just what’s happening to one molecule,”<br />

he says. “You’re not just asking, ‘Will protein X bind to protein Y<br />

and stop some sort of bad behavior?’ ...You have to be sure that by fixing<br />

one thing, you’re not disturbing something else.”<br />

Experiments from Erich’s lab and other groups throughout the world<br />

were beginning to fill in the map of protein interactions – like going<br />

from the ancient Greeks’ map of the world to that of Vasco da Gama,<br />

Erich says. “Those maps completely changed navigation and provided<br />

a basis for exploring the world,” Erich says. “Having a complete chart of<br />

human protein interactions – if it were high quality – would be the<br />

equivalent, it would be a fantastic tool for every kind of biological exploration.<br />

So developing this map has been a major focus of our efforts.”<br />

By 2005, systematic studies with the yeast two-hybrid system had<br />

given the lab a list of about 3,200 interactions between proteins, most<br />

of which had never been detected in other experiments. Within the<br />

257 Part four: From the protein village to the cosmopolitan city


next two years, 3,000 more interactions were added to the map. All of<br />

the information went into a database that the lab maintains on the<br />

Internet. The data is only the beginning, however; a major effort of the<br />

group has to verify that the proteins actually come into contact in<br />

cells, using other types of experiments.<br />

Erich shows me one of the maps, on which proteins are shown as<br />

round balls, connected by lines which reveal their interactions. “Can<br />

anyone read this?” I ask him.<br />

“That’s a big issue,” he agrees. “You can’t use it if you can’t read it. A lot<br />

of our effort has gone into designing tools to help scientists query the<br />

information in various ways. That begins with simple things, like looking<br />

up a protein you’re interested in and looking up the molecules it’s<br />

known to interact with. You can see whether it belongs to more than<br />

one signaling pathway in the cell. You can put in the names of two<br />

molecules and ask whether they bind to any common partners. You<br />

can zoom in from a protein’s role in the whole network to its functions<br />

in one tissue, at one developmental stage, in one disease.”<br />

Part four: From the protein village to the cosmopolitan city<br />

Ideally, Erich says, following a protein’s interactions will suggest new<br />

links to a cell signaling pathway or a disease process. “It’s a hypothesisgenerator,”<br />

he says. “Once you know where to look, you can confirm a<br />

protein’s role through other experiments, such as checking for mutations<br />

in patient tissues.”<br />

The value of the tool depends on the quality of its contents, for example<br />

by cutting down on the number of predicted protein interactions<br />

that turn out to be false. Erich and his colleagues are trying to teach<br />

the computer to recognize some of these cases itself, by presenting it<br />

with “gold standard” lists of proteins that interact and those that<br />

don’t.<br />

258<br />

��<br />

After a decade of work on protein interactions, everything is coming<br />

together in Erich’s current work on huntingtin. The group has<br />

found intriguing new leads to the causes of disease and points of<br />

attack for potential therapies. Initial experiments with the protein


showed that it might bind to hundreds of molecules. By 2004 postdoc<br />

Heike Göhler and other members of the lab had whittled the list down<br />

to 86 proteins that could link to huntingtin in at least 188 different<br />

ways. Most of the connections had never been seen in other experiments.<br />

Did these interactions have anything to do with the formation of fibrils?<br />

Heike and her colleagues did another screen in mammalian cells.<br />

This time they looked for proteins which, when brought together with<br />

huntingtin, made it aggregate.<br />

The story “A Swiss Army knife of the cell” recounts how Thomas<br />

Willnow’s group found that a molecule played a role in Alzheimer’s<br />

disease by drawing a membrane protein into the cell and sending it to<br />

a particular internal compartment. One of the proteins investigated by<br />

Heike, called GASP2, has a similar role in sorting membrane proteins. It<br />

binds to huntingtin in mammalian cells and moves through the cell<br />

with it.<br />

Another interesting molecule has emerged from a collaboration with<br />

Hitoshi Okazawa’s group at the Tokyo Medical and Dental University in<br />

Japan. “The Okazawa lab has established a cell culture model of<br />

Huntington’s disease,” Erich says. “I told you that mutant forms of<br />

huntingtin move into the nucleus, where they form fibrils. As they do<br />

so they trap molecules that are necessary for the cell to go about its<br />

normal business. Okazawa and his colleagues have been trying to<br />

understand how huntingtin gets into the nucleus and what happens<br />

once it gets there. We jumped in to look at other proteins that bind to<br />

huntingtin and to see if they have any influence on these processes.”<br />

The scientists compared cells with mutant forms of the protein to<br />

those with healthy huntingtin, and cells in which fibrils accumulated<br />

and others in which they did not. One pattern emerged over and over:<br />

whenever mutant proteins were present, levels of other molecules<br />

called HMGB1 and HMGB2 dropped.<br />

“These are incredibly important proteins,” Erich says. “They’re the most<br />

common molecules in the nucleus and you find them everywhere on<br />

259 Part four: From the protein village to the cosmopolitan city


the evolutionary tree. They are huge multi-functional tools that can<br />

bind to DNA. They also act as adaptor plugs for all kinds of proteins<br />

that need to attach themselves to DNA to repack it, repair it, activate<br />

genes... the list goes on and on. When mutant huntingtin is around,<br />

HMGB molecules seem to get stuck, and there are far fewer of the<br />

molecules floating around in the nucleus to do their jobs.” When the<br />

scientists checked cells taken from normal and diseased brains, they<br />

found the same connection.<br />

The results were exciting because they suggested that HMGBs might<br />

make good drug targets. When the scientists added extra copies of the<br />

molecules to cells with mutant proteins, cells normally affected by the<br />

disease were able to survive as well as healthy ones.<br />

The study revealed that HMGBs have another role that may play a<br />

direct role in Huntington’s disease. In healthy cells the proteins seem<br />

to pass signals that tell the cell when to stop dividing and repair its<br />

DNA. If mutant huntingtin proteins stop the proteins from doing this<br />

job, the timing may be thrown off and the cell will die. Adding HMGBs<br />

to the cells can restore the proper timing of the signals.<br />

��<br />

As Erich’s lab has learned more about the mechanisms that underlie<br />

Huntington’s disease, they have been searching for drugs or<br />

other substances that change the way cells handle the defective huntingtin<br />

protein.<br />

One strategy involves trying to bolster the cell’s natural defenses<br />

against problems. “When cells experience heat or other types of stress,<br />

one thing they do is produce more chaperone proteins,” Erich says.<br />

“These are ‘folding assistants’ whose jobs are to make sure that proteins<br />

are being made properly. Proteins are very sensitive to heat or<br />

other types of stress, and one of the first things that is likely to go<br />

wrong is for them to misfold.”<br />

An antibiotic called geldanamycin (GA) binds to a chaperone called<br />

Hsp90 and triggers the cell to respond to heat. “In one of our earlier<br />

studies we showed that treating cells with GA causes the cell to produce<br />

chaperones,” Erich says. “This stops the aggregation of huntingtin<br />

– the more antibiotic you add, the fewer the big clusters produced<br />

by the cell. But GA is so toxic that using it in a therapy would<br />

cause more harm than good.” Postdoc Martin Herbst has been working<br />

with chemically altered forms of GA that might be more useful.<br />

“Our studies turned up about 300 substances that reduce the clustering<br />

of huntingtin in the test tube,” Erich says. “One of these belonged<br />

to a family of molecules called the flavonoids, which are natural substances<br />

known to have a variety of health benefits.”<br />

Which brings us back to green tea.<br />

Part four: From the protein village to the cosmopolitan city<br />

The plant contains a flavonoid called epigallocatechin-gallate (EGCG).<br />

Alongside centuries of use as a traditional medicine in China, Japan<br />

and elsewhere, there is mounting new evidence that EGCG and other<br />

substances in green tea have anti-tumor effects. They may also help<br />

the body with other problems including microbial infections, allergies,<br />

and heart disease.<br />

Originally the health benefits of flavonoids were attributed to their<br />

“anti-oxidant” properties. They soak up overcharged atoms of oxygen,<br />

which are found in cancer and other diseases. These atoms undergo<br />

frenetic interactions with biological molecules, often breaking them<br />

apart. But some recent findings suggest that this effect may be much<br />

stronger in the test tube than in living organisms.<br />

Erich wondered if EGCG might interfere with the clustering of huntingtin<br />

proteins. That’s what his lab found in experiments in the test<br />

tube, cells and fly models. “EGCG potently blocks the process by which<br />

the repeat sequences in huntingtin ‘zip up’ to each other,” he says. “You<br />

remember Max Perutz’s hypothesis that such repeats in other proteins<br />

might also make them cluster and cause disease? That made us think<br />

that if EGCG was helpful in one case, it might be helpful in others. So<br />

we have tested β-amyloid, responsible for Alzheimer’s, and a protein<br />

called α-synuclein, which is found in Parkinson’s disease. EGCG prevents<br />

them all from accumulating into large fibers. So this natural substance<br />

recognizes a variety of proteins that might collect in large intracellular<br />

clumps and prevents them from doing so.”<br />

How does it manage this? “It doesn’t make the proteins go away,” Erich<br />

says. “And they do still cluster. But when we used the electron microscope<br />

to look at aggregation reactions that had happened in the test<br />

tube, we saw that α-synuclein and β-amyloid had grouped into small<br />

particles and then stopped. They didn’t go on to make the big, cumbersome<br />

particles that don’t dissolve. We followed up with a second<br />

experiment in which we added the different types of particles to cells.<br />

EGCG produced small particles that didn’t harm the cells, while the<br />

bigger fibril structure did.”<br />

The proteins go from a single molecule to a cluster containing various<br />

numbers of proteins and then big fibers. Somewhere along the way<br />

they kill neurons. “Early in the clustering process, EGCG steps in and<br />

groups the proteins in a new way,” he says. “That’s a detour from the<br />

normal path. EGCG binds to the unfolded molecules and prevents<br />

them from lining up in a β-sheet. That is a prerequisite for the formation<br />

of larger, toxic clusters.”<br />

It’s not a miracle cure, Erich cautions. Like most natural substances,<br />

EGCG and other flavonoids will have to be purified, developed, and<br />

extensively tested before they can be used in therapies.<br />

But EGCG has a lot going for it. In the first place, people have been<br />

drinking it for thousands of years – apparently with mostly positive<br />

260


effects. Secondly, it has already been approved for clinical testing in<br />

other diseases. Frauke Zipp of the <strong>MDC</strong> has recently started a clinical<br />

trial using EGCG in the treatment of multiple sclerosis. And maybe<br />

best of all, EGCG can slip out of the bloodstream and into the brain,<br />

which means it can reach cells in distress. Most other drugs fail at this<br />

because of a bodily defense mechanism called the “blood-brain<br />

barrier.”<br />

Erich and his colleagues aren’t just standing by, waiting to be discovered<br />

by pharmaceutical companies interested in turning their findings<br />

into new therapies. Once he felt the science was solid enough, Erich<br />

put together a major proposal and submitted it to the German Federal<br />

Ministry of Education and Research (BMBF). The program “GO-Bio” was<br />

created in 2006 when the BMBF realized that the relationship<br />

between basic research, pharmaceutical companies and clinics was<br />

changing. Scientists were having to work much harder to attract interest<br />

in promising projects and bring them to the marketplace. One<br />

option that researchers have is to start a company, but many small<br />

biotechnology firms fail because it usually takes many years to trans-<br />

form a discovery into a product. As a result, a huge number of promising<br />

leads are never taken beyond the lab.<br />

As a sort of mix between a research grant and a company start-up<br />

investment, GO-Bio supports very promising projects at an early stage<br />

so that researchers will have more time to develop their findings.<br />

Erich’s project was regarded as so promising that his lab was one of<br />

just a few selected for funding in the initial launch of the program. The<br />

money goes into an intensive search for drugs and substances – such<br />

as the molecules of green tea – that can be used as therapeutic or<br />

diagnostic compounds in the treatment of Alzheimer’s or<br />

Huntington’s disease.<br />

With all of this going on, you might expect his lab to resemble the<br />

Berlin train station on the first day of school vacations, or a Starbuck’s<br />

coffee shop on a Saturday afternoon. But the atmosphere in the group<br />

seems calmer, somehow. Maybe it’s the steadily falling levels of<br />

caffiene. Erich’s team hasn’t started drinking only green tea – not yet.<br />

But you see a lot of teapots and teacups. They seem to have become<br />

standard equipment at the bench.<br />

261 Part four: From the protein village to the cosmopolitan city


Playing the piano<br />

of planaria<br />

It seems to take more than ten fingers to play an étude by<br />

Alexander Scriabin – a pianist’s hands are constantly at work,<br />

evoking turbulent shadows that sweep by in grand gestures,<br />

exposing broken moments of light. Nikolaus Rajewsky is playing<br />

Scriabin to get to know the grand piano upstairs in the <strong>MDC</strong><br />

Communications Center. Watching his hands, it’s impossible for a<br />

non-pianist to imagine the feeling in those fingers, the combination<br />

of discipline and wildness needed to play such a piece.<br />

Moving on to Chopin, Nikolaus reaches most of the instrument’s<br />

88 keys. This is the standard number, but a few pianos have more;<br />

the Austrian firm Bösendorfer has made instruments with up to<br />

nine more tones in the bass. To keep from confusing performers,<br />

on some models the extra keys can be covered with a small lid. On<br />

others the colors of the lowest notes are reversed, replacing white<br />

ivory keys with black ebony ones. While modern music exploits the<br />

full range of the classical piano, few pieces require the extra notes.<br />

The long strings are there to add depth to the sound as they resonate<br />

with other tones.<br />

Part four: From the protein village to the cosmopolitan city<br />

262<br />

Nikolaus Rajewsky


Most biological processes make use of the full<br />

keyboard of the cell, Nikolaus says – meaning<br />

that they often involve dozens or hundreds of<br />

genes, RNAs and proteins. But until very<br />

recently, scientists haven’t been able to hear<br />

all the notes or see the keys that produced<br />

them. The discovery of microRNAs in the<br />

1990s was like suddenly discovering that<br />

pianists had been leaving out notes while<br />

practicing their scales.<br />

“For 50 years molecular biologists have been<br />

scrutinizing the process by which the information<br />

in genes is used to produce proteins,”<br />

Nikolaus says. “At every step along the route<br />

we have discovered ways that cells intervene<br />

to modulate the production of a molecule or<br />

alter it to change its functions. So it was<br />

amazing to find that there was an entire level<br />

of control in the cell that we had never been<br />

aware of. The question now is how widely<br />

Part four: From the protein village to the cosmopolitan city<br />

cells use this mechanism in processes like<br />

building a body, and what role it plays in disease.”<br />

MicroRNAs have been invisible to most of the<br />

tools used to investigate molecules. Rather<br />

than encoding proteins, like genes, they dock<br />

onto other RNAs and signal for them to be<br />

silenced. This process, described in the chapter<br />

“A very quiet cure,” represses the target<br />

RNAs from being used to make proteins. It’s<br />

hard to detect something by its absence – it<br />

would be like listening for the fingers that<br />

aren’t used during a particular measure of<br />

Scriabin’s études.<br />

But miRNAs have other functions which are<br />

even harder to grasp. Sometimes they block<br />

the production of a protein without destroying<br />

an RNA; instead, they obstruct the work of<br />

the protein-building machinery. Trying to<br />

observe this activity of the tiny molecules<br />

264<br />

across the level of the genome, Nikolaus says,<br />

has been a huge methodological challenge.<br />

Another problem is that the computer analysis<br />

of genomes, which has successfully identified<br />

a huge number of new genes, was initially<br />

not very successful at finding microRNAs.<br />

“In a way, they were too successful,” Nikolaus<br />

says. “Most of these original attempts led to a<br />

huge number of ‘false positives.’ At least half<br />

of the sequences predicted to be microRNAs<br />

in these early studies probably aren’t.”<br />

Finding the molecules and understanding<br />

their functions will require a combination of<br />

experiments in the test tube, cells, and organisms<br />

– and completely new ways of analyzing<br />

huge amounts of data. When Nikolaus arrived<br />

at the <strong>MDC</strong>, he brought along a vision for a<br />

new type of lab that would marry the<br />

approaches. That fit in well with the campus’<br />

vision for the Center for Medical Genomics.


He found himself in good company – one floor<br />

down is Norbert Hübner, who is taking an<br />

interdisciplinary approach to the problem of<br />

finding multiple genes responsible for disease,<br />

and upstairs is Thomas Jentsch, who is<br />

doing pioneering work on membrane proteins<br />

(see “The electrician’s toolbox”).<br />

Part of Nikolaus’ group is doing experimental<br />

work on cells and organisms. Another part is<br />

creating innovative computer algorithms to<br />

detect microRNAs. One way that they overlap<br />

is that a hypothesis about the identity and<br />

functions of microRNAs that has been generated<br />

in the computer can be tested in the lab.<br />

“But one doesn’t have the upper hand over<br />

the other,” Nikolaus says. “I’m just interested<br />

in problems, and in bringing the right methods<br />

to bear on them.”<br />

One project within the group is to establish a<br />

new, unusual model organism at the <strong>MDC</strong>: a<br />

small freshwater flatworm called the planarium.<br />

This animal has amazing regenerative<br />

powers that have fascinated scientists for over<br />

200 years. Cut into halves, or even smaller bits,<br />

the fragments can grow until they have reconstructed<br />

the entire body. Charles Darwin collected<br />

them during his five-year voyage on the<br />

Beagle, and Thomas Morgan wrote at least a<br />

dozen papers on planaria before devoting his<br />

laboratory to the fruit fly. Morgan discovered<br />

that it only took 1/279th of the worm’s body to<br />

rebuild an entire planarium.<br />

Over the past year Nikolaus’ interdisciplinary<br />

team has been settling into the newest building<br />

on campus, the Medical Genomics Center,<br />

a sleek black structure funded by the <strong>MDC</strong><br />

and FMP. When people work at night – which<br />

is not unusual – lights from the large office<br />

windows make it look like a futuristic vending<br />

machine.<br />

Science does not keep regular hours.<br />

��<br />

a very wide effort to discov-<br />

“Despite<br />

er and profile microRNAs, there’s<br />

no consensus about the number that might<br />

exist in the human genome,” Nikolaus says.<br />

“Estimates range from a few hundred to tens<br />

of thousands.”<br />

Theoretically these molecules can be captured<br />

in experiments, he says, but there are<br />

some problems. One is that cells may only<br />

produce a few copies of some microRNAs.<br />

Those may have important jobs in the cell, but<br />

they are easily overlooked when ten thousand<br />

copies have been made of another molecule.<br />

Computer scans of the genome might be<br />

another way to identify them, but even if the<br />

machine can be programmed to find their<br />

sequences, it won’t say what types of cells (if<br />

any) produce the molecules or what functions<br />

they might have. Finding that out requires<br />

experiments... which may not see microRNAs<br />

that are made in small amounts... “You get the<br />

point,” he says.<br />

New technology might offer a partial solution.<br />

Advances are so fast, he says, that even<br />

the most forward-thinking scientists find<br />

them hard to believe. “Technology is moving<br />

to the point that it will be possible to<br />

sequence a person’s entire DNA in an hour,”<br />

he says. “The first time that was done, it took<br />

15 years, and the job was just more or less<br />

completed five years ago. Obtaining individual<br />

genomes will not only help to answer a lot<br />

of fundamental questions about humans –<br />

for example, how much difference is there<br />

really between one person and the next –<br />

but it will also change a lot of aspects of<br />

medicine, from new diagnostics to new therapies.”<br />

The increasing speed and falling costs of<br />

sequencing makes it possible to solve some<br />

stubborn problems in new ways. “If you’re<br />

looking for something small and relatively<br />

265 Part four: From the protein village to the cosmopolitan city


are, one solution is simply to do a lot more<br />

sequencing,” Nikolaus says. It’s a bit like conducting<br />

a census to count the population of a<br />

city over and over again, in hopes of finding<br />

people who aren’t at home. If you do it<br />

enough times, eventually you’re likely to find<br />

almost everyone.”<br />

The method, called deep sequencing, is being<br />

applied to a variety of fascinating problems:<br />

to detect mutations in tumors, to sequence<br />

DNA from Neandertal homonids and the<br />

woolly mammoth, to study the genetics of<br />

social behavior in wasps, and to complete the<br />

genome of the grape Pinot Noir.<br />

Part four: From the protein village to the cosmopolitan city<br />

“There’s one catch to the power of deep<br />

sequencing,” Nikolaus says. “It’s heavily<br />

dependent on computing. In order to make<br />

sense of what you find, you have to be able to<br />

sort through immense amounts of sequence<br />

data and map them to specific locations in<br />

the genome.”<br />

The task is a bit like being given a fragment of<br />

text and having to find everywhere it appears<br />

in the Encyclopedia Britannica. That wouldn’t<br />

necessarily be a huge problem if you had the<br />

whole text of the encyclopedia on your computer,<br />

in a single file that could be searched.<br />

But the text of most genomes isn’t available<br />

in that form.<br />

266<br />

This has to do with the way most genomes<br />

have been sequenced and assembled. Instead<br />

of decoding human DNA letter by letter, from<br />

beginning to end, a “shotgun approach” was<br />

used. This would be like <strong>download</strong>ing the<br />

encyclopedia from the Internet onto your<br />

computer, in bits and pieces a few hundred<br />

letters long. If you did that (and you shouldn’t),<br />

it would be possible to reassemble the<br />

articles into a complete encyclopedia by<br />

matching up the beginnings and endings of<br />

texts and alphabetizing them. “Shotgun” DNA<br />

sequences can also be plugged into their correct<br />

positions in human chromosomes. But<br />

small entries might have been missed while


<strong>download</strong>ing the Encyclopedia Britannica,<br />

and one of those might contain the fragment<br />

you’re trying to find. That’s the situation for<br />

the human genome and other complex<br />

organisms that have been sequenced using<br />

the shotgun method. In fact, the only animal<br />

for which scientists have a full, beginning-toend<br />

DNA sequence is the worm C. elegans.<br />

So one of the computing challenges is to<br />

identify microRNAs by matching them to<br />

their proper places in the genome. Another is<br />

to deal with the “artifacts” that often arise in<br />

experiments. What may look like a microRNA<br />

may be a fragment of another molecule, an<br />

alternative version of part of a gene, or simply<br />

a mistake made while sequencing. If a computer<br />

program can’t tell the difference, it will<br />

find far too many false positives while looking<br />

at deep sequencing data.<br />

PhD student Marc Friedländer and other<br />

members of Nikolaus’ lab have found a way to<br />

address this problem in a program called<br />

miRDeep. “Basically we have taught the computer<br />

something about the biology of a<br />

microRNA,” Nikolaus says. “The cell processes<br />

premature microRNAs into active, mature<br />

micoRNAs. While doing so, it leaves a signature<br />

on which parts of the premature<br />

microRNA survive and which parts don’t.<br />

miRDeep detects these signatures in the deep<br />

sequencing data.”<br />

Each microRNA begins as part of a larger RNA<br />

molecule, which consists of a string of subunits<br />

called nucleotides. Some of the subunits<br />

are complementary to each other – they<br />

chemically attract each other. They are separated<br />

by a few nucleotides that don’t bind,<br />

instead making a hairpin-like fold. The shape<br />

of this fold is unique in microRNAs, which<br />

allows them to be recognized by a protein<br />

machine called the Dicer complex. It cuts the<br />

RNA into fragments. This leaves behind a<br />

microRNA, the loop at the top of the hairpin,<br />

and the short complementary sequence that<br />

the microRNA used to bind to. This last fragment<br />

is normally destroyed.<br />

Why can’t you find microRNAs through the<br />

chemistry of these folds? I ask. By scanning<br />

the DNA sequence for the right patterns?<br />

“False positives again,” Nikolaus says. “There<br />

are at least eleven million sequences in the<br />

human genome that would form hairpins if<br />

they were transcribed into RNA. Only a fraction<br />

of those are actually made into micro<br />

RNAs. So you can’t make predictions based on<br />

sequences alone. On the other hand, if you<br />

look at the small RNAs that have turned up in<br />

experiments, and find their positions in the<br />

genome, you can work backwards. You can ask<br />

if they belong to the right kind of hairpin.<br />

That’s an important indication that you’re<br />

dealing with a real active microRNA, and not<br />

just a random fragment of some other molecule.”<br />

As well as the microRNA, the other fragments<br />

left over by Dicer ought to appear in experiments.<br />

By comparison there ought to be more<br />

copies of the microRNA – after it is made, the<br />

cell protects it until it can be used.<br />

miRDeep uses all of these criteria and several<br />

others – for example, the fact that microRNAs<br />

are usually conserved through evolution, so<br />

the same sequences are found in many different<br />

species – to score the small RNAs found in<br />

deep sequencing experiments.<br />

When Marc and his colleagues finished the<br />

program, they put it through three tests. “We<br />

used deep sequencing data from the worm C.<br />

elegans because its genome is relatively small<br />

and the sequence is of high quality,” Nikolaus<br />

says. “MicroRNAs were discovered in that<br />

species and extensive studies have been carried<br />

out. We figured that if an analysis using<br />

miRDeep yielded the microRNAs that had<br />

already been found, it would validate the<br />

method. Then we carried out deep sequencing<br />

ourselves on white blood cells obtained<br />

from a dog, and a human cell line. We had different<br />

laboratories carry out the deep<br />

sequencing experiments and then ran them<br />

through miRDeep.”<br />

The tests had very positive results. The program<br />

revealed 89% of the microRNAs in C. elegans<br />

and found several new ones that were<br />

shown to be real by further experiments. It<br />

only proposed a handful of false positives.<br />

Marc and Nikolaus anxiously awaited the<br />

results from the other tests. They didn’t have<br />

anything to worry about; miRDeep proved<br />

itself in both.<br />

“The human genome is 30 times larger than<br />

that of C. elegans, but it has also been extensively<br />

studied in search of microRNAs,”<br />

Nikolaus says. “At the time of the study, 555<br />

microRNAs had been verified in the genome.<br />

No one kind of cell produces them all; the type<br />

we studied were known to produce 213. Well,<br />

miRDeep found 154 of them, nearly threequarters.<br />

If you work out what that means –<br />

in a single experiment with a single cell type,<br />

the program correctly identified 28 percent of<br />

all known human microRNAs. That’s pretty<br />

amazing.”<br />

The scientists were just as eager to see the<br />

results of the experiments on the dog cells.<br />

The genome of the domestic dog was completed<br />

in 2005, and it has become an increasingly<br />

important model for the study of<br />

human disease. Yet comparatively little work<br />

has been done on its genome, and only six<br />

microRNAs were known.<br />

The experiment with miRDeep has changed<br />

that picture dramatically. The program found<br />

three of the known microRNAs (four were present<br />

in the sample), and 203 more sequences<br />

passed the test. They will have to be verified in<br />

experiments, but so far miRDeep has done an<br />

excellent job of picking real microRNAs from<br />

other small molecules. Marc and Nikolaus are<br />

confident that nearly all of the sequences will<br />

turn out to be real. If so, they will confirm that<br />

the method works even for genomes whose<br />

microRNAs have not been studied. The dog is<br />

an important test case, Nikolaus says, because<br />

researchers hope to use it to understand<br />

microRNAs’ contributions to disease in an animal<br />

that is closely related to humans.<br />

267 Part four: From the protein village to the cosmopolitan city


He couldn’t resist trying the method out on<br />

planaria as well. “It’s an extreme case because<br />

of its very distant evolutionary relationship to<br />

humans and other animals with well-studied<br />

microRNAs. If you assume that the program<br />

has some built-in biases because it is based<br />

on familiar genomes, then data from planaria<br />

ought to point that out.”<br />

Quite a bit of experimental data was available<br />

on planaria microRNAs – 61 were known.<br />

miRDeep found 86 percent of them and proposed<br />

39 new ones. Of those, Nikolaus and his<br />

colleagues have tested 19 to see if they really<br />

function as microRNAs. So far 16 of them have<br />

turned out to be real.<br />

��<br />

None of these molecules could have been<br />

found without the marriage of laboratory<br />

and computational science that Nikolaus<br />

has drawn together in the Center for Medical<br />

Genomics.<br />

“A program can predict as many microRNAs as<br />

you want, but those predictions only have any<br />

value if they turn out to be real,” Nikolaus<br />

Part four: From the protein village to the cosmopolitan city<br />

says. “And search algorithms have to be finetuned<br />

by the biology. When the two<br />

approaches work together, we can learn a lot<br />

about microRNAs from deep sequencing<br />

experiments, even if the subject is a relatively<br />

unexplored genome like that of planaria.”<br />

Computer programs have become adept at<br />

looking for the genes in genomes; they are<br />

large sequences with complex patterns. But<br />

the lack of equally high-quality methods to<br />

identify microRNAs and their target genes<br />

has made it hard for researchers to get a grip<br />

on the problem. That’s something Nikolaus<br />

would urgently like to rectify, since experiments<br />

suggest that microRNAs play an<br />

important role in all kinds of biological<br />

processes – likely in the development of many<br />

diseases.<br />

Recently Nikolaus helped take a global look at<br />

the effect of microRNAs in a particular developmental<br />

process: the maturation of B cells,<br />

the white blood cells that produce antibodies.<br />

(They are a central theme in the story “Where<br />

slow rivers meet”) Making a B cell’s antibody<br />

is a complex process that involves cutting the<br />

268<br />

left: A fully-grown planarium<br />

right: a worm in the process of regrowing its head<br />

far right: The smaller worm has just regenerated its tail<br />

DNA strand, discarding long regions, and<br />

reassembling the sequence to form a new<br />

antibody gene. There are a huge number of<br />

pieces to choose from. Because different<br />

regions are removed and combined in each<br />

cell, it acquires a unique gene and builds a<br />

unique antibody. Parts of this process are not<br />

yet understood, and experiments suggested<br />

they might involve microRNAs. A group at<br />

Harvard decided to find out. They needed<br />

computational expertise in microRNAs, and<br />

Nikolaus was the logical choice. Besides being<br />

one of the world leaders in the field, he had a<br />

connection to the lab – it was run by his<br />

father.<br />

Klaus Rajewsky is one of Germany’s bestknown<br />

scientists. Before relocating to the<br />

Harvard Medical School in 2001, his lab carried<br />

out a great deal of pioneering work on the<br />

biology of B cells at the University of Cologne.<br />

Along the way the group created one of the<br />

most powerful tools in genetic research: the<br />

conditional knockout. Until the method was<br />

developed, researchers could not knock out<br />

many of the cell’s most crucial genes to study<br />

their functions, because removing them


caused embryos to die at a very early stage –<br />

often before they had even undergone the<br />

first cell division. Conditional knockouts<br />

allowed a gene to be shut down in a single<br />

type of cell at a precise time. It is now a standard<br />

tool in labs across the world and is the<br />

basis of several of the projects described in<br />

this book.<br />

Did microRNAs play an important role in the<br />

maturation of B cells and the creation of antibodies?<br />

Sergie Koralov and Stefan Muljo, two<br />

members of Klaus’ group, thought of a way to<br />

find out. They used a conditional knockout to<br />

delete the Dicer gene from immature B cells<br />

in mice. Since this is the key enzyme in the<br />

Dicer complex, its removal meant that RNA<br />

hairpins would not be cut to create<br />

microRNAs.<br />

The experiment had a dramatic impact on<br />

the mice. They failed to develop any mature B<br />

cells; the cells died in an early stage of development.<br />

Klaus and his colleagues suspected<br />

that this was due to the presence of molecules<br />

that would normally be blocked by<br />

microRNAs. They compared activity in the<br />

immature B cells to those of normal mice.<br />

Nikolaus and his PhD student Azra Krek analyzed<br />

the results.<br />

“We found 411 different molecules that were<br />

being produced at significantly higher levels<br />

in the experimental animals,” Nikolaus says.<br />

“The levels of some of them – maybe most of<br />

them – had probably changed because they<br />

were no longer being held in check by<br />

microRNAs. The only way to tell was to check<br />

the molecules to see if they had the features<br />

of microRNA targets, and dozens of them did.<br />

And the computation predicted that a single<br />

microRNA cluster had the largest effect in<br />

the Dicer knockout experiment. Subsequent<br />

analysis of potential targets of this cluster<br />

identified bim as a very likely target that<br />

should be specificially important for B cell<br />

maturation regulated by microRNAs.”<br />

Another lab had shown that a protein called<br />

bim is a key player in the survival of B cells.<br />

Cells that produce it die. They are supposed<br />

to – this keeps the body from making cells<br />

that will attack itself and cause autoimmune<br />

diseases. But not all B cells die; they are<br />

allowed to survive when they block the production<br />

of bim.<br />

“This is an essential step in making B cells<br />

which can participate in healthy immune<br />

responses,” Nikolaus says, “and the study<br />

strongly suggests that it’s partly the work of<br />

microRNAs.”<br />

It’s not the only role that the tiny molecules<br />

play in the immune system. A paper last year<br />

by Klaus’ lab – also with the help of Nikolaus<br />

– showed that microRNAs are equally important<br />

in the development of another type of<br />

white blood cell – T cells.<br />

��<br />

The Medical Genomics Center, home to<br />

Nikolaus’ group, has spacious rooms at<br />

the corners of the floors. The groups have put<br />

in comfortable sofas, coffee machines, and<br />

old copies of journals – not all about science.<br />

There’s a motorcycle magazine on a table. It’s<br />

a place to relax, talk, and think about other<br />

things. As Nikolaus lounges on a sofa his fingers<br />

drum a light pattern on the armrest.<br />

Maybe it’s Chopin, some difficult passage<br />

269 Part four: From the protein village to the cosmopolitan city


that he once rehearsed for hours and hours,<br />

until he had it in his fingers – no brain<br />

required.<br />

We have been talking about one of his main<br />

interests, a new style of science called systems<br />

biology. The expression is a fad, so widely used<br />

these days that it seems to mean everything<br />

and nothing. Nikolaus has a firm idea of what<br />

it means; it has been guiding the way he has<br />

put his group together. He says that his ideas<br />

and approach have been heavily shaped by his<br />

experiences in the United States, where he<br />

worked for nine years after obtaining his PhD<br />

from the University of Cologne.<br />

But a systems approach to science requires<br />

more than one group, and he has a bigger<br />

vision of things; while getting settled in, he<br />

proposed unifying the work of laboratories<br />

and medical specialists throughout Berlin<br />

into an institute that will be devoted to systems<br />

biology.<br />

Today he’s excited because he just received<br />

word that the vision will soon take on very<br />

concrete form. The German Ministry of<br />

Education and Research (BMBF) had<br />

announced a major new program to fund<br />

projects in research and innovation in the<br />

new German states. It was an opportunity not<br />

to be missed. Nikolaus and the directors of the<br />

institute put together a joint proposal on<br />

behalf of the <strong>MDC</strong> and the Charité and submitted<br />

it. Now he has been told that the<br />

BMBF has pledged 7.5 million Euros of support<br />

over the next three years. The idea is no longer<br />

just virtual; it will have a real roof and walls.<br />

“Over the last 50 years, molecular biology has<br />

been breaking things down to their most fundamental<br />

level, learning what the units are<br />

and how to control them,” he says. “Now we’re<br />

starting to make a serious attempt to put<br />

things back together. That doesn’t only mean<br />

figuring out how proteins work together in<br />

networks or machines. That’s part of it – to<br />

see the complexity of interactions at the scale<br />

of molecules. But molecules are embedded in<br />

Part four: From the protein village to the cosmopolitan city<br />

structures of larger scales – cells, tissues,<br />

organs, animals, even the environment.<br />

Ultimately all of those levels work together to<br />

read the information in a genome and turn it<br />

into an organism, with a lot of input from the<br />

environment. They also work together to<br />

determine whether you’re healthy or sick,<br />

whether over the long term you’ll develop<br />

Alzheimer’s – or whether you can learn to play<br />

the piano.<br />

Until very recently, Nikolaus says, technology<br />

has generally restricted a biologist’s view to a<br />

particular slice of things. “We have methods<br />

to look at the structures of molecules, which<br />

is the smallest scale; then you jump up to the<br />

level of biochemistry, which tells you whether<br />

proteins are interacting with each other.<br />

Another jump gives you the light microscope<br />

view of the world, which is good at watching<br />

major cell structures. Other techniques are<br />

used to examine larger structures such as tissues<br />

and organs. Eventually you reach the<br />

domain of the doctor and the surgeon.<br />

“Most scientists now have access to highthroughput<br />

tools that give them a very broad<br />

overview of what’s happening at a particular<br />

270<br />

level, like the DNA microarrays we use in deep<br />

sequencing,” he says. “It’s giving them the<br />

feeling that they’re looking at whole organisms.<br />

But obtaining a snapshot of the entire<br />

set of RNAs in a cell isn’t in itself systems biology,<br />

unless that information can be woven<br />

into higher levels. And it only gives you a small<br />

piece of the picture of what is going on at the<br />

molecular level.<br />

“One of our long-term goals is to be able to<br />

zoom in and out at different scales and to<br />

make realistic predictions like, ‘if we change<br />

one letter of the genetic code, what’s likely to<br />

happen at all of those larger scales?’ Or, ‘if a<br />

person takes this drug, how will that find its<br />

way into cells, change biochemical processes,<br />

and then work its way back up to influence his<br />

overall health?’ Or, ‘what does this disease in a<br />

mouse tell me about something similar that<br />

happens to humans?’”<br />

The marriage between laboratory and computational<br />

science has already paid off in the<br />

projects on microRNAs: turning the most<br />

basic level of information in the genome –<br />

tiny bits of DNA sequence – into crucial<br />

insights about immunity and other systems<br />

in the body.<br />

Another recent project, Nikolaus says, makes a<br />

good example of how everything comes<br />

together. “We’ve known that microRNAs act in<br />

two ways to block the synthesis of proteins,”<br />

he says. “One method is by triggering the<br />

destruction of their target RNAs. In the<br />

other way, the RNA stays around, but it’s prevented<br />

from being used to make proteins.<br />

How do you tell which RNAs are being handled<br />

in each way? You can’t do it molecule-bymolecule,<br />

because you don’t know what the<br />

targets are.<br />

He teamed up with Matthias Selbach, who<br />

can quantify protein levels of thousands of<br />

genes using the latest advances in mass spectrometry.<br />

In addition, by changing the procedure<br />

used to make measurements, they could<br />

for the first time quantify how many proteins


Catherine Adamidi


were synthesized after activating or repressing<br />

microRNA in human cells. Three months<br />

of pure measurement time and a year of computational<br />

and experimental follow-up analyses<br />

were carried out by PhD students Bjoern<br />

Schwanhaeusser from Matthias’ lab and<br />

Nadine Thierfelder from Nikolaus’ group, with<br />

Minnie Fang, also in Nikolaus’ lab. They arrived<br />

at a detailed picture how cells react, on the<br />

proteome, transcriptome, and sequence level,<br />

to microRNAs.<br />

The result, Nikolaus says, is the first ever picture<br />

of how proteins in cells respond – in<br />

totality – to changes in levels of an miRNA.<br />

“We saw that raising levels of a particular<br />

miRNA has an effect on the protein production<br />

of thousands of genes,” he says. “That<br />

was interesting in itself, but the way we had<br />

set up the study gave us answers to some<br />

other questions that have been very hard to<br />

answer. For example, we knew that most<br />

miRNAs don’t work like an on-off switch, com-<br />

Part four: From the protein village to the cosmopolitan city<br />

pletely shutting down a protein. Are they<br />

more like a volume control, adjusting how<br />

much gets produced? And we knew that<br />

miRNAs worked in these two ways – by causing<br />

the breakdown of their targets, or blocking<br />

translation. But it was completely unclear<br />

how widely this second mechanism was used<br />

by cells, and how important it was in volume<br />

control.”<br />

I’m suddenly reminded of the piano again.<br />

Once you’ve played a note you can’t take it<br />

back, but you can dampen it or change its<br />

sound by pressing one of the three pedals.<br />

Translational control was like a new pedal;<br />

Nikolaus wanted to understand how it<br />

worked and when it was used as the cell<br />

played its daily recitals.<br />

Answering his questions required a way to<br />

monitor the behavior of the cell’s entire set of<br />

RNAs and proteins as the activity of an miRNA<br />

changed. “We raised the production of an<br />

miRNA and studied its effects. The method<br />

272<br />

allowed us to watch the cell’s response, in<br />

terms of which proteins were being produced,<br />

and at what amounts. Then we did exactly the<br />

opposite – we lowered the levels of the<br />

miRNA. It was like using a volume control; the<br />

amount of the miRNA determined the<br />

amount of proteins. We saw direct effects on<br />

hundreds of genes, and indirect effects on<br />

thousands.”<br />

This gave Nikolaus and his colleagues the first<br />

global look ever at the importance of miRNA<br />

translation control in the cell. “Hundreds of<br />

target RNAs were blocked at the step of translation,”<br />

he says. “But the effects were relatively<br />

week. Usually they cause a less than two-fold<br />

drop in the protein output. So it’s a volume<br />

control, but a dial that is stuck in the middle –<br />

you can’t turn it all the way up or down.”<br />

Is there anything special about the RNAs that<br />

are controlled by blocking translation, as<br />

opposed to those that are simply destroyed<br />

when an miRNA docks onto it?


“The first type seems to happen most in the<br />

cases of molecules that are to be secreted, or<br />

inserted into cell membranes,” he says. He<br />

may even know why. But he’s not ready to<br />

explain, not today. “That’s for your next book.”<br />

��<br />

One of Nikolaus’ colleagues comes in and<br />

he introduces us. It’s Catherine Adamidi,<br />

a technician from France. “She’s in charge of<br />

the planaria,” he says. “Want to see some<br />

worms?”<br />

“Oh, that’s good, I was just about to feed<br />

them,” she says.<br />

Catherine leads the way to the lab. The planaria<br />

are slinking along the bottom of white<br />

plastic boxes, half-filled with water. I peer over<br />

the edge as she gives them fresh water.<br />

They’re easy to care for, she says, but the containers<br />

have to be kept clean.<br />

“You can find them almost anywhere there is<br />

fresh water, in ponds and in lakes,” she says.<br />

“These come from a fountain in Barcelona. I<br />

went wading.”<br />

The largest of the worms are about the length<br />

of my fingernail. They are thin and very flat,<br />

with triangular heads and two eyespots that<br />

make them look cross-eyed and slightly<br />

ridiculous, like cartoons. The spots, she says,<br />

are some of the most primitive eyes found in<br />

nature: two cups lined with cells called photoreceptors,<br />

which transform light into electrical<br />

stimuli. A planarium usually swims<br />

away from the light. They also don’t like to be<br />

disturbed very much. When she jostles the<br />

container, they curl up into tiny balls.<br />

Some species of the flatworm reproduce<br />

through sex, but most types create offspring<br />

without the help of a mate. From time to<br />

time, particularly when the worms are under<br />

some sort of stress, they spontaneously split<br />

into pieces. Each part grows into an entire<br />

new flatworm.<br />

“This ability to regenerate seems to come<br />

from the fact that they have one type of stem<br />

cell, and it’s able to do everything,” Catherine<br />

says. “Adult humans, you know, have many different<br />

types that are limited in what they can<br />

become. If you had only one type, too, it would<br />

need to be flexible enough to regrow any part<br />

of your body, the way an embryonic stem cell<br />

does.”<br />

It’s no wonder that planaria are a favorite of<br />

groups studying stem cells and regeneration.<br />

Maybe human cells could be taught to behave<br />

the same way, to regrow new organs when<br />

they fell prey to injuries, old age, or disease.<br />

This isn’t the only odd aspect of the flatworms’<br />

lives. “If you starve them, they don’t<br />

die,” Catherine says. “They shrink. And they<br />

regress – they repeat their process of development,<br />

only backwards. When you feed them<br />

again they develop and grow.”<br />

And then there’s memory. Planaria are simple<br />

animals, but for their body size they have relatively<br />

large brains and synapses. This means<br />

they can be trained to do things like run a<br />

maze, using light and mild electroshocks.<br />

What happens if the worm is trained and<br />

then cut in half? You’d expect the head half to<br />

remember the maze, but interestingly, when<br />

the tail grows a new head of its own, that<br />

worm seems to remember the maze, too.<br />

There isn’t much in the literature about this;<br />

researchers have shied away from the topic<br />

since the 1960s. Back then a researcher<br />

named James McConnell tried to prove that<br />

planaria could learn through cannibalism, by<br />

eating the ground-up remains of other flatworms<br />

that had learned the way through his<br />

mazes. It turned out to be a case of “what you<br />

want to see is what you get;” when the experiments<br />

were repeated with proper controls,<br />

the results could never be reproduced.<br />

It’s comforting to find out that cannibalism<br />

won’t make the worms any smarter.<br />

Otherwise Catherine would probably come in<br />

one day to find them all gone – except for one<br />

very fat, very smart planarium in the middle<br />

of the tank, plotting its escape. But I wonder<br />

about memories stored in tails, in cells. The<br />

planarium’s strange method of reproduction<br />

means that all of the worms are clones,<br />

regrown from the cells of some original<br />

ancestor that was born a long time ago, not<br />

too long after the origins of the first animals.<br />

So they never really die. It’s strange to be looking<br />

at an animal which, for all practical purposes,<br />

is immortal. Even stranger to think that<br />

it still might have tiny splinters of memories<br />

of some of those former lives.<br />

273 Part four: From the protein village to the cosmopolitan city


A rat in a tree and a<br />

wolf on the loose<br />

It was a bad year to be a British citizen traveling in the United States.<br />

The colonies were waging war against England, and spies were as<br />

thick as mosquitoes. In 1778 John Berkenhout, an English physician,<br />

arrived in the state of Pennsylvania pretending to be sympathetic to<br />

American independence. But by the fall he had been exposed as a spy<br />

and was languishing in a Philadelphia prison.<br />

Berkenhout was eventually released and lived to stake another claim to<br />

fame. Like many physicians of his day, he was an avid naturalist, devoting<br />

his free time to the search for new species of insects and plants.<br />

Europe – and particularly Britain – was caught up in a fever of collecting<br />

and classifying. This had started with the work of the great (and<br />

slightly crazy) Swede Carl Linneaus, who invented a magnificent system<br />

that catalogued the world of plants according to the structure of their<br />

sexual organs. It turned out to be a brilliant insight, compared to other<br />

attempts to classify species based on color or the shapes of their leaves.<br />

(Linneaus had the advantage of knowing that plants reproduced sexually,<br />

a fact which had only been discovered at the end of the 17th century.)<br />

Linneaus’ scientific work was based on precise observations, but his


Norbert Hübner


prose was full of poetic rhapsodies with heavy sexual overtones. The<br />

following passage, for example, comes from the opening of his dissertation<br />

submitted to the University of Uppsala in 1730:<br />

Words cannot express the joy that the sun brings to all living<br />

things. Now the blackcock and the capercailzie begin to<br />

frolic, the fish to sport. Every animal feels the sexual urge.<br />

Yes, Love comes even to the plants. Males and females, even<br />

the hermaphrodites, hold their nuptials (which is the subject<br />

that I now propose to discuss), showing by their sexual organs<br />

which are males, which are females, which hermaphrodites...<br />

The actual petals of a flower contribute nothing to generation,<br />

serving only as the bridal bed which the great Creator<br />

has so gloriously prepared, adorned with such precious bedcurtains,<br />

and perfumed with sweet scents in order that the<br />

bridegroom and bride may therein celebrate their nuptials<br />

with the greater solemnity. When the bed has thus been made<br />

ready, then is the time for the bridgegroom to embrace his<br />

beloved bride and surrender himself to her...<br />

Ah, if only dissertations could be written that way today.<br />

The tradition of mixing botany and sex was enthusiastically continued<br />

by the Englishman Joseph Banks, who sailed with James Cooke to the<br />

South Pacific as a naturalist. Bank’s amorous adventures with Tahitian<br />

women became the talk of high-society London. (Less known was the<br />

fact that the islanders were willing to trade sex for bits of metal, such<br />

as nails, which led Cooke’s crew to remove so many nails from the ship<br />

that it started to fall apart.) In any case, Bank returned to England and<br />

began preaching the classification system of Linneaus. This influenced<br />

generations of young English scientists, from John Berkenhout to<br />

Charles Darwin.<br />

One of the questions that interested naturalists was how species<br />

spread from place to place. Berkenhout’s book Outlines of the Natural<br />

History of Great Britain, published in 1769, attempted to provide an<br />

answer for a species of rat that infested Britain. Believing that the animal<br />

had been carried to the island 50 years earlier on Norwegian<br />

ships, he named it the Brown Norway rat. As it turns out, he was<br />

wrong; modern studies have showed that the animal probably originated<br />

in central Asia and arrived in England much earlier. But the<br />

name has stuck.<br />

Part four: From the protein village to the cosmopolitan city<br />

In the 20th century the Brown Norway and other strains of rats were<br />

brought into the laboratory, where they have become some of the<br />

most important modern organisms for medical research. Their value,<br />

says Norbert Hübner of the <strong>MDC</strong>, lies partly in the fact that they are<br />

more closely related to humans than the mouse is, and are larger and<br />

easier to dissect. “There have been exquisite studies of the animal’s<br />

anatomy and physiology,” he says. “Structures that are tiny and difficult<br />

to examine in the mouse are much easier to investigate in rats.<br />

That’s especially important if you’re trying to determine how genes<br />

affect development or cause disease.”<br />

276<br />

Henricke Maatz, Judith Fischer<br />

and Norbert Hübner


��<br />

Heavy inbreeding in the laboratory has led to the development of<br />

hundreds of strains of rats, many of which have problems that<br />

closely resemble human diseases. Often the symptoms appear to be<br />

the work of multiple genes. In nature, disease-causing genes brought<br />

together in individuals usually scatter again as animals find diverse<br />

mates. Dangerous combinations don’t entirely disappear, but they<br />

remain rare. That changes if there is heavy inbreeding in a population.<br />

Both parents are much more likely to carry disease-causing combinations<br />

of genes and pass them along to their offspring.<br />

All of this has made the rat useful in traditional medical research and<br />

pharmacology, Norbert says. “It’s a great model to study how body systems<br />

respond to diseases and drugs. Now we need to bring it into the<br />

modern age. That means developing the knowledge and tools to<br />

manipulate its genes as we do with the mouse.”<br />

If that can be done, he says, it should give scientists insights into a<br />

much more fundamental biological question. “There is an almost oneto-one<br />

match between the genes of humans and rats, but we’re still<br />

very different,” he says. “Small differences between genes must be<br />

responsible for that. But which genes cause which differences? That’s<br />

277 Part four: From the protein village to the cosmopolitan city


the question you ask if you want to compare two species. It’s also the<br />

issue if you’re comparing two individuals. One has high blood pressure;<br />

one does not. One responds to a drug; the other does not.”<br />

Most biological processes are collaborations between many genes, so<br />

most health problems are also likely to be due to combinations of particular<br />

varieties of molecules. The best strategy to discover them,<br />

Norbert says, is to find an animal for which many strains exist and for<br />

which there are already good models of complex human diseases –<br />

the rat is a perfect example. The next step is to carry out linkage studies<br />

in the animals, to identify genes or blocks of DNA that are inherited<br />

together in animals that have similar problems, and then to see if<br />

those patterns appear in people suffering from the human form of the<br />

disease.<br />

Before that could be done with the rat, however, scientists needed to<br />

get a look at the complete rat genome. This was accomplished in 2004<br />

by an international consortium of researchers who finished the<br />

genome sequence and analyzed it. Norbert’s lab participated. As other<br />

labs collected sequences, the <strong>MDC</strong> group helped plug them into their<br />

proper positions on the rat chromosomes. The process was like assem-<br />

Part four: From the protein village to the cosmopolitan city<br />

bling a puzzle with million pieces, each showing a tiny fragment of a<br />

map of Europe. The end result was a complete, gapless genome<br />

sequence.<br />

With this in hand, Norbert and his colleagues were ready to begin<br />

measuring diversity – in other words, to see how individuals and<br />

strains differ from each other and the “norm” of the genome. Norbert<br />

says there are two main kinds of variation.<br />

“The first type is a simple ‘spelling’ difference between the same<br />

sequences in two animals,” he says. “Often this involves just one letter<br />

– two regions of the code are identical except for a single base pair. The<br />

technical term for this is a single nucleotide polymorphism, or SNP.<br />

Within a species, there are normally several possible spellings for each<br />

gene – it usually contains a number of SNPs. Often this has nothing to<br />

do with disease – it’s just the normal variation that makes each of us<br />

unique.”<br />

Other differences are more significant. The story “Waking a Sleeping<br />

Beauty and other tales of ancient genes” tells how blocks of DNA<br />

sometimes jump to new positions in the genome. The blocks may con-<br />

278


Henricke Maatz, Judith Fischer and Katharina Grunz<br />

tain just a few letters, a whole gene that is thousands of bases long, or<br />

even longer strings of DNA. The positions of these blocks provide<br />

another way of comparing and classifying rats. Animals with the<br />

blocks arranged in the same way are said to belong to the same haplotype.<br />

Finding out what we want to know about the rat genome, Norbert<br />

says, means studying enough animals to discover the main haplotypes<br />

and listing the SNPs found at various positions in their genetic codes.<br />

The result will be a family tree or another type of diagram that shows<br />

how all the strains are related. That would be easier if scientists knew<br />

the origins of the rats first brought into the lab. But that is not the<br />

case, and nor do lab records describe how the early animals were interbred<br />

and inbred. Some of these questions can be cleared up by studying<br />

SNPs and haplotypes<br />

Wouldn’t you need a complete genome sequence from each individual<br />

animal? I ask.<br />

“That would be ideal,” he says. “In the future, as it gets easier and less<br />

expensive to sequence DNA, it will be feasible. For people as well as<br />

rats. Someday you’ll probably carry around a copy of your genome on<br />

something like a credit card or memory stick and give it to your doctor<br />

when you go in for a checkup. It will be helpful for diagnoses and other<br />

aspects of your medical care.”<br />

��<br />

For the moment, collecting and analyzing sequences at a somewhat<br />

lower “resolution” has given the lab enough to do. The rat<br />

genome gave the group and their colleagues across the globe an outline<br />

to begin cataloguing SNPs and haplotypes. Kathrin Saar, a ...<br />

in Norbert’s group, headed the project – which involved coordinating<br />

the work of 12 laboratories in Germany, the United Kingdom,<br />

France, Japan, Spain, and the Czech Republic. Their findings will<br />

go to press at about the time this book is printed.<br />

The study revealed that the strains can be clustered into ten clades –<br />

or ten main types. Nine of these are closely related, which means that<br />

their ancestors were similar to each other – or they arose from a common<br />

pool of animals. The Brown Norway rat turns out to be an exception,<br />

a more distant cousin. It’s good to have such an “outlier”, Norbert<br />

279 Part four: From the protein village to the cosmopolitan city


280<br />

says, when you’re trying to understand the causes and effects<br />

of diversity. “Just as if you wanted to get a feeling for human<br />

variety, you might start by looking at people from one region,<br />

but you would certainly want to include a few from far away.”<br />

Originally Norbert hoped that the study would permit a<br />

detailed reconstruction of the rats’ family tree, but that hasn’t<br />

been possible. “This probably means that the first rats brought<br />

into the lab were quite diverse, already coming from mixed<br />

strains, and that they were interbred in the early years, mixing<br />

things up even more.”<br />

The project was able to plot three million SNPs, obtained from over<br />

150 strains of rats, onto the genome. “That works out to about one<br />

variation per every 800 base pairs,” Norbert says. “We were also<br />

able to make predictions about the effects that roughly a tenth of<br />

these variations probably have on cells and tissues. In 56 cases we<br />

predict that the change has an effect on the function of a protein.”<br />

Seven of the variations involve rat molecules that are closely<br />

related to genes involved in hereditary diseases or cancer in<br />

humans. One is ALDH2, which has been linked to acute alcohol<br />

intolerance. Another is AFF4, which causes cancer if it doesn’t<br />

function properly. Several hundred of the variations probably<br />

affect the way cells cut and paste RNAs to allow different forms<br />

of proteins to be made from single genes. Others affect regions of<br />

DNA that control when and how genes are activated.<br />

The study has permitted the lab to identify parts of the genome<br />

that change faster than others. Every genome has such areas,<br />

and sometimes they reveal features that are being rapidly<br />

changed by evolution. Here they discovered an interesting parallel<br />

to the mouse: a great deal of variety was found in genes<br />

involved in the sense of smell. Scent is so central to a rodent’s<br />

life that even small changes in the processes that govern it can<br />

have a big influence on the animal’s ability to survive and pass<br />

along its genes.<br />

The haplotype map remains incomplete, but the data gathered<br />

so far confirms something that Norbert has suspected for a<br />

long time: a rat is not always a rat. “While the substrains can be<br />

clustered into ten groups with similar genomes, they aren’t<br />

nearly identical,” he says. “There’s still quite a bit of diversity. It’s<br />

not that obvious if you look at the genome as a whole, but if<br />

you look at regions that are changing the fastest, there is a lot<br />

of variation between the substrains. In a rat known as the Long-<br />

Evans strain, 29 percent of the base pairs in these areas have<br />

different spellings than you find in other strains.<br />

“This happens because until now, it hasn’t been possible for<br />

labs to cleanly identify what type of rat they started out with,”


Norbert says. “Then when they breed the animals with other rats –<br />

whose origins also aren’t clear – the offspring inherit genomes that<br />

are unique to each place. It means you have to be careful when you try<br />

to extend what you’ve found in one animal to others. A genetic problem<br />

may be unique to one strain because it has a particular genetic<br />

makeup. You may not be able to generalize to another strain of rat – let<br />

alone a human being. These projects should help labs ‘clean up’ their<br />

strains. At least now we have some standard reference points that can<br />

help you define the animal you’re working with.”<br />

Norbert quickly points out that all of this work is descriptive.<br />

“Successfully linking a SNP or haplotype to a disease is only the starting<br />

point,” he says. “The next step is to determine why one form of a<br />

gene – or a set of genes – lead to disease. That requires like conditional<br />

knock-outs that remove the gene in one tissue at a specific time.<br />

While those tools are well-developed in the mouse, most of them<br />

don’t yet work in the rat. There’s a lot of catching up to do.”<br />

��<br />

Norbert appeared in “The case of the short-fingered muskateer”<br />

for his role in helping Friedrich Luft’s group look for the genetic<br />

causes of brachydactyly and hypertension. That story shows how difficult<br />

it can be to pin down the causes of a disease, even when the problem<br />

seems to lie in a single region of DNA. With multiple genes, the<br />

task becomes staggering. I wonder what has motivated Norbert to<br />

tackle one of the hardest questions in today’s science.<br />

“I don’t really know,” he says. “I like complicated things. But my being<br />

here right now, doing this, is a bit of an accident.”<br />

In 1989 he was a medical student at the University of Heidelberg. He<br />

was looking for a part-time job in a laboratory and was hired on as a<br />

technical assistant by Detlev Ganten, who would soon move to Berlin<br />

to fully dedicate himself to the creation of the <strong>MDC</strong>. At about the time<br />

Norbert finished his first degree, another member of the group named<br />

Klaus Lindpaintner was getting ready to move to Harvard in the USA<br />

to set up his own group. He wanted to hire good people and invited<br />

Norbert to go along. So that’s what he did.<br />

After two years of work on his doctoral thesis at Harvard, Norbert<br />

moved back to Germany, this time to Berlin, to finish his medical studies<br />

at the Charité. That made him one of the few people to have<br />

earned both an MD and PhD. It also made him just the right type of<br />

person to find a job at the <strong>MDC</strong> – once again, he laughs, in the group<br />

of Detlev Ganten. This time as a postdoc.<br />

In 2003 he was given his own group, followed a year later by his<br />

appointment as director of the <strong>MDC</strong>’s “Gene Mapping Center.” Since<br />

2005 he has held a Professorship in Medical Genomics and Genetics at<br />

the Charité.<br />

One red thread that ties together all these phases of his life has been<br />

the rat.<br />

“Some complex processes are hard or impossible to explore in the<br />

other models we have,” Norbert says. “For example, mice don’t<br />

develop arteriosclerosis. We don’t have mouse models of natural types<br />

of hypertension or heart insufficiency. It may be that the tiny anatomy<br />

just makes some of these things hard to find – on the other hand,<br />

something in mouse biology may protect them. Another thing is<br />

that when the mouse was brought into the lab, it was mainly used<br />

to watch development. The rat, on the other hand, has been t<br />

he favorite in looking at human disease. These two ways of looking<br />

at an animal are deeply connected, and we need to bring them to -<br />

gether.”<br />

Norbert admits that he and his colleagues are just opening the book<br />

on diversity in rats and their link to human disease. But the first pages<br />

of that book have already shown how an intimate knowledge of rat<br />

strains and their defects can be a powerful tool in establishing such<br />

connections.<br />

“One of the problems in human disease is that it’s not only genetic<br />

factors which are intertwined – so are the symptoms and various body<br />

systems,” he says. “Think about one of the most common causes of<br />

death in humans: heart failure. That’s somehow connected to other<br />

things, like high blood pressure and hypertrophy (enlargement of the<br />

heart). High blood pressure is one of the major risk factors for heart<br />

failure. But is it the cause? Are both things caused by the same genes?<br />

Or are they separate problems, influenced by different sets of genes?<br />

In some patients they all seem to be connected, one symptom following<br />

on the heels of another. But other people suffer from hypertension<br />

and hypertrophy without suffering any of the clinical signs of<br />

heart failure. So we may actually be talking about different diseases.”<br />

One way to answer the question, he says, is to look at available rat<br />

models of disease and try to peel the problems apart by studying different<br />

strains. That’s what Jan Monti and Judith Fischer, two members<br />

of his group, have been up to with colleagues from the <strong>MDC</strong> and<br />

Charité, from the pharmaceutical industry, and from other institutes<br />

in the UK and the USA.<br />

They began with a strain known as the spontaneously hypertensive<br />

heart failure (SHHF) rat, which develops a disease that mirrors most of<br />

what is known about human heart disease linked to high blood pressure.<br />

“The animals not only develop heart failure after high blood pressure<br />

and hypertrophy of the heart,” Jan says, “but when you look at the<br />

activity of their genes, you discover patterns that strongly echo what<br />

happens in human tissue.” Another strain, the stroke-prone spontaneously<br />

hypertensive (SHRSP) rat, has similar problems of high blood<br />

pressure but does not suffer from heart failure.<br />

281 Part four: From the protein village to the cosmopolitan city


Patterns of heredity showed that the symptoms were based on genes.<br />

“That meant that if you mated the two types of rats, you would get different<br />

types of offspring,” Jan says. “Some would have all of the problems<br />

of both parents, and others would have various combinations.<br />

We thought we could use this to peel apart the genes and their symptoms<br />

and try to figure out their connections.”<br />

At the same time the lab crossed the SHHF (heart failure) rats with a<br />

control strain that had normal blood pressure. This would lead to different<br />

sets of genes and symptoms in the offspring – some of which<br />

would still suffer heart failure. If the strategy worked, Jan and Judith<br />

would have a variety of different animals with different sets of genes.<br />

It ought to be possible to find the molecule or molecules responsible<br />

for heart failure, and to see whether the disease only happened in animals<br />

that also had high blood pressure.<br />

The laws of genetics and the behavior of dominant and negative<br />

genes meant that the researchers had to wait two generations to get<br />

the right animals. They sorted the offspring into groups based on<br />

blood pressure and other factors, closely watched their health, and<br />

began looking for the common denominator between heart failure<br />

and DNA.<br />

While the rats’ health problems were made worse by high blood pressure<br />

and the enlargement of the heart, some mice experienced heart<br />

failure without the other symptoms. This hinted that a separate gene<br />

was at work. The scientists pinpointed the heart failure defect to a<br />

region of the rat chromosome 15, and eventually to a gene called<br />

Ephx2. The laboratory sequenced the gene in the animals and found<br />

several variations – SNPs – that would change the way cells made proteins<br />

from the Ephx2 gene. That in itself was promising. But things<br />

really got exciting when the scientists found out what the gene did.<br />

Part four: From the protein village to the cosmopolitan city<br />

“A search of the literature showed that Ephx2 works in the heart,”<br />

Norbert says. “It modifies a type of hormone called EET found in cardiac<br />

muscle. EETs play a role in relaxing blood vessels and they also step<br />

in to help protect the heart after some types of heart attacks. When<br />

we checked levels of these molecules in SHHF animals, we found high<br />

amounts of Ephx2 and significantly lower amounts of EETs.”<br />

All the evidence suggested that the researchers had found a new<br />

heart failure gene, but there was more to do. If Ephx2 was the culprit<br />

in heart failure, they might be able to protect the animals by shutting<br />

it down. But that couldn’t be done in rats.<br />

Norbert sighs. “Here was the classical case – a model for human disease<br />

that was working in the rat, but we didn’t have the genetic tools to<br />

follow up.”<br />

The only alternative was to turn to mice, which have a very similar version<br />

of Ephx2 that seemed to have the same functions in the heart.<br />

But switching model organisms always makes things a bit complicated.<br />

In this case, there was no such thing as a spontaneously hypertensive<br />

heart failure mouse. The researchers would have to delete the<br />

gene in healthy mice and then find a way to simulate the type of heart<br />

damage seen in rats.<br />

Removing the gene caused no obvious problems for the mice – they<br />

had normal blood pressure, and their hearts were the usual size. But<br />

differences appeared when the animals were given a drug that caused<br />

the muscle cells of the heart to beat irregularly.<br />

Normal mice experienced symptoms resembling a heart attack.<br />

Things were different in the animals without Ephx2.<br />

“It was much more difficult to cause an irregular heartbeat,” Norbert<br />

says. “That was true when we used drugs, also with the use of electrical<br />

stimulations. Even when you subjected the heart to irregular electrical<br />

pulses, it kept beating on in a steady rhythm.”<br />

One more thing had to be done to prove that the case had any connection<br />

to human heart disease. The scientists needed to see how Ephx2<br />

behaved in human patients. They obtained samples from the cardiology<br />

units at the Charité to check whether there was anything unusual<br />

about the protein.<br />

“Jan and Judith found that Ephx2 levels were lower than usual in<br />

patients recovering from heart failure,” Norbert says. “That makes a lot<br />

of sense if you realize that EET hormones are needed for the recovery<br />

process. What you would want to do in that case is lower the amount<br />

of Ephx2 – so that it stops repressing the activity of EETs.”<br />

This fits with what scientists believe about the functions of the hormones,<br />

Norbert says. They seem to control channels that let charged<br />

atoms pass into and out of heart cells. That flow is the major factor in<br />

coordinating the activity of billions of cells, making them contract at<br />

282


the same time to create a heartbeat. It also fits neatly with a recent<br />

study in which drugs were used to control the activity of Ephx2 in mice<br />

with enlarged hearts. A group at the University of California-Davis<br />

showed that treatment with molecules called sEH inhibitors could<br />

prevent the condition from developing. An unexpected side effect of<br />

the study was that the same procedure protected the mice from<br />

arrhythmic heartbeats.<br />

Now Norbert – and the rest of us – know why.<br />

��<br />

If you’ve been waiting for the wolf in this story – the one in the title<br />

– it’s probably not what you think. Norbert hasn’t proposed using<br />

wolves as model organisms in the search for disease genes. At least<br />

not yet. In this case the wolf is a disease called Systemic lupus erythematosus,<br />

or SLE, or simply lupus.<br />

The disease was first described in the Middle Ages, and there are various<br />

hypotheses about why it was given the Latin name for the wolf.<br />

One explanation is that patients often develop a reddish rash across<br />

their noses and cheeks – supposed to resemble either patterns of fur<br />

on the face of a wolf, or scratches or bites made by the animal. These<br />

are the mildest symptoms of the disease.<br />

Sometimes the survival of the body calls for the sacrifice of some of its<br />

own cells, such as those infected by viruses. White blood cells track<br />

them down and prompt them to self-destruct in a process called<br />

apoptosis. If for some reason healthy cells become targets, the result<br />

may be an autoimmune disease such as lupus (systemic lupus erythematosus,<br />

or SLE). Lupus is dangerous because it can spark spontaneous<br />

inflammation anywhere in the body, eventually leading to permanent<br />

tissue damage and death. Most often affected are the circulatory<br />

and nervous systems, joints, and the skin. That’s why the unusual<br />

rash often develops on the face.<br />

Norbert’s group became involved through a collaborator. Min Ae Lee-<br />

Kirsch, a clinical researcher at the Technical University of Dresden, was<br />

investigating a family that suffered from a rare form of hereditary<br />

lupus. Working with Norbert’s lab, she carried out a linkage study that<br />

revealed defects in a gene called TREX1.<br />

That made them wonder whether other types of lupus might involve<br />

mutations in TREX1. They examined samples provided by clinical partners<br />

throughout Europe and found changes in the genes of patients<br />

suffering from SLE, the most common form of lupus. None of the samples<br />

from non-lupus patients had these types of mutations.<br />

Why should defective TREX1 cause an auto-immune disease?<br />

“The gene encodes a protein that can enter the nucleus and chew up<br />

the cell’s own DNA – one of the steps of normal apoptosis,” Norbert<br />

says. “Normally it doesn’t do so because it is tethered to structures<br />

outside the nucleus. It stays there until it’s time for the cell to die. And<br />

there’s evidence that it may have a second job; it may act as a sensor<br />

that monitors the cell for DNA brought in by viruses.”<br />

The mutation seems affect the part of the protein that ties it to its<br />

duty station, so animals with defective TREX1 may not be able to hold<br />

it there. The molecule escapes and spreads to other places, including<br />

the nucleus. Healthy cells die, releasing DNA and other molecules from<br />

the nucleus at the wrong time. The immune system misinterprets<br />

them as foreign and builds antibodies against them, which then leads<br />

to attacks against healthy cells.<br />

TREX1 is not responsible for all cases of SLE. Defects in other DNA-digesting<br />

proteins have also been linked to the disease. But the study may<br />

explain doctors’ long-standing suspicion that some cases of lupus are<br />

triggered by viral infections. If TREX1 really is a DNA sensor, it may<br />

detect viral molecules and launch its own defense by killing the cell.<br />

Because this is not the normal way the body becomes aware of an<br />

infection, the result may be an autoimmune disease.<br />

Norbert’s office is on the ground floor of the Medical Genomics building.<br />

The floor-to-ceiling windows on the north side look out onto a<br />

grassy field and a stretch of woods. He peers outside, lost in thought,<br />

and I instantly know what’s on his mind.<br />

“Wolves,” Norbert says.<br />

“Don’t even think about it,” I say.<br />

“Dogs do make excellent models for human diseases,” he muses. “And<br />

there’s definitely enough space on campus...”<br />

Don’t even think about it.<br />

283 Part four: From the protein village to the cosmopolitan city


Walter Birchmeier,<br />

Scientific Director of the <strong>MDC</strong><br />

What makes the Max Delbrück Center unique?<br />

For 15 years the <strong>MDC</strong> has been devoted to doing excellent science<br />

around a concept of molecular medicine, primarily in collaboration<br />

with the Charité and the universities of Berlin, more recently with the<br />

arrival of the Leibniz Institute for Molecular Pharmacology on the<br />

campus. What is happening now in science makes this a very fertile,<br />

attractive theme for basic researchers and clinical scientists, and the<br />

campus is appealing because of our experience.<br />

I think we handle interdisciplinarity in a unique way – our researchers are<br />

not typically limited by borders of technology, models, or model systems.<br />

Our scientists have been exploring new model organisms such as the<br />

naked mole rat, which Gary Lewin has introduced in order to get a grip on<br />

sensory mechanisms in the nervous system, and Nikolaus Rajewsky has<br />

brought in planaria, which he is using to look at microRNAs and issues<br />

related to regeneration. Alongside the very classical systems of the mouse<br />

and the fly, we are also doing original work with more established organisms<br />

such as zebrafish and the rat. There genetic tools have lagged<br />

behind, but groups like that of Norbert Hübner have been doing some<br />

creative work to try to overcome those obstacles.<br />

This willingness to branch out into new territory allows researchers<br />

such as Thomas Willnow, whose original focus was on the cardiovascular<br />

system, to make a foray into a neurobiological theme when a<br />

knockout animal takes him there. Or scientists like Claus Scheidereit,<br />

who began as a classical biochemist, to make real contributions to our<br />

understanding of the progression of specific types of cancer. Carmen<br />

Birchmeier is co-head of our Neurodegenerative Diseases program,<br />

Interview<br />

but her work has increasingly taken her into muscle development. And<br />

the context of the <strong>MDC</strong> has led to valuable collaborations between my<br />

own lab and clinically-oriented groups such as those of Ludwig<br />

Thierfelder and Peter Schlag. It’s a very satisfying feeling for a cell biologist<br />

to participate in work that has a clinical relevance. These are only<br />

a few examples that I could give. There are many others in this book.<br />

One unique aspect of our existence is the generous funding that we<br />

receive through the BMBF and the State of Berlin. It allows us to concentrate<br />

on science – mostly! – and helps a great deal in making outstanding<br />

recruitments. The <strong>MDC</strong> can push for a high quality of science,<br />

and our regular external evaluations have confirmed that we seem to<br />

be on the right path.<br />

Obviously there are many more unique features to the <strong>MDC</strong>, including<br />

the historical and cultural aspects of the campus and its situation in<br />

Berlin, which I think are well represented in this book and some of the<br />

other publications we have produced over the years. Our staff have<br />

taken the initiative in writing about the history of the campus and science<br />

in Berlin – one publication has just come out on some of the<br />

great figures in genetics that have had a connection to Buch. There<br />

have been other books and pamphlets on the campus’ many artworks,<br />

biographies of the scientists who have worked here, and even the<br />

campus flora. Then there are our activities with schools, and a very<br />

strong campus participation in the Long Night of the Sciences...<br />

There have been tremendous changes in science over the past<br />

15 years. How has this affected what people mean by<br />

molecular medicine and how it is practiced?<br />

284


Stefan Schwartze and Walter Birchmeier<br />

One huge change has come through the explosion of new technologies,<br />

including high-throughput platforms that have moved us from a<br />

very narrow focus to a much broader one. When I started at the <strong>MDC</strong>,<br />

the norm was for one professor to work on one gene. Now, potentially,<br />

everyone can work on every gene. That’s a danger because to make<br />

progress you need to focus. Everyone is solving this problem in his own<br />

way. In the case of our lab, we have concentrated on Wnt/β-catenin<br />

and Met-Gab1 signaling. Claus Scheidereit has focused on NFκB;<br />

Thomas Jentsch on channels and transporters... I could go on.<br />

In this environment some traditional areas of focus no longer really<br />

make sense. There have been discussions about whether institutes, for<br />

example, should focus on one disease type – whether the German<br />

Cancer Research Center should “take over” cancer, while we concentrate<br />

on cardiovascular diseases. Well, everything we are learning<br />

shows that common pathways and mechanisms are implicated in<br />

embryonic development, cancer, autoimmune diseases, and so on.<br />

Understanding a disease will require looking at a process from many<br />

different angles – and understanding a process will require looking at<br />

the role it plays in making a wide variety of phenotypes.<br />

As well as broadening our themes, labs have expanded the repertoire<br />

of methods and techniques that they use on a daily basis. We have so<br />

many more means at hand now to perturb and thus understand bio-<br />

logical systems. It’s amazing to remember that the knock-out technologies<br />

that we use in virtually every paper are relatively new.<br />

Conditional knock-outs are newer than that, and now at the frontier<br />

are small interfering RNAs. We have also built a platform for smallmolecule<br />

screening in which we actively seek inhibitors for the pathways<br />

and processes that we are working on. Those are crucial tools as<br />

you develop animal models of disease and want to manipulate them,<br />

as well as in the search for new drugs.<br />

Taking advantage of these technologies means that groups have to<br />

integrate new types of expertise – mastering mass technologies,<br />

learning to cope with data on a genomic scale, working with bioinformatics<br />

and new ways of modeling processes. Not every group can do<br />

everything, which means we need to flexible about the question of<br />

whether platforms are maintained centrally or by groups. The <strong>MDC</strong><br />

has this flexibility, and here once again the context of the Helmholtz<br />

Association is very helpful.<br />

An explicit mission of “molecular medicine” raises hopes that<br />

there will be some fast translation of basic science into therapeutic<br />

tools. Sometimes a lot of “hype” is generated around a<br />

particular approach – for example, gene therapies. There have<br />

been a few attempts at trying out virally delivered gene therapies<br />

in human patients, and some have led to unexpected<br />

285 Interview


Laying the cornerstone of the new building to house the 7-Tesla MRT machine. Left to right: Prof. Ernst Otto Göbel, Prof. Walter Rosenthal, Dr. Peter Lange, Dr. Hans-Gerhard<br />

Husung, Prof. Walter Birchmeier, Dr. Siegfried Russwurm.<br />

side effects or have been disappointing in other ways. On the<br />

other hand, your own work has contributed to new diagnostic<br />

tools that have saved lives. Do you see “molecular therapies”<br />

becoming common tools in the treatment of patients<br />

anytime in the near future?<br />

Molecular approaches have already been extremely valuable in diagnostics,<br />

and I’m convinced that molecular therapies will become common<br />

tools in medicine. Yet we have to be humble and do things we<br />

can. It’s worthwhile to screen small-molecular-weight compounds<br />

with the FMP, but you have to be realistic – only one in 500 lead compounds<br />

make it into the clinic. On the other hand there have been<br />

some impressive successes – Herceptin from Genentech/Roche,<br />

Gleevec from Novartis. Many more are in the pipeline: antibodies, vaccines,<br />

therapeutic proteins, and therapeutic cells.<br />

You have to beware of hype and fads. They come and go, and I’m<br />

always a bit relieved when they go. They’re often associated with unrealistic<br />

expectations and pressure to apply technologies that may not<br />

be ready. Stem cells are a good current example. It’s obvious that they<br />

Interview<br />

have a powerful potential – that doesn’t mean they can be injected<br />

right away in a desperate measure to cope with neurodegenerative disease.<br />

That would be dangerous – in animals stem cells operate operate<br />

in very controlled environments, in small “stem cell niches.” Their behavior<br />

in other contexts is unpredictable and may be dangerous.<br />

What are the most important things clinicians need to<br />

understand about basic research?<br />

That it’s hard work and takes time to understand a biological problem.<br />

Although there is still a gap and a difference in mentality, I must say<br />

that we have some excellent MDs at the <strong>MDC</strong>: Ludwig Thierfelder,<br />

Bernd Dörken, Friedrich Luft and others. And many excellent young<br />

ones. I hope that the ECRC will attract very high-quality new MDs,<br />

both seniors and juniors.<br />

What do you consider to be the main bottlenecks and<br />

challenges for the campus in terms of achieving its ambitions<br />

in molecular medicine?<br />

Basic science is doing well. Many of the collaborative projects could be<br />

improved, and we need to start new projects. All of the health-related<br />

286


Helmholtz Centers do molecular medicine now and have “translational”<br />

centers. So to stand out we will need to be more successful, and<br />

develop new projects. But we are in an excellent position. Things are<br />

moving forward with the ECRC, and just this week we have taken a<br />

major step forward with plans to create a new institute in Berlin for<br />

systems biology. I could also envision a major push in the direction of<br />

stem cell biology. What we have to learn in this field is still immense –<br />

on stem cell niches, on cancer stem cells, on the pathways that control<br />

the various participating cells.<br />

Institutes become marked by the character and style of their<br />

directors. You are only the second director of the <strong>MDC</strong>, having<br />

taken over the reins from Detlev Ganten, who was obviously<br />

very strongly associated with the <strong>MDC</strong>. How have you put<br />

your own stamp on the <strong>MDC</strong>? What do you think your legacy<br />

will be?<br />

I was an intermediary, an internal appointment for five years, and during<br />

a lot of this time a search for a new director has been going on.<br />

That has limited my maneuvering room. I think that if I weren’t a scientist,<br />

the situation would have been very frustrating. But this situation<br />

has given us the chance to search intensively for an excellent new<br />

person to direct the institute for the future, someone with experience<br />

in the outside world, an excellent, widely respected scientist, who<br />

understands molecular medicine.<br />

One of my own goals was to bring more outstanding scientists to the<br />

<strong>MDC</strong>. For a while in the 1990s, this was a concern. I think here we have<br />

had excellent success in the recruitment of Thomas Jentsch, Nikolaus<br />

Rajewsky, Norbert Hübner, and Gary Lewin, as well as youngsters like<br />

Matthias Selbach, Oliver Daumke, and Ines Ibanez-Tallon. I could list<br />

many others.<br />

287<br />

I also wanted to push for excellent science, and the recent review of<br />

our proposals for the next five-year funding period was extremely positive.<br />

All of the programs had improved – in some cases very significantly<br />

– over the past five-year period. Overall, the <strong>MDC</strong> will certainly<br />

rank among the best – and will possibly have the best overall rating –<br />

of all five of the Helmholtz health-oriented centers. I am particularly<br />

pleased about the performance of the program on neurobiology,<br />

which had the highest rating of “seven” in the review. One consequence<br />

will be that nobody will try to restrict the institute’s focus to<br />

only cardiovascular disease.<br />

What advice will you have for your successor?<br />

To keep practicing his or her science.<br />

How will you spend all of your “new free time”?<br />

To rebuild my own lab. We have had some nice stories – Jolas’ work on<br />

wound healing, Martas’ work with NOMA-GAP, Ute’s Gab1 mutants,<br />

Klaus’ work on a Shp2 inhibitor, Dietmar’s studies of β-catenin in the<br />

brain, and the role of Wnt and Bmp in heart development from<br />

Alexandra. Getting such papers published in the best possible journals<br />

is vital to these young scientists’ careers, and it’s something that I can<br />

help with, if I have the time.<br />

What have been the hardest decisions you have had to make<br />

as director?<br />

To tell to some of our young group leaders that they will not be<br />

tenured. It’s heartbreaking, particularly since some of them have been<br />

so outstanding previously.


Walter Rosenthal,<br />

Director of the Leibniz Institute<br />

for Molecular Pharmacology<br />

What makes the FMP unique?<br />

The FMP is unique partly because of its scientific vision and the people<br />

and tools we have assembled to accomplish that vision; it is also special<br />

because of the way in which we operate as a member of the<br />

Leibniz Association, which gives us a great deal of scientific and organizational<br />

freedom.<br />

I came to the FMP in 1996. Like many other scientific institutes that<br />

had formerly operated under the GDR, the FMP was still in a process of<br />

evolution. Before the “Wende” it had been called the “Institut für<br />

Wirkstofforschung.” At that time there had been a good mixture of life<br />

scientists and chemists among the staff, because it had worked in conjunction<br />

with pharmaceutical companies. When a scientific council<br />

was appointed to guide the formation of the FMP, they originally<br />

wanted to make it into an institute with a focus on cell signaling. But<br />

the major development that we started in 1996 was to significantly<br />

expand structural biology within the FMP. My predecessor had already<br />

begun to push this area. Our vision was to establish a strong program<br />

in structural methods centered around NMR. That has worked out very<br />

well: scientists of the FMP have made very important contributions,<br />

not only in NMR applications, but also in the development of methods<br />

like solid-state NMR.<br />

In a subsequent phase, we wanted to strengthen the chemistry activities<br />

again. Michael Bienert was here, doing peptide chemistry. To that<br />

we have added a Medicinal Chemistry group focusing on combinatorial<br />

chemistry and the Screening Unit. Soon we’ll have a new group<br />

here that makes artificial proteins. What we have achieved is to estab-<br />

Interview<br />

lish a good mixture of cell biologists, signal transduction people,<br />

chemists and structural biologists. This mix represents a potential we<br />

should use to work on rather difficult projects. For example, several of<br />

our groups are working on membrane proteins such as receptors and<br />

channels; these are among the most important areas where structure<br />

elucidation and drug design are concerned, yet they are also among<br />

the most difficult. Finally, we now have groups working on<br />

protein-protein interactions, hoping to develop approaches to interfere<br />

with them pharmacologically. A research group funded by the<br />

German Research Council (DFG) has now been established with that<br />

focus.<br />

We work without restricting ourselves to particular organs or diseases.<br />

The ideal projects require experience from groups from all the<br />

subdisciplines represented here at the FMP. Teams from these areas<br />

already work well together under our roof, but many scientists are still<br />

unaccustomed to working this way. Our aim is that everybody in the<br />

house – from PhD students to established scientists – should take<br />

advantage of the resources of the entire institute. When we hire we<br />

look for people who will be able to do that. And interdisciplinarity is<br />

also a hallmark of our new Leibniz graduate school, which has just<br />

been established thanks to an application submitted by Bernd Reif. It<br />

has been set up so that the students get real exposure to a spectrum<br />

of platforms and scientific themes. It’s a good mirror of our overall<br />

?concept.<br />

You have said that institutes like the FMP are evolving a new<br />

role in the process of drug design...<br />

288


Today we have this vision of a “rational” drug design which will operate<br />

not on a basis of blind trial-and-error, but on our knowledge of<br />

how molecules are structured and how they behave. The plan is to find<br />

better and easier ways to identify good targets for drugs, to quickly get<br />

a look at the details of their surfaces, and to engineer artificial molecules<br />

that will change their behavior. Achieving this will require a great<br />

deal of work collecting basic knowledge about the structures and<br />

functions of molecules and processes in the cell. Large pharmaceutical<br />

companies will not invest in such high-risk basic issues, although it’s<br />

clear that this type of research is essential to drug design. It is better<br />

carried out in an academic setting. So we’re moving towards new<br />

types of partnerships between these communities.<br />

Science has been changing rather dramatically over the past decade.<br />

Most research groups have access to a wide range of technologies that<br />

have enabled them to bring their work much farther along the road to<br />

applications. But there are still some important gaps. For example, as<br />

far as I know, there are no open platforms in Germany where academic<br />

researchers can carry out screens for small molecule inhibitors of<br />

the molecules they are interested in. They might be looking for something<br />

that can be developed into a drug, or they may simply need<br />

some sort of a probe that can be used for basic investigations of the<br />

cell. They will only get access to a screening platform by paying for it –<br />

which is usually too expensive – or by getting a company interested in<br />

the project. In most cases that’s unlikely to happen.<br />

We’re helping to fill that gap by setting up a major new screening unit<br />

that will cater particularly to academic users. This will give a boost to<br />

their basic work; in the case of those interested in drug design, they<br />

will be in a better position to hook a pharmaceutical company. They<br />

may leave the facility with several hits – compounds that bind to<br />

potential drug targets. There are several more steps that have to be<br />

289 Interview<br />

Walter Rosenthal


taken after that, including optimizing the compounds and testing<br />

them in cell culture systems or laboratory animals. The chemistry<br />

groups we have put into place can help there. Again, the farther along<br />

a group brings a project, the more likely they are to catch the interest<br />

of a company.<br />

We’re working in a context where, over the last 20 years, Germany has<br />

steadily lost its position as the world’s major producer of pharmaceuticals.<br />

We hope that the platforms and expertise we have assembled<br />

will make the region attractive to those who want to invest in<br />

Germany and Berlin.<br />

How does the FMP fit into the research landscape in Buch and<br />

Berlin? In Germany? In Europe?<br />

Going back to the 1970s, there was the Institut für Wirkstofforschung,<br />

the IWF, a part of the GDR Academy of the Sciences, which was housed<br />

in Berlin Friedrichs?felde. After the reunification scientific reviewers<br />

came in to evaluate its activities and to think about its future. They<br />

were impressed by some of its hallmarks, especially the interdisciplinarity.<br />

There was already a tradition of working with animal models,<br />

doing chemistry, research into signal transduction, and clinical pharmacology<br />

work. There was even a small company alongside the institute,<br />

a unique “academic industrial complex.” That type of combination<br />

was a new approach for the West. The recommendation of the<br />

review was to transform the institute into the FMP, with a focus of<br />

molecular pharmacology. It was opened in January, 1992.<br />

It was clear that Friedrichsfelde wouldn’t do over the long term; the<br />

old buildings weren’t adequate for our needs. It was very difficult to<br />

set up cell biology labs there, for example. The only place big enough<br />

to house our NMR machines was a big open building with a huge iron<br />

beam running across the ceiling. When asked why<br />

it was there, they said it had been used to hoist dead elephants or<br />

something, as once the buildings were used by a veterinary institute<br />

that worked for the animal park nearby.<br />

The move to Buch was a very good decision. We have very strong partners<br />

on the campus that we can support and get support from, and<br />

are working closely with them on several very interesting projects. Our<br />

most important partner in Buch is the <strong>MDC</strong>. The research concepts of<br />

the two institutes complement each other: while the molecular medical<br />

research at the <strong>MDC</strong> is particularly dedicated to diseases or clinical<br />

symptoms and their molecular explanations, the FMP investigates<br />

the functional and structural characterization of proteins as well as<br />

the development of strategies for influencing them pharmacologically.<br />

The close connection to the <strong>MDC</strong> extends to the organizational level.<br />

Thus, large equipment is shared and jointly operated. Guest scientist<br />

contracts make it possible for scientists of one institute to use equip-<br />

Interview<br />

ment belonging to the other. Both establishments send representatives<br />

to important committees of the other. The planning of costly and<br />

long-term research projects as well as the appointment of leading scientists<br />

takes place in joint agreement. The <strong>MDC</strong> and the FMP arrange<br />

and finance joint events for those studying for their doctorates.<br />

In terms of Germany, the FMP is a member of the Leibniz Association,<br />

a national research organization similar to the Max Planck Society, the<br />

Helmholtz, and the Fraunhofer Associations. Leibniz institutes retain a<br />

great deal of independence but also profit from the association. They<br />

perform problem-oriented research of national interest and strive for<br />

scientific solutions for major social challenges.<br />

Another hallmark of Leibniz is the fact that we promote very strong<br />

links to the university system. At the moment the FMP has six full staff<br />

with joint appointments (three full, two associate and one honorary<br />

professorships).<br />

One particular feature of Leibniz is that its institutes are not guaranteed<br />

a permanent existence. Every seven years, the FMP – like every<br />

other Leibniz institute – undergoes a thorough scientific evaluation.<br />

The question is whether the institute is needed – whether it is fulfilling<br />

its mission, and whether the same scientific work could be accomplished<br />

without it. This challenges us to perform at a very high level<br />

and to be responsive to the shifting landscape of science. In some<br />

cases, of course, institutes are closed – or they leave the organization<br />

and find another mode of funding. But new institutes are joining. Very<br />

soon we will add a research laboratory in Berlin whose focus is<br />

rheumatology, and another new member will be the Berlin Natural<br />

History Museum.<br />

This system also allows for expansion of institutions that are doing an<br />

excellent job. Despite a period of financial problems in Berlin, we<br />

haven’t suffered – just the opposite. We have received budget increases<br />

and hope for more, particularly in investments into NMR and other<br />

major equipment. Recommendations for growth and other types of<br />

changes arise from the evaluations. We also have had quite strong<br />

support from the scientific council in achieving our plans.<br />

As far as our relationships to Europe are concerned, the FMP has participated<br />

in several European research networks over the last years.<br />

Since October 2006 Enno Klußmann has been the first FMP scientist<br />

to coordinate a project funded by the European community (STREP<br />

“Identification of the therapeutic molecules to target “compartmentalized<br />

cAMP signaling networks in human disease (thera-cAMP),”<br />

under the 6th framework programme).<br />

Facing the European deficit in Chemical Biology, the FMP coordinates<br />

activities in the field of small molecule screening. One of these activities,<br />

the infrastructure initiative “EU-OPENSCREEN”, will provide access<br />

to screening platforms for european academic groups. Thus, local<br />

290


expertises of research institutions like the Screening<br />

Unit of the FMP can be used throughout the entire<br />

European research area.<br />

What’s life like for scientists here?<br />

We have junior and senior groups. As of November<br />

2006, we standardized the system so that junior<br />

group leaders are not initially hired into tenure-track<br />

positions. They are very well funded for an initial period<br />

of five years; that can be extended for another four<br />

years. All of our junior group leaders have gone on to<br />

find positions, becoming chairs of departments or<br />

receiving C3 professorships. About two-thirds of our<br />

alumni move into academia; the remaining third go<br />

to work for drug and biotech companies.<br />

The fact that we use tools similar to the pharmaceutical<br />

industry means that people who receive training<br />

here should be interesting candidates for biotech<br />

companies. To promote this link, the campus has set<br />

up a summer school – “From target to market” – to<br />

help prepare postdocs and predocs for careers in<br />

industry. The next course will be offered in September.<br />

What do you consider to be the most significant<br />

bottlenecks for the institute and the field<br />

in general?<br />

Right now I would like to add a team of synthetic<br />

chemists who can work on hits which we have identified<br />

in our Screening Unit. There is also a general bottleneck<br />

in academia when it comes to support for cell<br />

biologists and biochemists to synthesize new tools.<br />

We have one excellent group here, but that’s not<br />

enough. As an important step toward a solution,<br />

we’re setting up a national network of chemists and<br />

biologists (ChemBioNet). Our ambition is to further<br />

expand those activities and to become a national<br />

platform to promote this symbiosis of chemistry and<br />

biology.<br />

In classical academia, you characterize a target and<br />

usually stop there. However, while maintaining the<br />

focus on basic research, we strive to extend our<br />

pipeline. We take projects to the point that we have<br />

small molecule inhibitors for a protein, including<br />

small molecules to disrupt protein-protein interactions.<br />

This is a significant contribution to narrowing<br />

the gap between pharmaceutical companies and<br />

academia.<br />

291


Detlev Ganten,<br />

CEO of the Charité,<br />

Universitätsmedizin Berlin and<br />

Founding Director of the <strong>MDC</strong><br />

The <strong>MDC</strong> was established with a vision of blending fundamental<br />

and clinical research in a new way, in hopes of making<br />

it easier to translate findings from basic science into something<br />

that would be medically useful. How did that vision<br />

evolve while you were at the <strong>MDC</strong>, and what is your take on<br />

the progress of molecular medicine today?<br />

When I came to Berlin-Buch in 1991, genomics was the name of the<br />

game. I was a pharmacologist who was already starting to take a<br />

“genomics” approach to hypertension in the work of my own lab, and<br />

many others were doing likewise – even though we didn’t yet have a<br />

national genome program in Germany at that time. But we were<br />

already beginning to expand our focus beyond small numbers of molecules<br />

and single processes, trying to see things in terms of complex<br />

relationships within genomes. Almost at the same time, in 1992, the<br />

idea of gene therapies started to generate a lot of enthusiasm. The<br />

idea was that we could exchange a bad gene for a good one, using a<br />

virus or some other type of delivery vehicle, a “gene taxi.” Well, the<br />

results of this work have been mixed, which is to be expected considering<br />

that the methods are in their infancy. But we had very stimulating<br />

and successful scientific meetings on gene therapy in Berlin-Buch.<br />

The focus has now shifted as we realize that genome-wide phenomena<br />

crystallize at the level of cell biology and systems biology and many<br />

of the things we want to do will require a thorough understanding of<br />

what happens at this level. I remember when experiments from<br />

Walter Birchmeier’s lab on a signaling protein called the hepatic scatter<br />

factor started us thinking about the mechanisms that underlie<br />

Interview<br />

metastases in liver cancer. At the time I wasn’t really sure how this<br />

would develop and whether it would be a basic mechanism that<br />

would be of interest to many people. But you see how that has turned<br />

out – everything we discover about the fundamental cell biology of<br />

these crucial processes has the potential of becoming something very<br />

powerful in diagnostics and possibly treatments. Another lab in which<br />

this relationship is very clear is that of Thomas Jentsch, who is clearing<br />

up basic issues about the functions of membrane proteins that play<br />

important roles in several diseases.<br />

The term “molecular medicine” implies that there will be detailed, indepth<br />

studies on basic molecular mechanisms, and there will be a<br />

patient-oriented medical outcome of your work. It gives you an orientation<br />

when you discover that a molecule such as β-catenin or NFKB<br />

controls migrations or the cell cycle. You don’t stop at saying, “What<br />

does this do during embryonic development?” – you go on to see<br />

whether the process also takes place in cancer cells, in cardiovascular<br />

tissue or in the brain. It might seem like an accident that a scientist<br />

like Thomas Willnow finds a connection between receptor proteins<br />

and Alzheimer’s disease, but it’s not, not really. If it weren’t Alzheimer’s,<br />

it would be something else. Today’s biggest killers – in the developed<br />

world, and increasingly in the rest – are diseases that stem from fundamental<br />

problems in cells. So to deal with them we’re going to have<br />

to bridge the gaps between molecules, cells, organ function and the<br />

whole body in health and disease.<br />

The idea was to gather clinicians and basic researchers on one site, to<br />

have them share students so that there would be an exchange of<br />

292


Detlev Ganten<br />

293<br />

methodologies and questions between basic research and<br />

the clinics. While the approach has been successful, we still<br />

haven’t completely solved the culture gap between patient<br />

care in the clinic and laboratory work. It’s especially gratifying<br />

to see how many clinicians have really become interested<br />

in the mechanisms of the diseases they confront every<br />

day, and unraveling connections between those and much<br />

rarer diseases. Maybe it’s easier for clinicians, who are concerned<br />

with practical problems and real patients, to<br />

become interested in the theoretical side of things. For<br />

researchers the goal has always been to publish in great<br />

journals. That has been the goal so long that it’s harder to<br />

escape. They may respect clinicians for their ability to treat<br />

patients, but that’s not the main criterion for success in science.<br />

So while it’s already a great step forward to collect<br />

people under one roof – and that’s essential – it is not a<br />

guarantee that they will develop the right mentality to<br />

really collaborate.<br />

But in Buch it works, and I can cite many examples. When<br />

Claus Scheidereit was at the Max Planck Institute he was<br />

working on NF-ΚB, doing basic biochemistry, without thinking<br />

much about any real applications. Then he came to the<br />

<strong>MDC</strong> and is now working with Bernd Dörken and Friedrich<br />

Luft, trying to understand how lymphomas develop and<br />

how blood vessels are damaged by hypertension and what<br />

kind of new therapies might be possible. The same thing<br />

happened with Achim Leutz, who converted from being a<br />

cell biologist to a person interested in leukemia. From the<br />

clinical side – it was the constellation of combining things<br />

that attracted people like Ludwig Thierfelder, a cardiologist<br />

who has become interested in experimental genetics. Peter<br />

Schlag, who arrived with a high-tech perspective on surgery,<br />

realized what could be gained through collaborations<br />

with basic science and cell biology.<br />

Those interactions should be pushed even more<br />

through the ECRC, which is now being created in<br />

conjunction with a Clinical Research Center on the<br />

Charité side...<br />

Yes. Clinical and experimental science have to be under one<br />

roof. They have to have labs next door to each other – even<br />

better is to have clinicians and basic scientists working<br />

together, within one lab. Originally, the two academic clinics<br />

in Buch were in danger after German reunification<br />

because they were small and science-oriented; they cost<br />

money. Fortunately, they became part of the Charité university<br />

hospitals. When HELIOS took over, there was a big dis-


cussion about where the new hospital would be built – one plan was<br />

to put it near the highway, which would have had some advantages in<br />

terms of infrastructure. But we fought for two years to get the hospital<br />

right next to the science campus where it is now. The clinics and<br />

research labs have to be one community.<br />

One thing you were involved in was the development of the<br />

Helmholtz Association itself as a national organization; in<br />

fact, you served as its president.<br />

You have to understand the history of Helmholtz, where it came from.<br />

After World War II, in the 1950s Germany was rebuilding its science<br />

infrastructure, and one key decision was to start working again on<br />

atomic and nuclear physics for peaceful purposes. Other types of scientific<br />

institutes that had been strong before the war – such as in aviation,<br />

energy research, etc. – also had to be rescued or restructured.<br />

Several institutes were founded; they were completely independent of<br />

the academic university structure and each of them was directly supported<br />

through government funding. Of course that caused problems<br />

– they were too strongly linked to politics, and they needed scientific<br />

and administrative freedom. Since a number of institutes were in the<br />

same boat, they decided to organize themselves as an association to<br />

represent their joint interests. Before I became president, they decided<br />

they needed a corporate identity of their own. That required long discussions,<br />

and we decided it would be the “Helmholtz Association of<br />

National Research Centers.” The government was not really amused by<br />

this move – they were afraid of losing control of the programs – and<br />

didn’t want to leave us alone. We agreed that we were bigger and scientifically<br />

as good as the Max Planck Institutes and needed academic<br />

independence.<br />

Then we formalized the institutions of the Helmholtz association, and<br />

created a senate. The ministry was present in the senate but didn’t<br />

preside over it, so we slowly developed independent structures. The<br />

Interview<br />

next step was to organize ourselves across the centers in scientific<br />

programs such as health, energy, and the environment. Each of the<br />

institutes was still independent and whenever something important<br />

happened we would have to go directly to the ministry because they<br />

provided the money. Now the situation is different. A major part of funding<br />

comes through grant money and third-party funding from applications<br />

made by groups working on common themes across different<br />

institutes – so for example a certain amount of program funding will be<br />

devoted to cardiovascular and metabolic research, most of it going to<br />

the <strong>MDC</strong> but some to the DKFZ and other groups. This is still an ongoing<br />

process; Helmholtz wants to develop a corporate identity of its own, and<br />

ensure its academic freedom and independence from politics.<br />

I took up the job of the Helmholtz presidency because I felt that in<br />

Germany the many research organizations and their separation from<br />

the universities leads to fragmentation and should be better integrated.<br />

Things are still evolving; it will be interesting to see how everything<br />

turns out. But the new president of Helmholtz is a strong integrator,<br />

bringing the centers and insitutions together.<br />

You invested a lot of effort in trying to build up Buch itself as a<br />

community, and to draw new institutes and businesses here...<br />

Everybody needs a long-term vision. After German reunification, there<br />

was enormous emotion going through the country. Buch is a nice<br />

place with a strong tradition and a lot of potential; it would have been<br />

a tragedy to lose that. Such places needed to be protected and developed.<br />

The town of Buch itself hasn’t evolved as well as we would have<br />

liked, but I believe that will change as the huge HELIOS complex develops<br />

its infrastructures, and as projects like the Biotechnology Park and<br />

the proposed new Life Sciences Center develop. The constellation we<br />

have achieved here is interdisciplinary and very special and is also<br />

receiving support from the Senate of Berlin and other directions. The<br />

current president of Humboldt University is a theologian, and under<br />

294


his presidency the interdisciplinary life sciences have become a major<br />

focus of the university’s development, which is also good for Buch.<br />

At the beginning things were different, and there was quite a bit of<br />

anxiety about how things would turn out. In the very beginning I<br />

talked to the architect Prof. Wischer, who did the general planning of<br />

the campus. We had meetings and walked around on the campus,<br />

looking at leaking roofs and we produced a big list of things that needed<br />

to be done. He said, “Don’t get too focused on small problems; think<br />

about long-term perspectives.” Prof. Wischer organized a meeting with<br />

architects and engineers, landscapers and gardeners. He asked them,<br />

“What should this campus look like in 50 or 100 years? Don’t think<br />

about money, think about possibilities.” We drew up a plan and then<br />

began to realize it piece by piece. At that time we would not have<br />

believed that our dreams would ever become a reality. But within the<br />

last 15 years we got the FMP, the communications center, the new animal<br />

facility, and the administrative building, the mensa, the new buildings<br />

of the technology park, and the medical genomics center. The<br />

biotechnology park is not to be underestimated because it has the<br />

potential of systematically linking basic research, clinical applications,<br />

and the economic and commercial use of research results, patenting<br />

licenses and so on. The Gläsernes Labor is a relatively small activity but<br />

with a huge impact in communicating with the public. The plans for a<br />

new “Life Sciences Center” are a logical consequence of it.<br />

You have been very engaged in promoting the public understanding<br />

of science; the Gläsernes Labor is a fantastic<br />

platform for reaching school children.<br />

Everything you do in this area is important and eventually returns benefits<br />

to you. Such a lab has a high value for the campus, and not only for<br />

children. I’ll give you an example: Walter Jens is a friend, the President of<br />

the Academy of Arts, and he is a very influential person in Germany. I had<br />

known him a long time. He is very interested in medicine, but he has<br />

been extremely critical of genomics, for sort of classical reasons. He felt<br />

that it would lead to genotyping and exposure of intimate private<br />

details about people’s lives. Well, I invited him to come with some of his<br />

people. He visited the Gläsernes Labor and learned what a genotype and<br />

phenotype meant and then we had lunch. He said, “I understand now –<br />

it’s just a methodology. I largely overestimated the conclusions that can<br />

be drawn from genome analysis.” The visit to the Gläserne Labor, and the<br />

hands-on experience with gene technology and modern genome<br />

methodology, converted him from a critic to a supporter. Convincing just<br />

two or three people like Walter Jens, who write books and for journals<br />

and are influential, can change German science politics.<br />

Regarding the industry side of the campus... Wasn’t the idea<br />

originally that the companies would be spin-offs of the lab, or<br />

provide services for them?<br />

In the beginning that was more the case. The first companies were<br />

spin-offs of the academy institutes or the <strong>MDC</strong>. The most prominent<br />

example is Eckert and Ziegler, who started up with intellectual property<br />

from the <strong>MDC</strong> and established a company and got on the market.<br />

Several others such as Mr. Bendzko and Dr. Fichtner started that way.<br />

Later there was an influx of companies from the outside who liked the<br />

campus, found the conditions there very attractive for start-ups, and<br />

were also enthusiastic about being in an environment with the<br />

research institutes and clinics. The Biotech campus is flourishing and<br />

there is still space for new spin-offs. Like always, success such as this<br />

depends on opportunities and institutions, but especially on people.<br />

Dr. Gudrun Erzgräber still is the energetic and creative motor behind<br />

the campus management group.<br />

Now you have moved to the Charité – which must give you a<br />

unique perspective on the relationship between the clinical<br />

and research worlds. What has been the biggest adjustment<br />

you’ve had to make?<br />

If you are long enough at an institute like the <strong>MDC</strong>, you know almost<br />

everybody. You have hired important people, you know the institute in<br />

depth. Maybe you haven’t read every paper produced by the institute, but<br />

you know people’s work, their reputations, the grants they have. A scientific<br />

institute like the <strong>MDC</strong> is managed in a very personal way by interactions<br />

directly with the people and helping to solve their problems.<br />

Well, then you come to the Charité and find yourself confronted by<br />

15,000 people on four different sites, across the city of Berlin, covering<br />

themes from basic research to clinics. There is a huge amount of<br />

administration and technical support to manage. You can’t even know<br />

all the important people. This type of organization is completely new<br />

for me and I learned a tough lesson: you can only lead strategically. You<br />

can define a direction and then you have to select a few, able people<br />

who are able to translate your ideas into the institution. If you don’t<br />

assemble the right team then you are lost.<br />

Fortunately the Charité is an excellent, unique institution with a great<br />

tradition and a fantastic potential and really excellent people in<br />

research and in the clinics, as well as in the administration. The Charité<br />

is a very important institution for Berlin and thus has a lot of political<br />

support. In the fourth year after the merger of the medical faculties of<br />

the Free University (West) and the Humboldt University (East), the<br />

“Charité-University Medicine Berlin” has made great progress. We have<br />

a development plan above 500 million Euro for the next seven years<br />

and Nature has called the Charité “the beacon of reform” in the<br />

German University system. So it appears we are reaping the fruits and<br />

it has been worth all the efforts.<br />

295 Interview


Gudrun Erzgräber,<br />

Managing Director,<br />

BBB Management GmbH<br />

What were the circumstances surrounding the founding of<br />

the BBB?<br />

The BBB was founded in the summer of 1995 as an offshoot of the<br />

Max-Delbrück Center. Why? There was a recommendation from the<br />

Scientific Council of FRG that, alongside its other functions, the campus<br />

should host related industrial research activities - that meant we<br />

needed to make an effort to bring in companies. The Council was well<br />

aware that a considerable amount of applied research had been carried<br />

out in the GDR institutes that had been located here, and saw that<br />

as a potential area for growth. In 1992 the <strong>MDC</strong> began operations on<br />

campus. At the time it was the only research institute here, and it<br />

acquired responsibility for managing the campus. To handle that it<br />

founded a local management group, and I was put in charge of organizing<br />

it. At the beginning the group was responsible for the site<br />

management, in terms of supply and the technical infrastructure –<br />

what we’re doing today, on a much smaller scale. At the time we still<br />

belonged to the <strong>MDC</strong>.<br />

Then came a point where we said, if our goal is really to attract companies,<br />

then we’ll have to be able to apply for external financial support.<br />

The GEBA – which is responsible for providing such support on<br />

behalf of the state of Berlin – said that these activities were outside<br />

the normal domain of a research institute and to be eligible, the <strong>MDC</strong><br />

would have to establish a company. So in 1995 the Senate passed a<br />

resolution that permitted the <strong>MDC</strong> to found the BBB. This was a new<br />

public mission, it led to the creation of the BBB as a spin-off firm, and<br />

the necessary start-up capital came from the state of Berlin through<br />

Interview<br />

the <strong>MDC</strong>. I brought along two colleagues from the <strong>MDC</strong> and that’s<br />

how the BBB got its start.<br />

On January 1, 1996, I was officially appointed business manager. At the<br />

time I had no idea what I was getting into! But there are two things<br />

about the job that I have really enjoyed: solving difficult problems –<br />

and there have been a few of those! The other thing is the opportunity<br />

to communicate with a wide variety of people – with the companies,<br />

with the scientific institutes, with politicians, and many others.<br />

We knew that things wouldn’t be easy.<br />

The biggest challenge for me as manager has been to handle the<br />

financial aspects of the company. If we were going to apply for financial<br />

support, it meant we would also have to bring our own money to<br />

the table, and that meant borrowing it and assuming debts. But we<br />

weren’t per se eligible for credit – of course we had backing from the<br />

<strong>MDC</strong>, but that wasn’t a usual situation for a firm. After complicated<br />

negotiations with the state, we obtained 5.5 million. We had to pay<br />

that back ourselves, including the interest, from rent income. We did<br />

obtain start-up money between 1996 and 1998, because it was obvious<br />

that there wouldn’t be much income in the initial period. Since then<br />

– since the beginning of 1999 – we have been financially independent.<br />

My goal was to see that we broke even, and maybe even do a little better.<br />

Not to make a profit – the BBB wasn’t there to make money for<br />

itself; that wouldn’t have worked anyway, since the buildings were 80to-90<br />

percent funded with state support. Instead, what we earned<br />

could be rolled back into the campus, to improve the infrastructure<br />

and make things easier for the companies. We’re proud that we have<br />

296


Gudrun Erzgräber<br />

managed – there was a very difficult period from 2003 to 2004. That’s<br />

when we opened the building we’re sitting in now, and at first there<br />

was only one occupant!<br />

I can remember when State Secretary Strauch came for the opening<br />

ceremony... I was standing in front of the building, it was a beautiful<br />

day in June, and I thought – “If they only knew how I feel right now!<br />

With this empty building...” The technology boom had crashed; there<br />

hadn’t even been any expressions of interest from companies... Very<br />

difficult. We had to stretch our accounts to the limit just to pay our<br />

bills.<br />

Then there was a revival of the incubation business. A succession of<br />

new businesses moved here. 2005 was our best year, an amazing year<br />

in which 12 new companies moved onto campus. In 2006 it was four,<br />

and 2007 four, to the point that today 88 percent of the available<br />

space is occupied. And new requests are coming in. Of course from<br />

time to time one of the companies leaves; some shrink, some grow –<br />

an example is Silence Therapeutics, which is currently in a growth<br />

phase. We have a very good anchor in Eckert and Ziegler. In addition to<br />

their own building, they rent space from us.<br />

What have been the major phases in the development of<br />

the campus?<br />

The first companies that came here in many cases worked with equipment<br />

that was left over from the former institutes; many of them worked<br />

part-time, in lab space that belonged to the <strong>MDC</strong>. I don’t know if<br />

you have seen pictures from those days, but everyone was in barracks<br />

– the campus was full of barracks. You can imagine that biotechnology<br />

companies in barracks, without proper lab equipment – that just<br />

doesn’t work. So our first goal was to improve the infrastructure and<br />

297 Interview


uild laboratories. The first stage was in 1998, with the renovation of<br />

the Oskar and Cécile Vogt House as well as house 79. Then we began<br />

building here on the “industrial strip.” In 2001 we opened buildings 80<br />

and 82, and in 2003 building 85 was finished. Those were the three<br />

major phases. Another important step was the establishment of the<br />

Gläsernes Labor, which now does such a wonderful job with school<br />

children. For this we made use of one of the oldest houses here.<br />

Professor Ganten had the idea of setting up these teaching activities,<br />

and he said, “See how you can finance this!” We renovated the house<br />

and rebuilt the upper floor, and set up an exhibit room as a meeting<br />

point for businesses, scientists, and clinical research. In doing so it<br />

became part of the state supported incubator facilities. Otherwise<br />

that would never have existed here – as valuable as it has been, it was<br />

simply too expensive in the start-up phase. There, too, we depended on<br />

public funding; the <strong>MDC</strong> could never have financed it on its own.<br />

When the FMP established itself on campus, in 2000, it “bought into”<br />

the BBB, by taking a share of 20%. Another very important moment<br />

came during the national BioRegio competition, and that brought<br />

Schering, which came in at another 20 percent – that’s now the phar-<br />

Interview<br />

maceutical concern Bayer Schering. That’s the status today – the <strong>MDC</strong><br />

has a 60 percent share and the other two partners the rest. The<br />

Charité has applied to become another shareholder. There are still formalities<br />

to be faced before that happens. I’m hoping it will come before<br />

the end of the year. But according to the rules, the <strong>MDC</strong> will remain<br />

majority shareholder.<br />

Why is it important today for there to be a significant company<br />

presence on campus, even if their primary mission is not to<br />

provide services to the research groups?<br />

It was fully clear to us that firms on campus would not subsist on services<br />

they provided to research groups, or be mainly dedicated to developing<br />

products that come from the research laboratories. It was never<br />

intended just as the incubator for firms arising from the <strong>MDC</strong> and<br />

FMP. There would never be enough intellectual property at the right<br />

stage of development to launch start-up companies on this scale.<br />

Some of the firms do have collaborations with the institutes, but overall<br />

the biotechnology park was intended to have a regional and superregional<br />

function. From the very beginning the companies would<br />

serve the clinics in Buch, but they are global actors, but could also<br />

298


serve local partners. The campus has a greater potential than just playing<br />

host to excellent research institutes. And as we develop that<br />

potential, it pushes the development of the institutes.<br />

The campus offers good infrastructures to companies. If you’re seeking<br />

collaborations, you can find them here. You can go to talks, to workshops,<br />

there is an excellent library, and an IT infrastructure...<br />

It’s a bit harder to say what the typical scientific group gets out of a<br />

strong company presence on campus. From my point of view, there<br />

could be more interest in this interaction. I understand that perfectly<br />

well – I have a research background myself. You tend to get very focused<br />

on getting good results and publishing good papers and you often<br />

forget to look at what’s going on around you. The most important<br />

thing is that when businesses express an interest in collaborations,<br />

that there’s a resonance from the side of the laboratories. And an<br />

important point, obviously, is that many of the scientists here don’t go<br />

on to academic careers. Those that go into industry have a good<br />

launch pad through campus firms.<br />

The ECRC may become a good platform for the integration of companies.<br />

I am a member of the council and know the concept. At the<br />

moment the project hasn’t yet evolved to the point that it would be<br />

really simple for companies to dock on. But as things progress, if the<br />

ECRC is open, and they are prepared to sit down with visitors and talk,<br />

it could function that way. Another point of contact will be core facilities,<br />

if they have extra capacity that can be used by companies.<br />

When a company expresses interest in coming, how do you<br />

make the campus attractive to them?<br />

There are the scientific strengths of the <strong>MDC</strong> in cardiovascular and<br />

metabolic diseases, cancer, and neurodegenerative diseases. That’s<br />

complemented very well by the structural and pharmacological work<br />

of the FMP. The presence of HELIOS with the possibilities of clinical<br />

testing. We also have the Academy of Health here, which offers a very<br />

extensive training program. About a thousand students are enrolled<br />

there now, being trained for a broad spectrum of health-related professions.<br />

It’s an association which carries out training as a service for<br />

hospitals, rehabilitation centers, and similar institutions. The students<br />

will become lab technicians, massage therapists, people who work in<br />

retirement homes, and so on. The Academy operates in an association<br />

including clinics, who co-finance its activities.<br />

We also show them the potential of Buch – there are other attractive<br />

campuses in the village, such as the area where the Franz Volhard<br />

Clinic used to be – that are now empty. We hope very much that they<br />

will be developed as partners in this complex of activities related to<br />

health.<br />

Buch does have a few disadvantages. The town isn’t in the<br />

city, which doesn’t always make things easy...<br />

During the first two phases of development on the campus, as the<br />

FMP was moving in, as everything was being renovated and built<br />

anew, we were worried that the discrepancy between a very progressive<br />

scientific campus and a village from the former GDR would become<br />

wider. We were getting more and more questions like, “Where can<br />

we go shopping here? What are the schools like?” For ten years, with<br />

the exception of this campus, Buch had been experiencing an exodus.<br />

Clinics were moving out; people were leaving, and there were empty<br />

houses. Then when HELIOS bought the state clinic in 2001, things suddenly<br />

started to move ahead. Several of us got together and decided to<br />

found a “Buch management association.”<br />

We convinced the region to do something similar; they created a regional<br />

management association. Their main responsibility was to identify<br />

the interests of the various partners and then to speak to the Senate<br />

with a unified voice about what has to be done for the region. One of<br />

their functions was marketing – to show outsiders the positive<br />

aspects of Buch. They have now received sponsoring for about six<br />

years and have done a lot for the town. These haven’t always been<br />

things that directly affected the campus, but they have been important<br />

for the infrastructure of the community and that has had important<br />

indirect effects. There have been historical walks through the<br />

Hoffmann campuses, business establishment workshops, and public<br />

meetings. We have met with potential investors and shown them the<br />

empty campuses and discussed possible uses. Some of these areas<br />

have now been bought and will be developed in a variety of ways –<br />

many of them will be related to the region’s overall health mission.<br />

One of them will house the new Life Sciences Center. I think all of<br />

these things together have helped Buch make a big step forward – and<br />

that in turn helps the campus.<br />

You’ll be retiring at about the time this book appears...<br />

Yes! Andreas Mätzold and Ulrich Scheller will take over my functions in<br />

the BBB. But I won’t completely disappear. One project I’ll be involved<br />

in is working on the Life Sciences Center – something I’m really pleased<br />

to be involved in, because this project has been an immense<br />

amount of work for almost eight years now. We have the other things<br />

we need: an excellent scientific campus, an enormous new hospital.<br />

While all the other projects that are being planned are important –<br />

new living areas and rehabilitation centers, housing for the families of<br />

patients – we also need something special. The Life Sciences Center<br />

has the potential of bringing a huge new public to Buch. As well as<br />

being something good just in itself, that will also give a strong boost<br />

to our local infrastructure. We think it will make an ideal new meeting<br />

point for all of the players here – from physicians and scientists to<br />

industry, children, and the general public.<br />

299 Interview


300<br />

Index<br />

Adamidi, Catherine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .271<br />

Agre, Peter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .186<br />

Alzheimer, Alois . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .169<br />

Andersen, Olav . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .170<br />

Anzenberger, Uwe . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .164<br />

Bähring, Silvia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .222<br />

Bairach, Rich . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .168<br />

Baumeister, Hans . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .132<br />

Behrens, Jürgen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .34<br />

Bendzko, Peter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .295<br />

Berkenhout, John . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .274<br />

Beyreuther, Konrad . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .169<br />

Bielka, Heinz . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .14<br />

Bienert, Michael . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .197, 288<br />

Bilginturan, Nihat . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .222<br />

Bilroth, Theodor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .72<br />

Birchmeier, Carmen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .28, 62, 154, 284<br />

Birchmeier, Walter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .22, 94, 168, 213, 284<br />

Blankenstein, Thomas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .119<br />

Blanz, Judith . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .206<br />

Bordag, Natalie . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .198<br />

Bright, Richard . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .220<br />

Broemer, Meike . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .243<br />

Burkitt, Dennis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .106<br />

Burmeister, Regina . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .166<br />

Cage, John . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .78<br />

Cajal, Santiago Ramón y . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .53<br />

Caliess, Christiane . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .44<br />

Carrel, Alexis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .90<br />

Chevelkov, Veniamin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .176<br />

Chitayat, David . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .230<br />

Chmielowiec, Jolanta . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .31<br />

Christély, Thomas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .79<br />

Ciechanover, Aaron . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .168<br />

Clevers, Hans . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .36<br />

Correns, Carl . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .13<br />

Couper, Archibald . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .11<br />

Crick, Francis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .30<br />

Danielczyk, Antje . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .136<br />

Darwin, Charles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .10, 248<br />

Dathe, Margitta . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .197<br />

de Vries, Hugo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .13<br />

De Vries, Hugo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .223<br />

Deichman, Galina . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .124<br />

Delbrück, Max . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .16<br />

Deter, Auguste . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .169<br />

Diekhoff, Britta . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .82<br />

Dörken, Bernd . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .243, 286, 293<br />

Eckert, Andreas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .134<br />

Edemir, Bayram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .189<br />

Eisai, Myoan . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .250<br />

Elsener, Karl . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .161<br />

Endruschat, Jens . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .82<br />

Ertl, Gerhard . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .51<br />

Erzgräber, Gudrun . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .167, 296<br />

Fichtner, Iduna . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .295<br />

Fischer, Judith . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .276<br />

Förster, Reinhard . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .66<br />

Frahm, Christina . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .155<br />

Franke, Werner . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .95<br />

Friedländer, Marc . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .267<br />

Furth, Priscilla . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .72<br />

Ganten, Detlev . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .281, 292<br />

Garratt, Alistair . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .154<br />

Gassman, Max . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .155<br />

Gerull, Brenda . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .92<br />

Giese, Klaus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .81<br />

Glass, Rainer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .59<br />

Göhler, Heike . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .259<br />

Goletz, Steffen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .134<br />

Golgi, Camille . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .53<br />

Gollasch, Maik . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .232<br />

Gomoll, Michael . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .190<br />

Gong, Maolian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .234<br />

Grossmann, Katja . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .95<br />

Grunz, Katharina . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .279<br />

Guise, Kevin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .142<br />

Hackett, Perry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .142<br />

Haeckel, Ernst . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .12<br />

Hammes, Annette . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .164<br />

Harris, Robert . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .184<br />

Helmholtz, Hermann von . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .10<br />

Henn, Volker . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .189<br />

Hertwig, Oskar . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .12<br />

Hinz, Michael . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .243<br />

Hoffmann, Ludwig . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .6<br />

Hollman, Andreas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .6<br />

Höpken, Uta . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .109<br />

Hübner, Norbert . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .65, 233, 274<br />

Huelsken, Jörg . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .35<br />

Humboldt, Alexander von . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .9<br />

Hundrucker, Christian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .189<br />

Ivics, Zoltán . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .141<br />

Izsvák, Zsuzsanna . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .141<br />

Jeffries, Alec . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .224<br />

Jens, Walter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .295<br />

Jentsch, Thomas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .86, 201<br />

Joshi, Mangesh . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .176, 213<br />

Kapitonov, Vladimir . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .146<br />

Kaufmann, Jörg . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .81<br />

Index


301 Index<br />

Kekulé, Auguste . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .11<br />

Keller, Sandro . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .194<br />

Kemmner, Wolfgang . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .72<br />

Kempermann, Gerd . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .58<br />

Kettenmann, Helmut . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .14, 53, 86<br />

Kettritz, Ralph . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .112<br />

Klaus, Alexandra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .39<br />

Klein, Eva . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .121<br />

Klein, Georg . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .121<br />

Klippel, Anke . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .81<br />

Klußmann, Enno . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .185<br />

Knut . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4<br />

Koch, Robert . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .11<br />

Koralov, Sergie . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .269<br />

Kries, Jens Peter von . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .213<br />

Kronenberg, Golo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .59<br />

Kujawa-Schmeitzner, Elisabeth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .24<br />

Leheste, Jörg-Robert . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .166<br />

Lenhard, Diana . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .66<br />

Leutz, Achim . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .40, 293<br />

Lewin, Gary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .31, 151, 284<br />

Lindbergh, Charles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .89<br />

Linser, Rasmus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .176<br />

Lipp, Martin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .66, 86, 103<br />

Luft, Friedrich . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .106, 112, 218, 281, 286, 293<br />

Luft, Friedrich . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .112<br />

Luft, Friedrich . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .218<br />

Luft, Friedrich . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .281<br />

Luft, Friedrich . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .286<br />

Luft, Friedrich . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .293<br />

Maatz, Henricke . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .276<br />

Mammen, Jeanne . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .16, 78<br />

Markovic, Darko . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .56<br />

Mätzold, Andreas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .299<br />

Maul, Björn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .87<br />

Melby, James . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .229<br />

Mendel, Gregor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .12<br />

Milenkovic, Nevena . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .155<br />

Miskey, Csaba . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .144<br />

Mo, Xianming . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .48<br />

Monti, Jan . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .281<br />

Morgan, Thomas H. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .223<br />

Muljo, Stefan . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .269<br />

Muller, Hermann . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .30, 223<br />

Müller, Johann . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .10<br />

Naegoe, Ioana . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .209<br />

Nägeli, Karl von . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .13, 61<br />

Naraghi, Ramin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .228<br />

Narayanan, Saravanakumar . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .180<br />

Nedvetsky, Pavel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .191<br />

Neeser, Philippe . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .250<br />

Nykjaer, Anders . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .172<br />

Oschkinat, Hartmut . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .210<br />

Pérec, Georges . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .18<br />

Perutz, Max . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .256<br />

Petersen, Helle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .165<br />

Podust, Larissa . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .215<br />

Porrasmillan, Pablo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .254<br />

Rajewsky, Klaus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .31, 244<br />

Rajewsky, Nikolaus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .237, 262<br />

Redel, Alexandra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .254<br />

Reich, Jens . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .247<br />

Reif, Bernd . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .174<br />

Rio-Hortegas, Pio del . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .56<br />

Rodin, Auguste . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .211<br />

Rosenthal, Walter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .186, 288<br />

Rossum, Barth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .212<br />

Rous, Peyton . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .46<br />

Scheidereit, Claus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .240, 284, 293<br />

Scheller, Ulrich . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .299<br />

Schlag, Peter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .70, 284, 293<br />

Schleiden, Mathias . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .10<br />

Schliemann, Heinrich . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .40<br />

Schliemann, Sophia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .42<br />

Schmidt, Roland . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .7<br />

Schmidt-Ulrich, Ruth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .248<br />

Schmieder, Peter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .212<br />

Schmitt, Andrea . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .144<br />

Schuster, Herbert . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .222<br />

Schwann, Theodor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .10<br />

Schwartze, Stefan . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .168, 285<br />

Scriabin, Alexander . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .262<br />

Seyfried, Salim . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .166<br />

Siefert, Steffi . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .60<br />

Sinzelle, Ludivine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .146<br />

Spoelgen, Robert . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .164<br />

Stein, Ulrike . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .72<br />

Synowitz, Michael . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .59<br />

Thierfelder, Ludwig . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .88, 284<br />

Timoféeff-Ressovsky, Nikolay . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .30<br />

Toka, Hakan . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .222<br />

Toka, Okan . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .222<br />

Tschernak, Erich von . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .13<br />

Uckert, Wolfgang . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .127<br />

Vargas, Carolyn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .217<br />

Vasyutina, Elena . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .62<br />

Vietinghoff, Sibylle von . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .112<br />

Virchow, Rudolf . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .10, 40<br />

Vogt, Cécile . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .30<br />

Vogt, Oskar . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .30<br />

Volhard, Franz . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .220<br />

von Kölliker, Albrecht . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .54<br />

Waaler, Jo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .217<br />

Wagner, Anne . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .255<br />

Wallace, Alfred . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .10<br />

Walther, Wolfgang . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .72<br />

Wanker, Erich . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .251<br />

Wartosch, Lena . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .204<br />

Watson, James . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .30<br />

Wengner, Antje . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .110<br />

Wiedenmann, Bertram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .85<br />

Wienker, Thomas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .221<br />

Willimsky, Gerald . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .125<br />

Willnow, Thomas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .160, 259, 292<br />

Wolf, Jana . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .247


Further reading<br />

Andersen, OM, Reiche, R, Schmidt, V, Gotthardt, M, Spoelgen, R, Behlke, J, von Arnim, CA F,<br />

Breiderhoff, T, Jansen, P, Wu, X, Bales, KR, Cappai, R, Masters, CL, Gliemann, J, Mufson, EJ,<br />

Hyman, BT, Paul, SM, Nykjær, N and T E Willnow. (2005) SorLA/LR11, a neuronal sorting receptor<br />

that regulates processing of the amyloid precursor protein. Proc. Natl. Acad. Sci. USA<br />

102, 13461-13466.<br />

Atanasoski, S, Scherer, S, Sirkowski, E, Leone, D, Garratt, AN, Birchmeier C, Suter U. (2006).<br />

ErbB2 signaling in Schwann cells is largely dispensable for maintenance of myelinated<br />

peripheral nerves and proliferation of adult Schwann Cells following injury. J. Neurosci. 26:<br />

2124-2131.<br />

Bähring, S, Kann, M, Neuenfeld, Y, Gong, M, Chitayat, D, Toka, HR, Toka, O, Plessis, G, Maass, P,<br />

Rauch, A, Aydin, A, Luft, FC. (2008) The inversion region for hypertension and brachydactyly<br />

on chromosome 12p features multiple splicing and patient-specific expression of non-coding<br />

RNA. Hypertension (in press).<br />

Ball LJ, Kühne R, Schneider-Mergener J, Oschkinat H (2005) Recognition of proline rich<br />

motifs (PRMs) by protein-protein interaction domains. Angew Chem Int Ed 44, 2852-2869.<br />

Bárány-Wallje E, Keller S, Serowy S, Geibel S, Pohl P, Bienert M, Dathe M (2005) A critical<br />

reassessment of penetratin translocation across lipid membranes. Biophys J 89, 2513-2521.<br />

Begay, V, Smink, J, and Leutz, A. (2004). Essential requirement of CCAAT/ enhancer binding<br />

proteins in embryogenesis. Molecular Cell Biol, 24, 9744-9751<br />

Blankenstein, Th (2007). Do autochthonous tumors interfere with effector T cell responses?<br />

Sem. Cancer Biol. 17: 267-274. Weinhold, M, Sommermeyer, D, Uckert, W and Blankenstein,<br />

Th. (2007). Dual T cell receptor expressing CD8+ T cells with tumorand self- specificity can<br />

inhibit tumor growth without causing severe autoimmunity. J. Immunol., in press.<br />

Böddrich, A, Gaumer, S, Haacke, A, Tzvetkov, N, Albrecht, M, Evert, BO, Futschik, M, Müller, EC,<br />

Lurz, R, Breuer, P, Schugardt, N, Plaßmann, S, Morabito, LM, Warrick, JM, Suopanki, J, Wüllner,<br />

U, Frank, R, Hartl, FU, Bonini, NM, Wanker, EE. (2005) An arginine/lysine-rich motif in ataxin-<br />

3 is responsible for the interaction with the molecular chaperone VCP/p97 that modulates<br />

aggregate formation and neurotoxicity. EMBO J. 25, 1547-1558.<br />

Cambridge SB, Geißler D, Keller S, Cürten B (2006) A caged doxycycline analogue for photoactivated<br />

gene expression. Angew Chem Int Ed 45, 2229-2231.<br />

Chaurasia G, Iqbal Y, Haenig C, Herzel H, Wanker EE, Futschik ME. (2007). UniHI: an entry<br />

gate to the human protein interactome. Nucleic Acids Research, 2007, 35, D590-4. Herbst,<br />

M, Wanker, EE. (2007): Small molecule inducers of heat shock response reduce polyQ-mediated<br />

huntingtin aggregation. A possible therapeutic strategy. Neurodegener Dis. 4(2-3),<br />

254-60.<br />

Chen, K, Rajewsky, N. (2006). Natural selection on human microRNA binding sites inferred<br />

from SNP data, Nature Genetics 38, 1452-1456.<br />

Chen, K, Rajewsky, N. (2007). The evolution of gene regulation by transcription factors and<br />

microRNAs. Nature Reviews Genetics 8, 93-103.<br />

Chmielowiec, J, Borowiak, M, Morkel, M, Stradal, T, Munz, B, Werner, S, Wehland, J, Birchmeier,<br />

C, Birchmeier, W. (2007). c-Met is essential for wound healing in the skin. J. Cell Biol. 177,<br />

151-162.<br />

Ehrnhoefer, DE, Duennwald, M, Markovic, P, Wacker, JL, Engemann, S, Roark, M, Legleiter, J,<br />

Marsh, JL, Thompson, LM, Lindquist, S, Muchowski, PJ and Wanker, EE. (2006). Green tea (-)epigallocatechin-gallate<br />

modulates early events in huntingtin- misfolding and reduces<br />

toxicity in Huntington’s disease models. Hum Mol Genet. 15(18), 2743-2751.<br />

Engels, B, Nößner, E, Frankenberger, B, Blankenstein, T, Schendel, D, Uckert, W. (2005).<br />

Redirecting human T lymphocytes towards renal cell carcinoma-specificity by retroviral<br />

transfer of T cell receptor genes. Hum. Gene Ther. 16, 799-810.<br />

Further reading<br />

Engels, B, Uckert, W. (2007). Redirecting T lymphocyte specificity by T cell receptor gene<br />

transfer – A new era for immunotherapy. Mol. Aspects Med. 28, 115-142.<br />

Gerull , B, Heuser, A, Wichter, T, Paul, M, Basson, CT, McDermott, DA, Lerman, BB, Markowitz,<br />

SM, Ellinor, PT, MacRae, CA, Peters, S, Grossmann, KS, Drenckhahn, J, Michely, B, Sasse-<br />

Klaassen, S, Birchmeier, W, Dietz, R, Breithardt, G, Schulze- Bahr, E, Thierfelder L. (2004).<br />

Mutations in the desmosomal protein plakophilin-2 are common in arrhythmogenic right<br />

ventricular cardiomyopathy. Nat Genet. 36, 1162-4<br />

Glass R, Synowitz M, Kronenberg G, Markovic D, Wang LP, Gast D, Kiwit J, Kempermann G<br />

and Kettenmann, H. (2005). Gliomainduced attraction of endogenous neural precursor<br />

cells is associated with improved survival, J. Neurosci., 25, 2637-2646.<br />

Grego-Bessa, J, Luna-Zurita, L, del Monte, G, Bolós, V, Melgar, P, Arandilla, A, Garratt, AN, Zang,<br />

H, Mukouyama, Y, Chen, H, Shou, W, Ballestar, E, Esteller, M, Rojas, A, Pérez-Pomares, de la<br />

Pompa, JL. (2007). Notch signalling is essential for cardiac ventricular chamber development.<br />

Developmental Cell 12: 415-429.<br />

Haas B, Schipke C G, Peters O, Söhl G, Willecke K. and Kettenmann H. (2006) Activitydependent<br />

ATP-waves in the Mouse Neocortex are Independent from Astrocytic Calcium<br />

Waves. Cereb Cortex, 16, 237-46.<br />

Hammes, A, Andreassen, TK, Spoelgen, R, Raila, J, Hubner, N, Schulz, H, Metzger, J,<br />

Schweigert, FJ, Luppa, PB, Nykjaer, A, Willnow, TE. (2005) Role of endocytosis in cellular<br />

uptake of sex steroids. Cell 122(5): 751-762.<br />

Hanisch U-K and Kettenmann K (2007) Microglia – active sensor and versatile effector cells<br />

in the normal and pathologic brain, Nature Neuroscience, 10,1387-1394.<br />

Heidemann A C, Schipke C G, Kettenmann H. (2005) Extracellular application of nicotinic<br />

Acid adenine dinucleotide phosphate induces Ca2+ signaling in astrocytes in situ, J. Biol.<br />

Chem., 280, 35630-35640.<br />

Henke, N, Schmidt-Ullrich, R, Dechend, R, Park, JK, Qadri, F, Wellner, M, Obst, M, Gross, V, Dietz,<br />

R, Luft, FC, Scheidereit, C, and Müller DN. (2007). Vascular endothelial-specific NF-κB suppression<br />

attenuates hypertension-induced renal damage. Circ Res 101, 268-276.<br />

Hinz, M, Broemer, M, Çöl Arslan, S, Dettmer, R, Scheidereit, C. (2007). Signal-responsiveness<br />

of IB kinases is determined by Cdc37-assisted transient interaction with Hsp90. J. Biol.<br />

Chem. in press.<br />

Höpken, UE, Wengner, AM, Loddenkemper, C, Stein, H, Heimesaat, MM, Rehm, A, Lipp, M.<br />

(2007) CCR7 Deficiency causes ectopic lymphoid neogenesis and disturbed mucosal tissue<br />

integrity. Blood 109, 886-95 [Epub 2006 Oct 3]<br />

Hübner, N, Wallace, CA, Zimdahl, H, Petretto, E, Schulz, H, Maciver, F, Muller, M, Hummel, O,<br />

Monti, J, Zidek, V, Musilova, A, Kren, V, Causton, H, Game, L, Born, G, Schmidt, S, Müller, A,<br />

Cook, SA, Kurtz, TW, Whittaker, J, Pravenec, M, Aitman, TJ. (2005). Integrated transcriptional<br />

profiling and linkage analysis for disease gene identification. Nature Genetics. 37, 243-253.<br />

Hübner, N. (2006). Expressing physiology. Nature Genetics 38, 140-141.<br />

Humphries, JD, Schofield, NR, Mostafavi-Pour, Z, Green, LJ, Garratt, AN, Mould, AP,<br />

Humphries, MJ. (2005). Dual functionality of the anti-1 integrin antibody, 12G10, exemplifies<br />

agonistic signalling from the ligand-binding pocket of integrin adhesion receptors. J. Biol.<br />

Chem. 280: 10234-10243.<br />

Hundsrucker C, Krause G, Beyermann M, Prinz A, Zimmermann B, Diekmann O, Lorenz D,<br />

Stefan E, Nedvetsky P, Dathe M, Christian F, McSorley T, Krause E, McConnachie G, Herberg<br />

FW, Scott J, Rosenthal W, Klußmann E (2006) Highaffinity AKAP18-delta-protein kinase A<br />

interaction yields novel protein kinase A anchoring disruptor peptides. Biochem J 396,<br />

297-306.<br />

302


Ivics, Z, Izsvák, Z. (2006). Transposons for gene therapy! Curr. Gene Ther. 6, 593-607.<br />

Ivics, Z, Katzer, A, Stüwe, EE, Fiedler, D, Knespel, S, Izsvák, Z. (2007). Targeted Sleeping Beauty<br />

transposition in human cells. Mol. Ther. 15, 1137-1144.<br />

Janz, M, Hummel, M, Truss, M, Wollert-Wulf, B, Mathas, S, Jöhrens, K, Hagemeier, C, Bommert,<br />

K, Stein, H, Dörken, B, Bargou, RC. (2006). Classical Hodgkin lymphoma is characterized by<br />

high constitutive expression of activating transcription factor 3 (ATF3) which promotes viability<br />

of Hodgkin/Reed- Sternberg cells. Blood. 107, 2536-2539.<br />

Jentsch TJ. (2008) CLC chloride channels and transporters: from genes to protein structure,<br />

pathology and physiology. Crit Rev Biochem Mol Biol. 43(1): 3-36.<br />

Joshi M, Vargas C, Moelling K, Boisguerin P, Diehl A, Schmieder P, Krause G, Hagen V, Schade<br />

M, Oschkinat H (2006) Making protein-protein interactions drugable: discovery and 3D<br />

structure of low-molecular-weight ligands complexed with the AF6 PDZ domain. Angew<br />

Chem Int Ed 45, 3790-3795.<br />

Jundt, F, Raetzel, N, Müller, C, Calkhoven, CF, Kley, K, Mathas, S, Lietz, A, Leutz, A, Dörken, B.<br />

(2005). A rapamycin derivative (everolimus) controls proliferation through down-regulation<br />

of truncated CCAAT enhancer binding protein and NF-κB activity in Hodgkin and<br />

anaplastic large cell lymphomas. Blood. 106, 1801-1807.<br />

Jüttner S, Wissmann C, Jöns T, Vieth M, Hertel J, Gretschel S, Schlag PM, Kemmner W, Höcker<br />

M. (2006) Vascular endothelial growth factor-D and its receptor VEGFR-3: Two novel independent<br />

prognostic markers in gastric adenocarcinoma. J. Clin. Oncol. 24, 228-240.<br />

Kasper D, Planells-Cases R, Fuhrmann JC, Scheel O, Zeitz O, Ruether K, Schmitt A, Poët M,<br />

Steinfeld R, Schweizer M., Kornak U, Jentsch TJ (2005) Loss of the chloride channel ClC-7<br />

leads to lysosomal storage disease and neurodegeneration. EMBO J. 24, 1079-1091.<br />

Kaufman, CD, Izsvák, Z, Katzer, A, Ivics, Z. (2005). Frog Prince transposon-based RNAi vectors<br />

mediate efficient gene knockdown in human cells. Journal of RNAi and Gene Silencing 1,<br />

97-104.<br />

Keller S, Heerklotz H, Blume A (2006) Monitoring lipid membrane translocation of sodium<br />

dodecyl sulfate by isothermal tritation calorimetry. J Am Chem Soc 128, 1279-1286.<br />

Keller S, Heerklotz H, Jahnke N, Blume A (2006) Thermodynamics of lipid membrane solubilization<br />

by sodium dodecyl sulfate. Biophys J 90, 4509-4521.<br />

Keller S, Sauer I, Strauss H, Gast K, Dathe M, Bienert M (2005) Membrane-mimetic nanocarriers<br />

formed by a dipalmitoylated cell-penetrating peptide. Angew Chem Int Ed 44, 5252-<br />

5255.<br />

Kharkovets T, Dedek K, Maier H, Schweizer M, Khimich D, Nouvian R, Vardanyan V, Leuwer R,<br />

Moser T, Jentsch TJ (2006) Mice with altered KCNQ4 K+ channels implicate sensory outer<br />

hair cells in human progressive deafness. EMBO J 25, 642-652.<br />

Klaus A, Birchmeier W. (2008) Wnt signalling and its impact on development and cancer.<br />

Nat Rev Cancer, 8(5), 387-98.<br />

Klaus A, Saga Y, Taketo MM, Tzahor E, Birchmeier W. (2007) Distinct roles of Wnt/β-catenin<br />

and Bmp signaling during early cardiogenesis. Proc Natl Acad Sci 104(47), 18531-6. Epub<br />

2007 Nov 13.<br />

Klaus, A, Saga, Y, Taketo, MM, Tzahor, E and Birchmeier, W. (2007). Distinct roles of Wnt/catenin<br />

and Bmp signaling during early cardiogenesis. Proc. Natl. Acad. Sci. USA. 104, 18531-<br />

18536.<br />

Krek, A, Gruen, D, Poy, MN, Wolf R, Rosenberg L, Epstein EJ, MacMenamin P, da Piedade I,<br />

Gunsalus KC, Stoffel M, and Rajewsky N (2005). Combinatorial microRNA target predictions.<br />

Nature Genetics 37, 495-500.<br />

Krutzfeldt J, Rajewsky N, Braich R, Rajeev KG, Tuschl T, Manoharan M, Stoffel M. (2005).<br />

Silencing of microRNAs in vivo with ‘antagomirs’. Nature 438, 685-689.<br />

Lall, S, Grun, D, Krek A, Chen K, Wang Y-L, Dewey DN, Sood P, Colombo T, Bray N,<br />

MacMenamin P, Kao H-L, Gunsalus KC, Pachter L, Piano F, Rajewsky, N. (2006). A genome-<br />

Wide Map of Conserved MicroRNA Targets in C. elegans. Current Biology 16, 460-71.<br />

Lange PF, Wartosch L, Jentsch TJ, Fuhrmann JC (2006) ClC-7 requires Ostm1 as a beta-subunit<br />

to support bone resorption and lysosomal function. Nature 440, 220-223.<br />

Lee-Kirsch MA, Gong M, Chowdhury D, Senenko L, Engel K, Lee YA, de Silva U, Bailey SL, Witte<br />

T, Vyse TJ, Kere J, Pfeiffer C, Harvey S, Wong A, Koskenmies S, Hummel O, Rohde K, Schmidt<br />

RE, Dominiczak AF, Gahr M, Hollis T, Perrino FW, Lieberman J, Hübner N. (2007). Mutations<br />

in the gene encoding the 3’-5’ DNA exonuclease TREX1 are associated with systemic lupus<br />

erythematosus. Nature Genetics 39, 1065-7.<br />

Lietz, A, Janz, M, Sigvardsson, M, Jundt, F, Dörken, B, Mathas S. (2007). Loss of HLH transcription<br />

factor E2A activity in primary effusion lymphoma confers resistance to apoptosis. Br J<br />

Haematol. 137, 342-348.<br />

López-Bendito, G, Cautinat, A, Sánchez, JA, Bielle, F, Flames, N, Garratt, AN, Tagmale, D, Role,<br />

LW, Charnay, P, Marín, O, Garel, S. (2006). Tangential migration controls axon guidance: a<br />

role for Neuregulin-1 in thalamocortical axon navigation. Cell 125, 127-142.<br />

Martinez-Salgado C, Benckendorff AG, Chiang LY, Wang R, Milenkovic N, Wetzel C, Hu J,<br />

Stucky CL, Parra MG, Mohandas N, Lewin GR (2007) Stomatin and sensory neuron mechanotransduction.<br />

J Neurophys 98, 3802-8.<br />

Mathas, S, Janz, M, Hummel, F, Hummel, M, Wollert-Wulf, B, Lusatis, S, Anagnostopoulos, I,<br />

Lietz, A, Sigvardsson, M, Jundt, F, Jöhrens, K, Bommert, K, Stein, H, Dörken, B. (2006). Intrinsic<br />

inhibition of E2A by overexpressed ABF-1 and Id2 is involved in reprogramming of the neoplastic<br />

B cells in classical Hodgkin lymphoma. Nature Immunol. 7, 207-215.<br />

Mathas, S, Jöhrens, K, Joos, S, Lietz, A, Hummel, F, Janz, M, Jundt, F, Anagnostopoulos, I,<br />

Bommert, K, Lichter, P, Stein, H, Scheidereit, C, Dörken, B. (2005). Elevated NF-{kappa}B p50<br />

complex formation and Bcl-3 expression in classical Hodgkin, anaplastic large cell, and<br />

other peripheral T cell lymphomas. Blood. 106, 4287-4293.<br />

Mehrhof, FB, Schmidt-Ullrich, R, Dietz, R, Scheidereit, C. (2005). Regulation of Vascular<br />

Smooth Muscle Cell Proliferation: Role of NF-κB Revisited. Circ Res 96, 958-964<br />

Milenkovic N, Frahm C, Gassmann M, Griffel C, Erdmann B, Birchmeier C, Lewin GR, Garratt<br />

AN (2007) Nociceptive tuning by Stem Cell Factor/c-Kit signaling. Neuron 56: 893-906.<br />

Miskey, C, Papp, B, Mátés, L, Sinzelle, L, Keller, H, Izsvák, Z, Ivics, Z. (2007). The ancient mariner<br />

sails again: Transposition of the human Hsmar1 element by a reconstructed transposase<br />

and activities of the SETMAR protein on transposon ends. Mol. Cell. Biol. 27, 4589-600.<br />

Mo X., Kowenz-Leutz E, Xu H, Leutz A. (2004) Ras induces mediator complex exchange on<br />

C/EBP. Molecular Cell, 13: 241-250<br />

Mo, X, Kowenz-Leutz, E, Laumonnier, Y, Xu, H, and Leutz, A (2005) Histone H3-tail positioning<br />

and acetylation by c-Myb but not the v-Myb DNA binding SANT Domain. Genes &<br />

Development, 19: 2447-2457<br />

Moeller, H, Jenny, A, Schaeffer, H-J, Schwarz-Romond, T, Mlodzik, M, Hammerschmidt, M, and<br />

Birchmeier, W. (2006): Diversin regulates heart formation and gastrulation movements in<br />

development. Proc. Natl. Acad. Sci. USA 103, 15900-15905.<br />

Narayanan S, Kamps B, Boelens W, Reif B (2006) B-crystalline competes with Alzheimer´s<br />

disease beta-amyloid peptide for peptide-peptide interactions and induces oxidation of<br />

Abeta- Met35. FEBS Lett 580, 5941-5946.<br />

Narayanan S, Reif B (2005) Characterization of chemical exchange between soluble and<br />

aggregated states of β-amyloid by solution state NMR upon variation of the salt conditions.<br />

Biochemistry 44, 1444-1452.<br />

Nedvetsky P, Stefan E, Frische S, Santamaria K, Wiesner B, Valenti G, Hammer III JA, Nielsen<br />

S, Goldenring RJ, Rosenthal W, Klußmann E (2007) A role of myosin Vb and Rab11-FIP2 in the<br />

aquaporin-2 shuttle. Traffic 8, 110-123.<br />

Oeckinghaus, A, Wegener, E, Welteke, V, Ferch, U, Çöl Arslan, S, Ruland, J, Scheidereit, C,<br />

Krappmann, D. (2007). TRAF6 mediated Malt1 ubiquitination triggers IKK/NF-κB signalling<br />

upon T cell activation. EMBO J., in press<br />

Park TJ, Lu Y, Jüttner R, Smith ES, Hu J, Brand A, Wetzel C, Milenkovic N, Erdmann B,<br />

Heppenstall PA, Laurito CE, Wilson SP, Lewin GR. (2008) Selective inflammatory pain insensitivity<br />

in the African naked mole-rat (Heterocephalus glaber). PLoS Biol. 6: e13.<br />

Pocock, J M and Kettenmann H. (2007) Neurotransmitter receptors on microglia, Trends<br />

Neurosci., 30, 527-535.<br />

Poët M, Kornak U, Schweizer M, Zdebik AA, Scheel O, Hoelter S, Wurst W, Schmitt A,<br />

Fuhrmann JC, Planells-Cases R, Mole SE, Hubner CA, Jentsch TJ (2006) Lysosomal storage<br />

disease upon disruption of the neuronal chloride transport protein ClC-6. Proc Natl Acad<br />

Sci USA 103, 13854-13859.<br />

303 Further reading


Poet, M, Kornak, U, Schweizer, M, Zdebik, AA, Scheel, O, Hoelter, S, Wurst, W, Schmitt, A,<br />

Fuhrmann, JC, Planells-Cases, R, Mole, S, E, Hübner, CA, Jentsch TJ. (2006). Lysosomal storage<br />

disease upon disruption of the neuronal chloride transport protein ClC-6. Proc. Natl. Acad.<br />

Sci. USA 103, 13854-13859.<br />

Qi, M-L, Tagawa, K, Enokido, Y, Yoshimura, N, Wada, Y-ichi, Watase, K, Ishiura, S-ichi, Kanazawa,<br />

I, Botas, J, Saitoe, M, Wanker, EE and Okazawa, H. (2007). Proteome analysis of soluble nuclear<br />

proteins reveals that HMGB 1/2 suppress genotoxic stress in polyglutamine diseases.<br />

Nature Cell Biology 9(4), 402-414.<br />

Rasheed, AU, Rahn, H-P, Sallusto, F, Lipp, M, and Muller, G (2006) B helper T cell activity is confined<br />

to CXCR5hiICOShi CD4 T cells and independent of CD57 expression. Eur. J. Immunol.,<br />

36, 1892-1903<br />

Rat Genome Sequencing Project Consortium. (2004). Genome Sequence of the Brown<br />

Norway rat yields insights into mammalian evolution. Nature. 428, 493-521.<br />

Rogaeva, E, et al. (2007) The neuronal sortilin-related receptor SORL1 is genetically associated<br />

with Alzheimer’s Disease. Nat Genet 39, 168-177.<br />

Rust MB, Alper SL, Rudhard Y, Shmukler BE, Vicente R, Brugnara C, Trudel M, Jentsch TJ,<br />

Hubner CA (2007) Disruption of erythroid KCl-cotransporters alters erythrocyte volume and<br />

partially rescues erythrocyte dehydration in SAD mice. J Clin Invest 117, 1708-1717.<br />

Sasse-Klaassen, S, Probst, S, Gerull, B, Oechslin, E, Nurnberg, P, Heuser, A, Jenni, R, Hennies, HC,<br />

Thierfelder, L. (2004). Novel gene locus for autosomal dominant left ventricular noncompaction<br />

maps to chromosome 11p15. Circulation. 109, 2720-2723<br />

Schaeper, U, Vogel, R, Chmielowiec, J, Huelsken, J, Rosário, M, and Birchmeier, W. (2007).<br />

Distinct requirements for Gab1 in Met and EGF receptor signaling in vivo. Proc. Natl. Acad.<br />

Sci. USA, 104, 15376-15381.<br />

Scheel O, Zdebik AA, Lourdel S, Jentsch TJ (2005) Voltagedependent electrogenic chlorideproton<br />

exchange by endosomal CLC proteins. Nature 436, 424-427.<br />

Scheel, O, Zdebik, AA, Lourdel, S, Jentsch, TJ. (2005). Voltage dependent electrogenic chlorideproton<br />

exchange by endosomal CLC proteins. Nature 436, 424-427.<br />

Schmidt-Ullrich, R, Tobin, DJ, Lenhard, D, Schneider, P, Paus, R, Scheidereit, C. (2006). NF-B<br />

transmits Eda A1/EdaR signalling to activate Shh and cyclin D1 expression, and controls<br />

post-initiation hair placode down-growth. Development 133, 1045-1057.<br />

Schneider. MA, Meingassner, JG, Lipp, M, Moore, HD. and Rot, A (2007) CCR7 is required for<br />

the in vivo function of CD4+ CD25+ regulatory T cells. J. Exp. Med., 204, 735-745.<br />

Sieber MA, Storm R, Martinez-de-la-Torre M, Muller T, Wende H, Reuter K, Vasyutina E,<br />

Birchmeier C. Lbx1 acts as a selector gene in the fate determination of somatosensory and<br />

viscerosensory relay neurons in the hindbrain. J Neurosci. 2007 May 2;27(18):4902-9.<br />

Singh RK, Kefala G, Janowski R, Mueller-Dieckmann C, von Kries JP, Weiss MS (2005) The<br />

high-resolution structure of LeuB (Rv2995c) from Mycobacterium tuberculosis. J Mol Biol<br />

346, 1-11.<br />

Sommermeyer, D, Neudorfer, J, Weinhold, M, Leisegang, M, Charo, J, Engels, B, Nößner, E.<br />

Heemskerk, M, Schendel, DJ, Blankenstein, T, Bernhard, H, Uckert, W. (2006). Designer T cells<br />

by T cell receptor replacement. Eur. J. Immunol. 36, 3052-3059.<br />

Sommermeyer, D, Neudorfer, J, Weinhold, M, Leisegang, M, Engels, B, Noessner, E, Heemskerk,<br />

M, Charo, J, Schendel, D, Blankenstein, Th, Bernhard, H and Uckert, W. (2006). Designer T cells<br />

by T cell receptor replacement. Eur. J. Immunol. 36, 3052- 3059.<br />

Stefan E, Wiesner B, Baillie GS, Mollajew R, Henn V, Lorenz D, Furkert J, Santamaria K,<br />

Nedvetsky P, Hundsrucker C, Beyermann M, Krause E, Pohl P, Gall I, Maclntyre AN, Bachmann<br />

S, Houslay MD, Rosenthal W, Klußmann E (2007) Compartmentalization of cAMP-dependet<br />

signalling by phosphodiesterase- 4D is involved in the regulation of vasopressinmediated<br />

water reabsorption in renal principal cells. J Am Soc Nephrol 18, 199-212.<br />

Stein U, Arlt F, Walther W, Smith J, Waldman T, Harris ED, Mertins SD, Heizmann CW, Allard<br />

D, Birchmeier W, Schlag PM, Shoemaker RH. (2006) The metastasis-associated gene S100A4<br />

is a novel target of b-catenin/T-cell factor signalling in colon cancer. Gastroenterology 131,<br />

1486-1500.<br />

Stein U, Walther W, Stege A, Kaszubiak A, Fichtner I, Lage H. (2007) Complete in vivo reversal<br />

of the multidrug resistance (MDR) phenotype by jet-injection of anti-MDR1 short hairpin<br />

RNA-encoding plasmid DNA. Mol. Ther. (Epub ahead of print).<br />

Further reading<br />

Synowitz M, Glass R, Faerber K, Markovic D, Kronenberg G, Herrmann K, Schnermann J, Nolte<br />

C, van Rooijen N, Kiwit J, and Kettenmann H. (2006) A1 Adenosine Receptors in Microglia<br />

Control Glioblastoma-Host Interaction. Cancer Res 66, 8550-8557.<br />

Thirunarayanan N, Cifire F, Fichtner I, Posner S, Benga J, Reiterer P, Kremmer E, Kölble K, Lipp<br />

M. (2007) Enhanced tumorigenicity of fibroblasts transformed with human herpesvirus 8<br />

chemokine receptor vGPCR by successive passage in nude and immunocompetent mice<br />

Oncogene. 14, 523-532 [Epub 2007 Mar 23]<br />

Vasyutina E, Lenhard DC, Wende H, Erdmann B, Epstein JA, Birchmeier C. RBP-J (Rbpsuh) is<br />

essential to maintain muscle progenitor cells and to generate satellite cells. Proc Natl Acad<br />

Sci U S A. 2007 Mar 13;104(11):4443-8.<br />

von Vietinghoff, S, Busjahn, A, Schonemann, C, Massenkeil, G, Otto, B, Luft, FC, Kettritz, R.<br />

(2006) Major histocompatibility complex HLA region largely explains the genetic variance<br />

exercised on neutrophil membrane proteinase 3 expression. J Am Soc Nephrol. 17, 3185-3191.<br />

von Vietinghoff, S, Tunnemann, G, Eulenberg, C, Wellner, M, Cardoso, MC, Luft, FC, Kettritz, R.<br />

(2007) NB1 mediates surface expression of the ANCA antigen proteinase 3 on human neutrophils.<br />

Blood. 109, 4487-4493.<br />

Walisko, O, Izsvák, Z, Szabó, K, Kaufman, CD, Herold, S, Ivics, Z. (2006). Sleeping Beauty transposase<br />

modulates cell-cycle progression through interaction with Miz-1. Proc. Natl. Acad.<br />

Sci. USA 103, 4062-4067.<br />

Walisko, O, Izsvák, Z, Szabó, K, Kaufman, CD, Herold, S, Ivics, Z. (2006). Sleeping Beauty transposase<br />

modulates cell-cycle progression through interaction with Miz-1. Proc. Natl. Acad.<br />

Sci. USA 103, 4062-4067.<br />

Walther W, Arlt F, Fichtner I, Aumann J, Stein U, Schlag PM. (2007). Heat-inducible in vivo<br />

gene therapy of colon carcinoma by human mdr1 promoter-regulated tumor necrosis factor-alpha<br />

expression. Mol. Cancer Ther. 6, 236-243.<br />

Walther W, Minow T, Martin R, Fichtner I, Schlag PM, Stein U. (2006) Uptake, biodistribution<br />

and time course of naked plasmid- DNA trafficking after intratumoral in vivo jet-injection.<br />

Hum. Gene Ther.17: 611-624.<br />

Wengner, AM, Höpken, UE, Petrow, PK, Hartmann, S, Schurigt, U, Bräuer, R, and Lipp, M. (2007)<br />

CXCR5- and CCR7-dependent lymphoid neo-genesis in a murine model of chronic antigeninduced<br />

arthritis. Arthritis & Rheum. 56, 3271-3283<br />

Wetzel C, Hu J, Riethmacher D, Benckendorff A, Harder L, Eilers A, Moshourab R, Kozlenkov A,<br />

Labuz D, Caspani O, Erdmann B, Machelska H, Heppenstall PA, Lewin GR (2007) A stomatindomain<br />

protein essential for touch sensation in the mouse. Nature 445:206-209.<br />

Wildner H, Muller T, Cho SH, Brohl D, Cepko CL, Guillemot F, Birchmeier C. dILA neurons in the<br />

dorsal spinal cord are the product of terminal and non-terminal asymmetric progenitor cell<br />

divisions, and require Mash1 for their development. Development. 2006 Jun;133(11):2105-13.<br />

Willem, M, Garratt, AN, Novak, B, Citron, M, Kaufmann, S, Rittger, A, Saftig, P, De Strooper, B,<br />

Birchmeier, C, Haass, C. (2006). Control of peripheral nerve myelination by the gamma-secretase<br />

BACE1. Science 314: 664-666.<br />

Willimsky, G and Blankenstein, Th. (2005) Sporadic immunogenic tumours avoid destruction<br />

by inducing T-cell tolerance. Nature 437,141-146.<br />

Willimsky, G and Blankenstein, Th. (2007). The adaptive immune response to sporadic cancer.<br />

Immunol. Rev., in press.<br />

Willnow, TE, Hammes, A, Eaton, S. (2007) Lipoproteins and their receptors in embryonic<br />

development – more than cholesterol clearance. Development, in press<br />

Zimdahl, H, Nyakatura, G, Brandt, P, Schulz, H, Hummel, O, Fartmann, B, Brett, D, Droege, M,<br />

Monti, J, Lee, YA, Sun, Y, Zhao, S, Winter, E, Ponting, C, Chen, Y, Kasprzyk, A, Birney, E, Ganten, D,<br />

Hübner, N. (2004). A SNP map of the rat genome generated from transcribed sequences.<br />

Science. 303, 807.<br />

304


Credits<br />

Written by Russ Hodge<br />

Photography by Maj Britt Hansen<br />

Graphic design by Nicola Graf<br />

Paintings by Jeanne Mammen, from the collection on the campus in Berlin-Buch<br />

Additional photography by<br />

Silvia Bähring and Hakan Toka pages 226-234, 236, 237, 241<br />

Peter Himsel pages 289, 290, 297<br />

David Aussenhofer page 119<br />

Catherine Adamidi pages 268-269<br />

Ewan St. J. Smith page 153 (naked mole rat)<br />

Russ Hodge pages 75, 95, 100, 103 (Helios), 137<br />

Sharon Hodge<br />

Lindbergh Picture Collection,<br />

Manuscripts and Archives,<br />

page 6 (Knut)<br />

Yale University Library, images on<br />

Dr. Tobias Engelsing and Carolin<br />

Schulz, the Museum of Konstanz,<br />

provided the image of the unique<br />

page 90<br />

Leiner Collection in Konstanz page 59<br />

Robert Schröter page 126 (Berlin Wall)<br />

For more information about the campus:<br />

www.mdc-berlin.de<br />

www.fmp-berlin.de<br />

www.charite.de<br />

www.bbb-berlin.de<br />

Exposure and printing by ColorDruck GmbH, Leimen<br />

www.colordruck.com<br />

305 credits

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!