Chasing Memory

Los Angeles Times
Sunday, August 19, 2007

Chasing Memory / First of four parts
One man’s epic quest for understanding
* What happens when we encounter a new experience that enables us to recall it later at will? And what goes wrong

By Terry McDermott, Times Staff Writer

The first time I spoke with the neuroscientist Gary Lynch, the conversation went something like this:
Me:
I’m interested in spending time in a laboratory like yours, where the
principal focus is the study of memory. I’d like to explain how memory
functions and fails, and why, and use the work in the lab as a means to
illustrate how we know what we know.
Lynch: You’d be welcome to come here. This would actually be a propitious time to be in the lab.
Me: Why’s that?
Lynch: Because we’re about to nail this mother to the door.
Lynch is a neuroscientist at UC Irvine, where he has spent 37 years trying to uncover the biochemical mechanisms of memory.
He
has, for almost the length of his career, been trying to answer
essentially a single pair of questions: What happens in the brain when a
human being encounters a new experience so that he or she can recall it
at will tonight, tomorrow, in 2025? And what goes wrong when we can’t
remember?
This second question has in the last several years taken on great weight.
We
are on the verge of a dementia pandemic. It is estimated that by 2040,
100 million people worldwide will suffer from Alzheimer’s disease,
Huntington’s, Parkinson’s or some other form of dementia. Science has
been able to do precious little to combat these diseases, in large part
because the understanding of the underlying cognitive processes has been
meager. Thousands of scientists have spent countless years seeking and
largely failing to unearth the secrets within the human brain.
Medical
advances have allowed more and more people to live longer but have been
unable to relieve longevity of its principal bane — the breakdown of
mental processes, especially memory. When memory loss occurs, it seldom
fails to impress upon its victims and those who know them the extent to
which our memories constitute our selves.
That breakdowns occur is
not surprising. Consider: You’re 50 years old. What’s your time in the
100-yard dash? How does that compare to the 18-year-old you? Why would
your brain be exempt from declining in analogous ways? It isn’t. So much
goes wrong so often that many malfunctions are considered ordinary and
are often referred to collectively as normal cognitive decline.
Before
that first conversation with Lynch, I already knew that he had been an
often polarizing figure in his field, that he had a reputation for being
pugnacious, and that he had been uncannily right about a lot of things
over a very long time.
In the subsequent two years, I spent a great
deal of time in his lab. I spoke with the other scientists who worked
there and observed their experiments; I read papers they and others
published; I learned how to perform some of the most rudimentary tasks
of their basic experiments. But what I did mostly was talk to Lynch. Or,
more to the point, listen as Lynch explained mammalian biology and
brain science.
Listening to Lynch often entailed following swooping,
exhilarating flights over time and intellectual terrain. Bear with me,
he sometimes said, this might not seem connected to what we’ve been
talking about, but it will circle back. Ten, 20 or 30 minutes later,
often after side trips to ancient Rome or Yankee Stadium or Bismarck’s
Germany, it did.
Lynch almost always spoke in such a way that his
huge ambition, self-regard and lack of pretense were vividly displayed.
He was unreserved, witty, juvenile, insightful and learned in ways that
were surprising. He was as apt to quote Cormac McCarthy as Gregor
Mendel. He made on-the-fly references to, among many other things, left-
handed relief pitchers, Moses, British naval history, the stock market,
Kaiser Wilhelm II, Maxwell’s equations, the ur-city of Ur, Darwin,
Dylan, Kant, Chomsky, Bush, Titian, field theory, drag-racing, his
father’s perpetual habit of calling him — intentionally — by the wrong
name, his career as a gas jockey at an all-night service station,
Pickett’s charge at Gettysburg, Caesar crossing the Rubicon, and the
search for the historical Jesus.
Christine Gall, Lynch’s frequent
collaborator and longtime significant other, said: “Gary just has more
RAM than other people. He can access lateral information that most
people can’t. It isn’t like he has to think and remind himself. It’s
right there. He has access to it. To have that available to inform you,
to make the next cognitive leap — that’s his strength.”
That leaping
ability has earned Lynch as much trouble as reward. He never shies from
proclamation based on his intuitions, nor from criticizing those not
privy to his insight. “That is what amazes me,” he said. “People will
walk in who are very sensible and intelligent biologists and tell you,
‘Memory is this.’ And you go, ‘How in the hell could it possibly be
that? I didn’t think it was that when I was back at Our Lady of Fatima
Grade School. I mean, I didn’t think it was that when I was working at
the all-night gas station. For crying out loud!’ “
One result of this
perhaps excessive straightforwardness has been a constant war with the
neuroscience establishment, with university administrators and
colleagues at Irvine. But whatever his difficulties, Lynch has slogged
along, making hard progress documented in more than 550 published
papers, some of which are considered classic and are among the most
frequently cited works in all of neuroscience.
As a corollary to his
basic research, Lynch has sought ways to counter the various afflictions
that erode the brain’s abilities. Working with chemist Gary Rogers, he
invented a new class of drugs called ampakines, which, if they worked,
would not only improve memory, but would make the brain perform better
in numerous other ways. Drugs of this sort, called cognitive enhancers
or, more simply, smart pills, have been the Holy Grail of brain research
for a century.
Like much contemporary drug research, ampakine
development has been slow going, but by the time I met Lynch, versions
of his drugs were being considered by the Food and Drug Administration
for a series of clinical trials, which should largely determine whether
their substantial promise could be fulfilled. Success would be a signal
moment in neuroscience history.
By chance, the ampakine drug trials
would get underway at the same time the memory research in his lab
seemed headed toward its own finale. Lynch had a sense that answers he
had spent a career chasing were at hand.
He was alternately eager at
the opportunity and despondent at the likelihood of failure. He knew, as
every research scientist does, that almost everything almost always
goes wrong. If, over time, science can be viewed as the steady
extinction of ignorance, in the near term, on most days, ignorance wins
hands down.
“If you’re good, if you’re any good at all, you put
yourself in a situation where reality could come around and — WHACK! —
knock you down. That’s what you really are afraid of. If you don’t have
that, you’re not playing science,” Lynch said.
He was definitely
playing science now. With the drug research and the fast-approaching end
to his torturous journey in what he once characterized as a gulag of
unyielding biology, he had a rare opportunity — a shot on goal, he
called it.
“Come to the lab,” he said. “This could get interesting.”
Lynch Lab
Save for lynch, Lynch Lab was empty just before New Year’s 2005. Much to Lynch’s chagrin, everyone was vacationing.
The
lab had just developed a new technique that he thought would allow
researchers to visualize the physical trace of memories, and in doing so
resolve long-standing, fundamental debates in neuroscience.
This new
technique promised to answer conclusively what had been supposition,
and to answer it in such a way that you would literally see the result.
And people went on vacation?
Most of the space in Lynch Lab was taken
up by two parallel ranks of standard lab benches, complete with
faucets, hoses, beakers, stocks of chemicals, pipettes, scales,
reference books and undergraduates. Lynch and the lab’s senior
scientists had offices on the perimeter, but most of the experimental
work was done out on the benches.
Lynch could almost always be found
in his office, writing or reading, and chewing on a cigar if he had one
or a plastic cafeteria fork if he didn’t. No matter whose name was on
them, almost all of the journal papers that issued from the lab were
written by him.
He has an open, almost guileless face, so helplessly
expressive that your first impulse is to invite him to a poker game. The
years have begun to accumulate, however, cutting deep lines. He is
about 6 feet tall, rail-thin but for the beginnings of a belly, with
tangled, graying hair that has relaxed considerably from its Charlie
Manson heyday. He usually dresses in high-quality, untucked, casual
clothes — Klein, Boss and Zegna shirts and jeans and well-worn chukka
boots.
His corner office is spare and clean — a large glass-top,
metal-frame desk; a dual-monitor Mac workstation; a few potted plants
along broad, undraped windows. He has a telephone on his desk, but it is
often unplugged. He sometimes goes for weeks without reading e-mail.
The only decorations on the walls are a single small plaque honoring him
because his papers were so often cited by other scientists, and a pair
of large abstract paintings of brain interiors, which are mostly purple
and surprisingly pleasant.
Except for a congregation of Starbucks
decaf cups, he is fastidious. There is almost never more than a single
pen and a pad of paper on the desktop; he keeps a spray bottle of glass
cleaner handy to scrub it, which he does religiously. He usually has a
bottle of whiskey and a brace of glasses stowed among the plants. Before
serving, he scrubs the glasses with the same care he applies to the
desk.
The duties of a university scientist leading his own lab are
manifold. Foremost, the existence of the lab depends on his ability to
fund it. He is an employee of the university, but also a profit center.
He must attract grants, from which the university takes a significant
cut, to pay the basic expenses of his laboratory: salaries, equipment,
supplies.
The overwhelming majority of grant money comes from the
federal government, most through the National Institutes of Health. The
competition for money is intense and often leaves normally placid
scientists swearing like deckhands. Lynch, whose lab has been funded
mostly by the federal government at around $1 million per year for
decades, was no exception.
In part because of the constant threat of
extinction, neuroscience labs — even those that don’t have Lynch in
them — are not the happiest places. There is tension and fear and
jealousy and a near-constant sense that careers are about to be made or,
more commonly, missed. Such fraught situations call for careful,
considered management.
Due to a lack of interest, or possibly
ability, which can be the same thing, Lynch seemed to run his lab like a
man on a midnight beer run, running pell-mell down the aisle, throwing
things, many of them unhealthful, into the cart and hoping there would
be enough for everybody when he got back to the house.
Which is to say, although it was obvious to him, it wasn’t always clear to others what Lynch was up to.
In
addition to providing money, the lead scientist, in the academic world
called the principal investigator, is the intellectual leader of the
team. Lynch did very few experiments himself, but designed, assigned or
approved virtually all of what everyone else did. He would hate to admit
it, but he was a dictator.
The ‘free-ride guy’
The youngest son
in a disintegrating Irish Catholic working-class family in Wilmington,
Del., Lynch earned scholarships to Catholic high school, then — “always
a free-ride guy” — the University of Delaware.
The ride ended
abruptly when he was kicked out for partying. He worked odd jobs until
he was readmitted. Because Lynch’s main interest in college was to have a
good time, something had to change. When he came back, he changed
majors from engineering to psychology for, he said, two reasons —
engineering students spent weekends building electric circuits and, as
important, there were very few girls among them. Lynch chose psychology,
he said, because there were plenty of girls and no weekend work.
Lynch,
in spite of his professed laziness, excelled and earned a graduate
scholarship to Princeton, where he quickly determined he was much more
interested in mucking around inside the head than standing outside it
and asking questions.
He earned his doctorate in psychology in 1968,
just three years after enrolling. Soon after, he received a job offer
from UC Irvine. The university was so new it hadn’t yet graduated its
first class.
The offer was to teach in the psychobiology department.
Lynch had not completed a single college course in biology. (Too many
details, he said.) He had never been to California. But one of the first
of those now ubiquitous lists of the best universities had been
recently published. UC Berkeley ranked No. 1 in the world.
“The
thrill I felt was — it’s the people’s university,” said Lynch. “That’s a
public university. Oxford, Cambridge are down here; Harvard’s down
here; Princeton’s down here. The best university in the world is a
public university. I thought, ‘Man, we are so on the right track.’ That
inspired me. . . . I thought, ‘This is it; this is finally it.’ In the
face of people working on great things together in the sunshine, in the
eternal summer of California, privilege falls away. What could be more
beautiful?”
Irvine then was not far removed from its ranch-land past.
There were cattle grazing on the hills above the campus and cowboys
chasing them. Almost overnight, the university became a center of brain
research.
Neuroscience, too, was young, and there was a sense broadly
shared that the human brain, one of the great frontiers of science, was
about to be colonized — although from what direction or by whose army
was unclear. Biologists, chemists, anatomists, psychologists,
mathematicians, even philosophers and physicists, all suddenly calling
themselves neuroscientists, plunged into the field. No one knew where
they were going, and no one wanted to be left behind.
Memory as a
subject of inquiry and wonder is as old, perhaps, as man. The ancient
Greeks variously proposed that memory and other mental processes were a
function of the heart, the lungs or the brain, which eventually became
the agreed-upon site. Beyond locale, however, little was learned about
the processes of mental activity for the next 2,000 years.
Although
the great Spanish anatomist Santiago Ramon y Cajal proposed in the late
19th century that the brain was composed of tiny cells called neurons
and that memory might be stored at connections between neurons, there
were plenty of scientists who thought the whole mental apparatus too
ineffable, too mysterious a subject to yield to laboratory examination.
The
seminal event in the modern history of memory research occurred by
accident in 1953. In an effort to stop horrific epileptic seizures
afflicting a young Connecticut man, a neurosurgeon named William
Scoville removed a portion of the man’s brain. The surgery stopped the
seizures but rendered the man, known in the literature as H.M.,
incapable of forming new memories. His memory of events before the
surgery was uninhibited.
A main portion of the brain that Scoville
removed was a temporal lobe structure called the hippocampus. The fact
that H.M. could no longer form memories but could recall older ones
suggested strongly that the hippocampus was crucial to making but not
storing memory. It immediately made the hippocampus the central focus of
memory research, a position it had maintained when Lynch, just 26,
arrived in Irvine in 1969.
Lynch was wild-eyed, bushy-haired and
bearded, a man of his time — a bit too fully, perhaps. It was the ’60s,
it was Southern California, land of eternal light and endless good
times. Lynch lived in a party pad on Balboa Island.
By every account,
including his own, Lynch ate badly, drank heavily and slept hardly at
all. There were days he seemed to consume more cigars than calories.
“Gary doesn’t sleep,” said Michel Baudry, a 10-year veteran of Lynch’s lab. “He’s incredible. I don’t know how he survives.”
Another
researcher, Kevin Lee, recalled that for a period in the 1970s, the
only things he ever saw Lynch eat came out of a vending machine, a
single vending machine. His main meal consisted of salted peanuts mixed
into soft drinks.
“You know, Gar,” Lee recalled telling him, “you
might think about diversifying your diet. Nothing radical, but hey, man,
try a new machine. Have some chips.”
Lynch’s diet was of a piece
with his extreme work habits, which typically included seven days a week
of 12-hour or longer shifts in the lab, often followed by monumental
bouts in the nearest bar.
Given 400 square feet of lab space and $900
to equip it, Lynch quickly made discoveries having to do with the
brain’s ability to repair some damage to itself after injury. It had
generally been thought that the brain was static, that it did not
produce new cells or structures after it reached maturity. Lynch and
others began to wonder whether the brain did not possess more
malleability, what was called plasticity.
In 1973, just as Lynch was
expanding his investigation of brain plasticity, a pair of scientists in
Europe discovered that when they stimulated the hippocampus with
electric current intended to simulate brain activity, connections
between hippocampal cells were strengthened and, more important, those
strengthened connections could be retained indefinitely. They called the
phenomenon long-term potentiation (LTP).
The combination of brief
stimulation and long-lasting effect matched the key characteristics
scientists had long associated with memory. Lynch and others wondered
whether LTP was the biochemical process underlying memory. A global race
was on to prove it. Thirty-two years later, Lynch hoped he was near the
end of it.
‘A strange place’
Lynch lab has been staffed over the
years by a succession of visiting scientists, grad students,
postdoctoral researchers, dope peddlers, English majors and whoever else
was swept up in Lynch’s often irresistible aura.
All of the
inhabitants have been very bright, some brilliant. A number have gone on
to chair university departments, to found successful companies or to
publish distinguished papers, but when they were in the Lynch Lab, there
wasn’t much to recommend them to civil society. Any hint of future
distinction was obscured by the chain-gang grind of life in the lab.
Lynch’s
extraordinary drive and ability to make every person feel that he or
she was working on the single most important experiment in neuroscience
history was the oxygen the lab lived on. Especially in the early years
— a period Lynch called “the boy lab” because of its
testosterone-driven internal competitions — the lab was a woolly place,
not far removed in its culture from a Neanderthal cave. The guy with
the biggest club generally got his way. Lynch, while not at all
physically imposing, had a ferocious temper and never left a shadow of a
doubt about his willingness to swing whatever was at hand. The history
of the place was littered with battered telephones and drywall with
holes suspiciously the size of baseball bats and fists.
“That’s part
and parcel of the fire that burns in him,” said Lee, who now chairs the
neuroscience department at the University of Virginia. “The phone on the
wall? It just looked like a baseball sometimes.”
“He never really
hurt anybody physically but himself. Although there were people with
emotional scars, I can tell you,” said John Larson, now of the
University of Illinois at Chicago.
Lynch said: “That lab was a
strange, strange place. A lot of weird, weird, different kinds of
people. The dean would look at it and say, ‘That’s a strange damn
place.’ I’d answer: ‘Have you looked at me?’ “
Amy Arai, a native of
Japan, recalled the culture shock she felt when she joined the lab in
the 1980s. “In Japan, everything is very formal. Scientists wear jackets
and ties to work every day. Here in Irvine, nobody did that,” she said.
“I had a hard time even locating Gary. . . . I wandered around looking
for him. There were lots of people wandering around, including one
particularly scruffy guy I saw in the hallways, shirt always untucked
and dirty. I’d sort of hold my breath when I passed him in the hall. I
thought he was a janitor.”
One day, weeks after arriving, Arai was
summoned to Lynch’s office, which was removed from the rest of the
faculty offices in a double-wide trailer next to a parking lot. Arai
walked in and found the trailer empty except for the “janitor,” who was
sitting behind a desk smoking a cigar. It was Lynch.
Baudry, a
Frenchman, toured labs in the United States for five weeks in 1978, then
went back to Paris and told his professors he was going to join Lynch.
Baudry recalled the reaction of Jean-Pierre Changeux, the rising star of
French neuroscience: “He looked at me. He said, ‘You’re crazy. Gary
Lynch? The hippie of neurobiology?’ I said, ‘I’ll take my chance.’ I
went to Gary’s lab, and it really was something different in its
ambience. All these wild people. The contrast with Paris — fields, cows
around the campus. I thought, I have to give this a shot. It really was
the Wild West. And Gary really was this wild person.”
Lynch still
draws an off-kilter collection of researchers. His latest lab — the
“girl lab,” as he described it — included a grad student who wasn’t
officially assigned to the lab, a postdoc who ended up there by virtue
of being kicked out of her original department, and a preternaturally
talented undergrad who was hanging out only long enough to decide which
med-school scholarship to accept. The senior scientists, except for one
man who never left his private office, were three women, who seemed to
speak with one another as seldom as possible.
Work was assigned
largely by Lynch’s judgment of who could do what. If an undergrad was
able, he would find himself in the middle of crucial experiments.
The
lab has changed locations and varied in size over time — anywhere from
three dozen people to as few as six or seven. In January 2005, there
were around a dozen regular members, with students floating in and out.
Much
of the work was some variation of two basic LTP experiments. One
involved isolating single neurons, which, using high-powered
microscopes, were identified, then pinched with a clamp to hold them in
place. This was exceptionally tedious. Researchers could go entire days
without successfully clamping a single cell.
The other experiment
entailed placing a thin slice of a rat’s hippocampus in a nutrient bath
in which it stayed alive for hours, then imposing one of a variety of
conditions on the slice — usually, infusing it with chemicals known to
inhibit or incite certain molecular reactions — then stimulating the
slice with a precisely timed, placed and quantified electric impulse and
measuring what happened to that impulse.
What the scientists were
trying to find out by blocking or inviting the action of certain
molecules was what role they played in LTP. Theoretically, you could
determine all of the principal agents by this process of elimination.
In
practice, people spent extraordinary amounts of time — hours at a
sitting, days or weeks in succession — staring at graphical renderings
of the results on computer screens. It was not work filled with obvious
drama or even, except for making the occasional note in a lab journal,
movement. The lab was quiet — no music; no telephones; low
conversations, when there were any at all.
Lynch lived in dread of
being scooped on discoveries. The residents of the lab did not gush in
praise of his patience. He strode among the benches several times a day
to see how much progress was — or, more usually, was not — being made.
Lynch
talked often about hating the day-to-day process of science, the actual
experiments. He could hardly bear to wait for them to be done to prove
what he suspected to be true.
One day, explaining his distaste, Lynch
said, “There is so damned much housekeeping. The problem is, biology is
a very horizontal science. You have this result over here, that one
over there. None of it lines up.”
His lack of enthusiasm for working
on the bench meant that he needed others who were both capable and
willing to do it. No wonder he was unhappy about the rash of holiday
vacations.
‘Shadow land’
The person lynch was most unhappy with
was Eniko Kramar, a postdoc neurophysiologist who was running the
crucial experiment Lynch expected to prove his basic theory of memory
encoding. Kramar could hardly be regarded as a slacker. She typically
worked longer and harder than anyone in the lab, excepting Lynch.
Having
come relatively late to neuroscience, she was approaching a point in
her career where she needed to make discoveries, then move on to lead
her own lab, or remain locked in subordinate roles. She had become, like
Lynch, a virtual scientific monk, paring away other activities in her
life until all that remained was the lab. Unlike Lynch, she had actually
had a wide range of outside interests — family, friendships,
athletics.
Although it seemed to her at times that the more she did,
the more Lynch demanded, they were in important respects a good team. He
was a synthesizer. She was a pointillist, a technically minded bench
scientist who took care to not extrapolate beyond the results on her
screen. She sometimes found even those suspect, wondering if some
mistake hadn’t deceived her into false optimism.
When Kramar returned
from her brief Christmas holiday, she plunged back into the experiment,
which she had been planning since the previous summer. It involved
using a novel staining technique that would let the researchers actually
see changes in neurons.
A key part of Lynch’s conception of LTP, and
thus memory, was that the process initiated a micro-scale remodeling of
the interior skeleton of cells at synapses.
It is generally agreed
that memory is somehow built out of networks of brain cells called
neurons. How those networks get built is the central question of memory
research.
Researchers have established that when you experience a
sensation in the outside world — perhaps seeing, smelling or touching
something — the sensation is translated by the sensory organs into an
electrical signal that is routed to neurons in the brain, where, if the
signal is strong enough within individual neurons, it causes chemicals
called neurotransmitters to be released onto neighboring cells.
Neurons
are not physically connected to one another. There are tiny spaces
called synapses between them. The neurotransmitters travel across the
synapses. Think of the neurotransmitters as keys. On the surface of the
neighboring neurons are molecules that receive the neurotransmitters.
These are called receptors. Think of the receptors as locks. When
neurotransmitters attach to receptors on the surface of a receiving
cell, when the key opens the lock, channels open into the cell.
It is
because the neurons are not physically connected that communication
between them is never certain. You never know whether a key is going to
find a lock. This is thought to be why any cognitive activity, including
memory, is approximate. Sometimes the connections are made; other times
they are not.
The LTP hypothesis can be summarized by saying: After
two neurons have successfully made contact once — that is, after the
neurotransmitters have attached to receptors — the next time the
original cell releases its neurotransmitters, there is a much greater
chance the neighboring cell will receive them. There is a greater chance
a key will find a lock.
Lynch’s longtime goal was to figure out why.
The general outline of his hypothesis was this: Once a neurotransmitter
attaches to a receptor, opening a channel into the cell, calcium pours
through the channel, setting off a chemical cascade inside. The end
result of that cascade is an interior reorganization of the cell.
A
key molecule involved in the interior remodeling is called actin, which
is a structural protein used throughout mammalian biology to build
internal cell scaffolds. In the same way the outside of a house reflects
the shape of the frame beneath it, when an internal cell scaffold is
altered, the exterior of its cell is changed too. In this case, Lynch
thought a portion of the cell would become squatter, with more surface
area. The greater surface area provides space for more receptors. The
greater the number of receptors, the greater the chance of a
neurotransmitter finding one and making a connection between the two
cells.
The lab had recently developed a method in which the actin
scaffold proteins could be labeled with a dye. The labeling would occur
only after the actin changed shape; in lab terminology this was referred
to as polymerized actin.
The idea of Kramar’s experiment was that
after inducing LTP with the usual electric stimulus, portions of the
cells would restructure, creating polymerized actin. Because the actin
was stained, you could actually see it under a microscope. If you could
see it, it would mean Lynch had been correct in proposing that the whole
physical remodeling, the actin polymerization, was the end result of
LTP.
That reorganization, in turn, strengthened the connection
between cells; networks of those neurons with strengthened connections
constituted the underpinning of memory.
When Lynch had originally
proposed this sort of rapid structural change at synapses, many in the
field were skeptical. Eventually, most researchers came around to the
view that some sort of structural change occurred, but it was taken more
as a matter of faith. Even many who believed the structural rebuilding
occurred thought newly synthesized proteins from the cell nucleus had to
be sent to the synapse to do it, and they spent an awful lot of time
looking for those proteins.
Lynch thought it would take too long for
the proteins to be manufactured in the cell nucleus; events were already
underway, and the material needed to complete the job was on hand.
Imagine
a construction crew framing a building. If the protein synthesis
believers were right, the carpenters would have to call a warehouse
every time they needed a nail. Lynch proposed that the crew had the
nails right there in their belts. This experiment was intended to
provide proof.
“We’re in the penumbra, the shadow land,” Lynch said. “And now comes the moment of moments.”

________________________________________

Los Angeles Times
Monday, August 20, 2007

Chasing Memory | Second of four parts
Trials, and a series of errors, in the brain lab
*
Gary Lynch’s UC Irvine research team struggles to understand how
memories are made. It is crucial to treating age-related declines.

By Terry McDermott, Times staff writer
The
myth of modern science, that it proceeds carefully, rationally,
incrementally, building bit by bit from rock-solid foundations to
impregnable fortresses of fact, comes unraveled in contemporary
neuroscience. Fortresses, entire kingdoms of neuroscience have been
built on what turn out to be frail premises that get swept away entirely
when the next new thing comes along.
A few years ago, a huge amount
of effort was spent researching the then-thought marvelous qualities of a
humble molecule called nitric oxide. This molecule, better-known in the
broader world as the key element in laughing gas, was celebrated as a
vital actor in human memory and cognition.
Science Magazine, as if honoring a rock star or president, put the thing on its cover and declared it Molecule of the Year.
By
the end of the next year, nitric oxide had fallen off the end of the
Earth. Little of what had been claimed on its behalf turned out to be
true. This was but one example in a long, sad tradition of a science, as
if gripped by mass hysteria, going off the deep end and pretending it
knew how to swim.
There was no guarantee, neuroscientist Gary Lynch
liked to say, that something was important just because you happened to
study it.
“You always imagine those animals out in a herd, the
wildebeests — they’re running along, and a lion jumps up and takes out
this guy named Clyde,” Lynch said. And the world proceeds as if Clyde
never happened. “They don’t talk about Clyde anymore. It’s just not good
form to talk about him.”
Lynch, who runs a lab at UC Irvine, has
spent three decades studying a phenomenon known within neuroscience as
long-term potentiation, or LTP, which can be very loosely defined as a
process in which electrical stimulation strengthens connections between
brain cells. Lynch had taken up the study of LTP because it had
characteristics strikingly similar to human learning and memory. It
seemed to take place in parts of the brain where memory was thought to
occur, and like memory, it occurred in an instant and could last a
lifetime.
The practical reality of memory — that human beings, from
very young ages on, learn and store information — had been established
and studied for millenniums. How it happened, however, remained a dark
continent yet to be mapped.
When people, even scientists, talked
about memory, they likened it to objects or concepts in everyday life.
They talked about filing cabinets, and photographs, and videotape
replays. They almost never talked about what memories really, physically
were. Why? They didn’t know. LTP seemed an excellent candidate to be
that physical, molecular underpinning of memory.
Beyond the pure
scientific intrigue of it, memory research has grown more important as
medical advances allow more people to live into old age. With longevity
has come an epidemic of memory failure among the aged. Alzheimer’s
disease and other forms of crippling dementia threaten to make living
longer less a blessing than a curse.
Things can go radically wrong
inside an old brain, and unless you understood how the physical
processes of memory worked, Lynch thought, you’d never be able to fix it
when it broke.
Brain scientists generally agreed that networks of
neurons somehow wired together in the brain were fundamental elements of
memory. LTP was hypothesized as the means by which that wiring
occurred.
Lynch had bet his career that he could work out the details
of LTP, and that what he found would matter — that it would turn out
to be something beyond “an interesting little bit of biology.” There
were too many similarities between LTP and memory, he thought. The gods
were unpredictable, he said, but seldom that cruel.
LTP had been
discovered in 1973 not as a naturally occurring phenomenon, but in the
artificial and arbitrary conditions of a laboratory experiment. Because
of this, not everyone was convinced LTP had significance outside the
lab. As one of Lynch’s rivals, Nobel laureate Eric Kandel of Columbia
University, said: “You know what LTP is? It’s an artificial way of
stimulating your brain. Who knows if this is what happens in learning
and memory?”
That no one had figured out LTP was due largely to the
inherent complexity of brain biology. Seth Grant, a neuroscientist at
the Sanger Institute outside London, has counted more than 1,000
proteins thought to be involved in memory. If even half of that number
actually were involved, isolating and understanding the behavior of each
would be a Herculean undertaking.
Lynch was more prosaic. “It’s a bitch and two-thirds,” he said. “And stupid too.”
Other
scientists had moved on. “The boys,” as Lynch routinely referred to the
neuroscience establishment, turned en masse to the exploration of what
genes might be involved. That they were able to find such genes almost
at will was read by gene proponents as a reason to stay and look for
more. To Lynch, it made no sense. “It’s like trying to understand a
computer by studying it a transistor at a time. Not only will it take
forever, it will never work. You’ll never get there unless you
understand the programming.”
“I asked myself: ‘Why am I following you down this alley?’ So I didn’t.”
Lynch
instead rode out the LTP bet. In January 2005, after decades of
studying and teasing out its details, he was in the midst of an
experiment to determine once and for all whether LTP was a mere
laboratory curiosity or the real thing — the means by which neurons
were wired together to form memories. Lynch, in other words, was about
to find out whether he was a candidate to stand Nobel shoulder to Nobel
shoulder with Kandel. The alternative? He was Clyde.
Lynch had long
ago proposed that the end result of LTP was a micro-scale physical
remodeling of neurons that allowed them to communicate better with one
another. The lab had just developed a new technique that Lynch thought
would allow researchers to visualize this remodeling, in fact, to see
the physical trace of memories.
This new technique promised to answer
conclusively what had been supposition, and to answer it in such a way
you would literally see the result.
Neuroscience comprises many
distinct disciplines, or tribes, as Lynch called them, ranging from
mathematicians to evolutionary biologists. Eniko Kramar, a senior
scientist in the lab, was actually going to run the experiment. Kramar
was a neurophysiologist, meaning she studied the function of brain
cells. Physiologists, generally, can be thought of as engineers. They’re
practical people, interested in how stuff works.
At the moment, the
stuff in question was synapses in a rat brain. Most brain research labs
used animals in their work, the main reason being the lack of human
subjects willing to have their brains dissected. Lynch Lab used rats
almost exclusively.
Other labs used mice or simpler creatures —
fruit flies and sea slugs. Genes that performed certain known functions
in fruit flies did the same or similar work in humans; the genes were
conserved, scientists say — natural selection winners passed up the
evolutionary chain. Use of these animal models is a daily expression of
unquestioned trust in evolution as a central fact of human history. Even
as debates might rage in broader society over the idea that human
beings are descended from apes, there was a strong conclusion in biology
labs that human antecedents go back way past the apes to the flies and
beyond.
Still, Lynch wondered about the practicality of studying
memory in nonmammals that, in human terms, didn’t have any. “Memory is
an emergent phenomenon,” he said. “Steam is an emergent phenomenon. If
you want to study steam, you better study hot water. You ain’t going to
get steam out of mud.”
Visualizing success
Kramar’s experiment
began with the death of a rat, which she accomplished using a small
guillotine. It took her less than five minutes to decapitate — or, as
she put it, sacrifice — the animal, cut open its skull, remove the
brain and separate the hippocampus, a portion of the temporal lobe
thought to be involved in memory, from the rest of its cortex. She then
sliced the hippocampus, which in a rat is about the size of a clipping
from a thick thumbnail, into five very thin sections.
The slices were
transferred to a small, circular Plexiglas chamber centered on a
workbench under a microscope. The chamber was fed by separate lines
carrying a nutrient-rich warm liquid and oxygen, which together kept the
brain alive and in some sense functioning.
The top of the chamber
had cutouts that allowed electric probes to be placed into the brain
slices, one for stimulating and one for recording. The stimulating
electrode could be set to deliver currents of precise timing and
duration. Lynch and colleagues had discovered decades before that LTP
was optimized when initiated by electric currents that mimicked a
naturally occurring rhythm within the human nervous system known as
theta rhythm. This coincidence — that the best way to obtain LTP in the
lab was to mimic actual real-world biology — had, more than anything,
convinced Lynch that LTP was real.
“That day — the day we found theta — our mouths fell open,” said John Larson, who worked with Lynch on the discovery.
Kramar’s
chamber was situated on a table equipped with shock absorbers to
prevent the rumble of a truck or car outside from disturbing the queasy
equilibrium of the experiment. Because the electrical measurements
needed to be precise, any equipment that might interfere with them was
grounded, and connections were shielded with aluminum foil to prevent
stray signals from intervening.
With all the foil and electrical tape
and ground wires, the whole apparatus, which the scientists referred to
as a rig, had a kind of jerry-built, Rube Goldberg quality.
After
the slice was stimulated to mimic the theta rhythm, the current would
pass along previously identified pathways, setting off biochemical
reactions. The recording electrode measured the current as it exited the
slice. The data were fed automatically into a computer program that
translated and graphed the results. If LTP occurred — that is, if more
neurons were wired together — more current would move through the slice
after the stimulation than before.
Then, by using chemicals to block
the actions of different molecules within the brain slice, the
scientists could tell whether those molecules were essential to LTP. By
laborious process of elimination, they ought to be able to unveil the
entire process. The essential parts of that process would be the
fundamental building blocks of memory. They had done a great deal of the
work already, identifying what they thought were the key steps. They
hoped their new visualization technique would allow them to actually see
some of those steps.
Kramar was an exacting person, naturally
fastidious in setting up the experiment. She knew what was at stake. She
didn’t need Lynch leaning over her shoulder to tell her this was
important. Theirs was a fraught relationship. In one important way they
were complementary. Lynch was given to big-picture conceptualization,
while Kramar lived at the level of the brush stroke. Their temperaments
were so utterly different as to be nearly opposed.
Kramar punched a
key on her computer, initiated the electric pulse and waited. The
recording electrode was picking up interference from elsewhere,
overwhelming the readings. Kramar tried, patiently at first, to isolate
and banish the interference. She spent hours looking, but never found
the source. The brain tissue died in the chamber.
The day’s work
ended before it ever got underway. Lynch wouldn’t be happy and Kramar
knew it, but she was more upset that she had killed a rat to no good
end.
The seemingly random interference was the kind of thing that
drove the scientists crazy. It wasn’t enough that they were opposed,
they felt, every step of the way by the complexity of the biology. They
had to fight their way to even get to the biology. Kramar had run slices
in other experiments in this rig for months without anything like this
ever happening.
“You do the same thing every day for a year, and then
one day, for no reason at all, you can’t do it,” she said. “It makes no
sense, but you just have to come back and do it again.”
The next
day, the interference was gone, a ghost vanished. Kramar ran the
experiment, stimulating the slices, taking her readings, then infusing
the slices with a dye that would stain only the portions of the neurons
that had been changed in the experiment. They would show whether Lynch’s
hypothesized remodeling had occurred.
Afterward, she packed the slices on ice and took them to a nearby lab, where they would be prepped and mounted on slides.
Two
days later, Kramar, with her own and Lynch’s great anticipation, got
the mounted brain slices back. They were worthless. Either the
experiment had failed to produce any effect on the slices — unlikely —
or the slices had been improperly mounted. She would have to start
over.
Lynch said: “You’re always surprised or horrified or pleased or
something. It’s not what you expected. It’s always a bunch of crap.”
A whimper, not a bang
The
next day, Lynch said: “Several years ago, I sent a student out and
said, ‘Your job is to find out what the boys know about assembly.’
That’s what grad students are for. They’re the cannon fodder of science.
You throw them at problems that have no chance of being solved. One
day, the student came back and said a new thing — integrins.”
Integrins
tie cells down to a particular place. They fix, for example, blood
cells into place so that a cut will clot. Think of them as a kind of
cellular thread that stitches cells into place.
As is typical in
biology, molecules that perform a specific function in one place often
perform some variation of the same function elsewhere. So Lynch presumed
with integrins. He made them a key part of his investigation, and the
lab had since reported that integrins in the brain fastened neurons in
place, locking in the changes LTP created.
“The only thing that keeps
the neurons in your brain from rolling out your nose is the fact that
they’re stuck together at adhesion junctions. The adhesion junctions are
actually the synapses,” Lynch said. “It’s the boring biology of
wound-healing, of blood platelets clotting. . . . That’s what I love
about it, you know. It’s like the T.S. Eliot thing — when it’s all over
it’ll be a whimper, it won’t be a bang. It won’t be a magic protein, it
won’t be a special gene. It won’t be any of that crap. It’ll be
watching a cell crawl across the dish.”
Just then, Ted Yanagihara, a
gifted undergrad who was working with Kramar, poked his head inside
Lynch’s office. It was a mark of the meritocratic nature of the lab that
Yanagihara, just a kid, really, was entrusted with such work.
“Bad news,” he said. “I have a result, and it’s not a good result.”
Kramar,
in addition to the visualization experiment, was working with
Yanagihara trying to gather further evidence that integrin molecules
were one of the building blocks of LTP. They used chemicals, called
antibodies, to stop integrins from having any effect. If they did, they
hypothesized, they would block the final stage of LTP.
Kramar was
doing one version of this test in her slice experiments. Yanagihara was
doing another working with single cells. Everybody was on edge. Some
days the methods failed and they couldn’t gather results. On days their
methods worked, the data were wrong or confusing. Sometimes the
integrins were blocked, sometimes they weren’t. This continued for two
weeks.
Said Lynch: “That’s the trouble with biology: There are just too damned many variables.”
“It’s a wonder anything ever gets done,” Yanagihara said.
Lynch said: “If you want clean results, go be a physicist.”
Lynch
accepted the repeated failures with surprising equanimity. He’d been
through worse droughts before. One experiment in the early 1980s took
two years. It had turned an entire cohort of grad students into a
contemporary legion of the damned, but the legion kept marching and
eventually the experiment succeeded.
Lynch amused himself. He had
been shopping for a new car, his first since a 1987 Ford Mustang that
was now falling apart, day by day, piece by piece, in the parking lot.
Lynch had been a drag-racer as a kid, and cars, along with single-malt
Scotch whiskey and good books, were among the very few possessions he
cared much about.
He wanted a brand-new Chevrolet Corvette
convertible, a formidable machine, but he couldn’t find the model he
wanted nearby. While everything was falling apart in the lab, he found a
dealership in San Jose that had it. He hopped a flight, wrote a check
and drove the car home that night.
Kramar took the setbacks more
personally. By the end of the month, she was exhausted and did the
unthinkable — she took a weekend off. “I thought, I’m not even going to
show my face. When I get like that, I have to back away from
everything. I went out and bought books, then sat home and read them.”
The mood in the lab had grown very dark. One day, Yanagihara said to Kramar: “This is a nervous moment.”
Kramar replied: “You’re nervous? It’s my career at stake.”
She
laughed dryly; no one laughed along. The integrins were a crucial part
of the hypothesis, and apart from what would be the acute embarrassment
of having to retract previously published conclusions, Lynch had no
alternative explanation. His entire research program would be a
shambles.
It was remarkable that so much could go so wrong all at
once. Some days the electric probes were too noisy to produce reliable
results. One day a computer melted, smoke rising from its innards.
Programs crashed. A day’s work was halted when a grad student couldn’t
make it to the lab.
“The answer is sitting there waiting for you, and
you can’t do anything about it because your graduate student got his
car impounded,” Lynch said, then went off on a long rant about the
torture of academic biology. “If I never go to another meeting, get
involved in another symposium, I’ll be happy. I don’t care if I ever
train another graduate student. Don’t get me wrong. I’m pleased with the
way they’ve turned out. Lots of them have gone on to do interesting
things. But I want to be done. Done. Over.”
There was an occasional ray of hope.
Yanagihara
one day, finally, working with the brains of young rats, got his
experiment to work right, and found the result he was expecting to find.
“If we get it tomorrow in middle-aged rats, it’s great,” he said.
“If you see a garbage can flying out of the lab onto the hedge, you’ll know we didn’t,” Kramar said.
The
next day, the trash cans remained inside, but only because nobody had
the energy to throw them out the window. The experiment had failed
again.

________________________________________

Los Angeles Times
Tuesday, August 21, 2007

Chasing Memory
Third of four parts
Breakthroughs, and new crises, in the lab
* As Gary Lynch’s team starts piecing together the story of memory, his health prompts him to look into his own brain.


By Terry McDermott, Times Staff Writer
Lynch
Lab sits between a toll road and the UC Irvine main campus, in an
office park of indistinguishable low-rise, beige-on-beige stucco
buildings. Neuroscientist Gary Lynch had moved his lab and office — for
a while, just a desk in a hallway — numerous times during his Irvine
career, often as the result of some feud or slight. He ended up in the
office park largely because everybody — including him — concluded all
parties would be better-served if there were physical distance between
Lynch and his university peers.
The lab is at 101 Theory Drive, a developer’s idea of a scientific street name that Lynch found presumptuous.
It
is a mark of the difficulty of life sciences — biology and its many
descendants — that to call something a theory is to honor, not slight
it. Theory, evolutionary biologist P.Z. Myers has written, is what
scientists aspire to. Lynch, for all of his bombast, was respectful of
the intellectual protocols of his science.
“I would have called it Hypothesis Drive,” he said.
The
hypothesis is the fundamental organizing principle in scientific
research. Its “if this, then that” structure underlies almost all
scientific experiments. The work in Lynch’s lab has been driven by a
single overriding hypothesis Lynch first published in 1980.
Lynch
proposed that the fundamental act by which a memory was encoded involved
a nearly instantaneous physical restructuring of portions of brain
cells, called neurons. That restructuring allowed neurons to be built
into small networks. Each small network would be a memory, he thought.
Lynch’s
research focused on a particular area of the brain, a structure called
the hippocampus, long thought to be involved in memory. Most neurons in
the hippocampus have roughly triangular bodies. Slender fiber extensions
called dendrites sprout from the top and bottom. The branches coming
out of the top are called apical dendrites. Those coming from the bottom
are called basal dendrites.
Also coming out of the bottom is a
single larger extension called an axon. All along their lengths, the
dendrites are marked by microscopic nubs called spines, thousands of
them per dendrite. The axons of one neuron extend to meet the dendritic
spines of other neurons. These dendrite-axon junctions are the synapses.
Lynch
proposed that the dendritic spines at these junctions changed shape
during a process known as long-term potentiation (LTP), which resulted
in the strengthening of the bond between a dendrite and an axon. The
remodeled dendrites, he said, were the base elements of memory.
Lynch
acknowledged that the details of the biochemical interactions that
caused the shape change were complex and not well-understood — at the
time he originally proposed it, in fact, not understood at all.
But
the crux of the hypothesis was that human interaction with the
environment — a glimpse of blue ocean, the touch of a silk scarf —
resulted in an actual physical change in cells in the brain, and that
those changes were the underpinning of memory.
Two notable properties
of memory are its vast size and that it can be made in a moment, yet
last essentially forever. Any attempt to describe the physical
components of memory had to account for those properties.
There are
about 100 billion neurons in the human brain. Each neuron has dozens of
dendrites, and each dendrite has thousands of potential synapses. So the
synapses offered immense storage capacity. But how could storage be so
long-lasting? “For me it was really, really obvious it had to be
structural, but beyond that, what could I tell you?” Lynch said.
Aside
from the pure scientific achievement of understanding the proposed
memory mechanism, the value to ordinary people has become more apparent
almost every day since Lynch proposed it.
We are in the midst of a
brain failure epidemic. Worldwide, it is estimated that by 2040, more
than 100 million people will suffer some form of dementia. The physical
mechanisms of memory break, and do so with frightening frequency. Lynch
was fond of saying that you had no hope of fixing it if you hadn’t first
figured out how it was supposed to work.
To understand why it can
take so long to figure out, imagine a vast pile of broken plates. A
hypothesis is what someone, after surveying the pile, might say about
putting the pieces back together.
If the hypothesis holds for a
while, survives challenge and criticism, much of it improbably hostile,
it might eventually come to some rough, general acceptance and be joined
to other hypotheses to form something more far-reaching.
Neuroscientists habitually use a particular word to categorize such a
body of thought — or collected wisdom — in a part of their science.
They don’t, as a layperson might, refer to it as a theory. Instead, they
call it a story.
Lynch was perilously close to believing he knew the story of human memory — why it exists, how it works, how it fails.
Eric
Kandel, a Nobel laureate for his own memory research and a competitor,
said of Lynch’s work, “The current view of LTP is Gary Lynch’s view”
meaning that much of what was known about the process was what Lynch had
discovered.
“He was the first person to fully appreciate the
significance of LTP as a physiological phenomenon — very prescient,”
said Richard Morris of the University of Edinburgh.
At the time it
was published, Lynch’s hypothesis was a lonely view. By January 2005,
after he and others around the world had spent decades collecting
evidence, the hypothesis was widely shared but by no means proven.
Lynch
and his colleagues then embarked on a series of experiments intended to
prove or disprove his hypothesis once and for all. And, as if to taunt
Lynch, almost everything the lab did for a month didn’t work. Computers
crashed, fundamental experiments backfired, grad students went AWOL. The
whole research program seemed imperiled. The lab was not a happy place.
Middle-aged aging
What
brain scientists love about the hippocampus, apart from its presumed
significance, is that it is largely a good neighborhood in which to
work. It is, compared with much of the rest of the brain, orderly,
neatly layered and segregated. It is well-mapped; its regions named with
straightforward simplicity — CA1, CA3, etc. If you chose to poke
around there, you generally knew what you were poking.
There is a
place in the hippocampus, however, where the naming scheme as well as
all ideas about the function of the thing ran aground. The place is
called, in an appropriately Middle Earthian way, the mossy fibers.
“It’s
the strangest connection in the brain — the strangest thing in the
mammalian brain — right in the middle of the memory structure,” Lynch
said. “There’s no hypothesis as to why these things are sitting in the
middle of the memory structure.”
“These things” are very long,
faintly furry axons extending from neurons in an area of the hippocampus
known as the dentate gyrus. That they are there bugged Lynch to no end.
Here sat something dead-center in the thing he had studied for decades,
and he was utterly perplexed by it. Clues were gathering, however.
Laura
Colgin, a postdoc in the lab, was intrigued by weak electric pulses
that apparently originated in the same area as the mossy fibers. Other
researchers had reported similar low-frequency waves occurring elsewhere
in the brain during sleep and periods of wakeful rest. They called them
sharp waves. No one knew what the sharp waves did until Colgin
discovered that in the right circumstances, they seemed to erase LTP. In
other words, if LTP was the mechanism for memory, sharp waves could be a
mechanism for forgetting.
As counterintuitive as it seemed, the idea
that there might be an active forgetting mechanism made sense. No one
remembered or would want to remember everything. There had to be a way
to get rid of stuff.
“The brain is set up to detect patterns and
causality, even when they don’t exist,” Lynch said. Because the brain
does this, Lynch thought, it would be useful for the brain to have a way
to rid itself of patterns that turned out to be misleading.
“The
assumption that contiguity means causality probably flows from LTP, and
controlling it provides a good excuse for the erasure process,” Lynch
said.
Say, for example, you scratched your nose at the same time a
dog barked. The brain might read the fact that these two events occurred
at the same time as evidence they were related. There was a need to rid
yourself of these sorts of false associations.
While Eniko Kramar
was leading the troubled effort to validate Lynch’s broader LTP
hypothesis, Colgin’s sharp wave research prompted Lynch to begin a
broader investigation of ordinary, non-disease-related memory decline,
which had been documented in the psychology literature for decades. The
general consensus of that research was that memory started to decline
not long after humans reached physical maturity. You peaked at about 20
years of age. After that, in terms of your ability to remember new
things, it was a long, slow, steady slide. Some studies indicated the
rate of decline was pretty much a straight line. You lost as much
ability to memorize between the ages of 20 and 30 as you did between 50
and 60.
To neuroscientists, forgetfulness was largely seen as a
behavioral curiosity, and was not much studied as a biological
phenomenon. When it was, it was generally regarded less as a process
itself than as the failure of the memory process.
In Lynch’s work, it
was clear that LTP was greatly diminished in elderly rats, but there
were no experimental data to support the notion of middle-aged decline.
If LTP were the underpinning of memory, and memory began declining in
middle age, then LTP’s decline should mirror that of memory.
Lynch
decided to do a more systematic search. He assigned the project to a
young graduate student, Chris Rex, who was a student of Lynch’s
collaborator, Christine Gall. Rex had come to Lynch Lab to learn
physiology, and never left. He fell in love, he said, with the direct
contact to the biology that he felt doing physiology experiments. “It’s
more like having a conversation where the challenge is really asking the
right questions,” he said.
Lynch had a vague idea that one reason no
one had ever found LTP decline in middle-aged animals was that everyone
had looked in the wrong places. That they did so was largely a matter
of convenience. The apical dendrites are more numerous and easier to
study, so most research focused there.
Lynch sent Rex off into the
basal dendrites of middle-aged rats, where — wonder of wonders — Rex
found distinct failure of LTP.
Since Colgin’s discovery that
forgetting could be the result not of a failure but of an active
process, Lynch had begun formulating a broader view of LTP. He began to
see it as the result of an exquisitely balanced set of inputs. Some of
the inputs encouraged LTP. Others inhibited it. Such systems are common
in mammalian biology. They can fail from either direction — too little
incitement or too much inhibition. Maybe the failure of memory was
simply the result of this mechanism getting slightly out of kilter. It
could be as simple, Lynch thought, as too much of one protein or too
little of another.
Lynch had known since the early 1990s that too
much of a molecule called adenosine outside a neuron interfered with
LTP. He suggested Rex administer a drug known to block adenosine to
brain slices of middle-aged rats where LTP was inhibited.
After a
couple of weeks of false starts, including another computer crash and
some difficulty administering the drug, Rex decided to wait until after
LTP was induced to block the adenosine. Everybody else in the lab
thought that was a weird idea that would never work. Lynch shook his
head. “Crazy kid,” he said.
At 1 a.m. on a Saturday in February 2005,
Rex erased the LTP deficit. It was completely, utterly gone. Full LTP
was restored. In brain science, even many successful experiments have
vague results that could be read in various ways. Seldom is anything
this clear-cut.
“It should never have worked,” Lynch said.
Rex, a second-year grad student, grinned and shrugged his shoulders. “Pretty lucky,” he said.
The
result was astonishing. What Rex seemed to have discovered was a major
cause — if not the major cause — of one of the most persistent,
widespread real-world effects of aging: forgetfulness. And it seemed to
be caused mainly by too much of a single molecule, adenosine.
After
weeks of repeated failures on almost every other front, Lynch was
ecstatic. “You mean this crap actually works?” he said. “You don’t
expect to see a result this black-and-white. You expect ambiguity. Aging
does not occur uniformly even across a single neuron. It’s an instant
default explanation for memory loss. It’s getting to the point where we
might have to start believing we were right.”
A good rain
Rex’s long-shot success dislodged some karmic plug in the LTP universe. Things started to work throughout the lab.
Kramar
had struggled for weeks with the experiment that was intended to
definitively validate Lynch’s hypothesis that the dendrites of neurons
were physically reorganized during LTP. She had endured a weird barrage
of difficulties: problems with her experimental apparatus, outdated
chemical reagents, improperly prepared specimens.
Suddenly, the
difficulties all disappeared. The experiment started working precisely
as planned, and her results were almost too good to be true.
“The data are perfect,” Lynch said.
Kramar,
still shaken by the preceding bleak weeks, spent hours alone in the
imaging room, studying her results; she was afraid to believe what she
was seeing. She produced stunning images showing the cellular
reorganization Lynch had hypothesized as the end stage of LTP, the step
that locked a memory in.
She did a series of experiments to block
LTP, and the cellular reorganization disappeared. She incorporated Rex’s
adenosine findings. The results were clear-cut. Adenosine blocked the
reorganization. Take it away, and the process worked perfectly. She
blocked the integrins, the molecules that stitched everything else into
place. The LTP disappeared. Every crank of the wheel churned out another
supporting result.
After decades of struggle, all of the pieces were
falling into place. Lynch’s long-standing hypothesis was being borne
out to the smallest detail. He could hardly believe it.
“I have to
say, I’m flabbergasted,” Lynch said. “I genuinely believe that what
we’re staring at is the exact thing that occurs in adult mammals as they
lay down memories. The exact thing. That’s it. That’s what I wanted. I
wanted to see the thing itself. . . . It’s just colossal. It’s a very
hard thing to believe.”
The mundane nature of the molecules involved
made the findings more convincing, Lynch thought. No divine
intervention, no magic gene — “just another lift from the parts bin,”
as he termed it.
“You give up grandeur, but in return you get confidence,” he said.
One
morning, not long after, Lynch woke up and could barely get out of bed.
He had no balance. He couldn’t walk down the hall. Within the week, he
developed an acute respiratory infection. He had an attack of gout.
Another unrelated ancient ailment — caused by a chronic spinal
condition — recurred. It was like a perverse illustration of his
constant complaint that you never knew what was about to go wrong.
The
lack of balance was thought to be the result of a viral infection of
the inner ear; his doctor sent him to get an MRI to rule out problems
deeper inside.
Each image of a typical MRI shows a very thin slice of
whatever body part is being examined. A brain MRI produces in digital
form what you would get if you were able to take a very large kitchen
mandoline and work your way down, slice by slice, from the top of a
skull to the bottom. The resulting stack presents a digital photo album
of the inside of the head. After the exam, Lynch asked for copies of the
images.
He left the clinic that day with a CD-ROM containing the
interior images of his head in the pocket of his black cotton jacket. He
hopped in his brand-new cobalt-blue 400-horse Chevrolet Corvette
convertible and headed toward Irvine.
Lynch is a torque man. He
drives very fast, especially in the lower gears, where the experience of
speed is visceral. Unless he’s on the freeway, he seldom gets out of
third gear. Of course, in the Corvette, third gear can mean flying 100
mph down a blind alley, giggling like a schoolgirl.
Lynch took the
CD-ROM back to his lab, where he fed the images into a computer program
that allowed him to scroll from top to bottom — like riding the Magic
School Bus with Ms. Frizzle — through his brain. Upon first sight of
his own brain, Lynch began to make some not altogether happy noises.
There were low whistles, smacked lips and much muttering. He shook his
head a couple of times. He grew uncharacteristically somber.
MRIs, as
useful as they can be, remain crude tools. They allow one to see larger
structures inside the body. Unfortunately, the work of the brain occurs
mainly at the micro scale. The MRI would give Lynch a flyover from
35,000 feet; what he really wanted was to blow down through the
anatomical weeds, low to the ground in first gear in the Vette.
Even
so, the MRI revealed cause for concern. The human brain contains in each
hemisphere large cavities — literal holes in the head — called
ventricles, where cerebrospinal fluid is produced. The ventricles in
Lynch’s brain were enlarged. This in itself came as no great surprise.
Ventricle enlargement often accompanies aging. The crucial questions
were how much expansion and from what cause.
As he sat in his office,
looking at his brain blown up to quadruple scale on his giant Mac
monitor, he exhaled, shook his head, pointed, and said, “Boy howdy. That
doesn’t look very good.”
Lynch and a company he co-founded were that
very month trying to get Federal Drug Administration approval to begin
testing the drugs he had invented, called ampakines, in humans. The
ampakines were intended to help alleviate a wide variety of brain
malfunctions.
He slumped back in his chair and said, “You better hope
we come up with something on them ampakines. Normal or not, you don’t
want this.”
Despite the siege of illnesses, Lynch continued going to
the lab every day. One Friday, he realized he had forgotten to submit to
the National Institutes of Health crucial data supporting his request
for renewed funding. The data were long past due. Lynch was stricken.
That
afternoon, everybody in the lab gathered to celebrate the incredible
run of good fortune. Somebody dug out three bottles of Sierra Nevada
Pale Ale and divided it up among a dozen or so people. With no drinking
glasses at hand, they poured the beer into test tubes and tiny chemical
beakers.
Lynch, standing in the middle of the celebration, raised his beaker:
“We
do adenosine, Eni’s integrin experiments. We ran the table. We ran the
table. Then I realize: I forgot my grant. I forgot to send the
supplemental material. I’m a chronic screw-up. I promise you, I’m the
only neuroscientist in history who forgot his grant. This is a screw-up
of biblical proportions. Even I have to say it — that’s a screw-up.
This grant is cursed.”
Lynch stood there, swaying back and forth. His
face, expressive even when becalmed, now seemed about to stretch beyond
the bones beneath it. His jaw worked slowly from side to side, his grin
shifting with it. He leaned on a lab bench for support, to keep
standing.
He stood with his little beaker of beer and grinned. He stood there like that, grinning and quiet and swaying, for a long time.
“It
doesn’t matter,” he said. “I can barely walk a straight line, and I
blew my grant. I’m a chronic screw-up. Who cares? I have this beautiful
science raining down all around me.”

________________________________________

Los Angeles Times
Wednesday, August 22, 2007

Chasing Memory / Final of four parts
Success, with a big dose of rejection
* Gary Lynch expects his lab’s work to bring ‘all the tribes of neuroscience to the same campfire.’ But he meets resistance.


By Terry McDermott, Times Staff Writer
Reflecting
in the spring of 2005 on his lab’s recent successes, which he regarded
as a culmination of decades of work, UC Irvine neuroscientist Gary Lynch
said: “This will be a moment when all the tribes of neuroscience come
to the same campfire.”
He was wrong. There was no reaction. Nothing.
Initially, he couldn’t even get a short paper on a crucial visualization
experiment published. Lynch envisioned the experiment as a grand
confirmation of his notion that a change in the physical structure of
brain cells at the connections between them was responsible for the
encoding and persistence of memory.
It had taken 20 years to acquire
the tools to execute, and when Eniko Kramar, a senior scientist in
Lynch’s lab, produced a series of spectacular microscopic photographs
depicting where and how the change occurred, Lynch awaited the triumphal
acclamation of the lab’s success.
The tribes were not at the same campfire. Many apparently hadn’t yet learned that fire had been discovered.
When
a paper is submitted to a scientific journal, the journal editors send
it for review to panels of scientists. Peer review is the backbone of
contemporary scientific legitimacy and lauded by everyone involved. It
is also an opportunity for mischief and misunderstanding.
Lynch’s
history of antagonizing his peers sometimes made peer review more a
gantlet than a critique. Richard Thompson of USC, a renowned
neuropsychologist, said he had more than once nominated Lynch to
membership in the prestigious National Academy of Sciences, but was told
by other members Lynch would not be elected so long as they lived.
“There’s
a reason for his paranoia. There are a lot of people out there who
don’t like him. Gary doesn’t suffer fools gladly,” Thompson said, then
paused for a moment. He chuckled and said: “And there are a lot of fools
in the world.”
The reviews on Kramar’s paper seemed not to even
acknowledge its main point — that the lab had for the first time
demonstrated the physical reorganization of cells that occurred in the
final stage of long-term potentiation, or LTP, which Lynch believed was
the biochemical process underlying memory.
One reviewer, in
recommending against publication, complained that the scientists had
only looked at a specific set of synapses, which was inexplicable as
criticism. They looked there because that’s where they were doing the
experiment, that was where the condition they were examining existed. It
was as if a traffic engineer, having proposed adding carpool lanes to
the San Diego Freeway, was asked why he hadn’t examined four-way stop
signs in Barstow.
Lynch was irate, and for a couple of days everybody
avoided him. Then one morning, he was at his desk, smirking like a boy
with the key to the cookie jar.
What happened? I asked.
“I can’t tell you,” he said, his grin growing.
But,
of course, it was Lynch; he had to tell. He pointed at his computer
monitor on which was displayed information on a company called Memory
Pharmaceuticals, founded by Nobel laureate Eric Kandel, and a competitor
of Lynch’s biotech company, Cortex Pharmaceuticals.
“I’m shorting Eric’s stock,” Lynch said and cackled.
Kandel
was the god king of contemporary neuroscience. He won the Nobel Prize
in 2000 for investigations of synaptic activity that occurred during
reflex learning in sea snails. He had also almost single-handedly made
the study of protein synthesis a major focus of neuroscience.
Lynch
thought the emphasis was wrongheaded, but there was little he could do.
Kandel, for his part, had nothing but nice things to say about Lynch. He
barely acknowledged they were competitors, despite the fact that the
two had led opposing armies in a 1990s war over where at the synapse the
crucial actions of LTP occurred.
Lynch turned out to be right and won that battle, but he lost the war. Kandel received the Nobel Prize; Lynch went to ground.
For
reasons not entirely clear even to Lynch, he retreated to his Irvine
lab. He focused on his research, continued to publish voluminously, but
largely absented himself from the numerous academic conferences and
symposiums at which neuroscience findings were presented and debated
and, not insignificantly, reputations made and maintained. He declined
to meet with visiting researchers and fought with administrators and
colleagues.
“It got to be very hard for me to keep playing with the boys,” he said.
Few
things are more punishing to an ambitious man than to be right and
unappreciated. The latest fight over the Kramar paper was but one more
slight. In the end, Lynch reacted as he had before: He complained
bitterly then went back to work.
One of the things he devoted time to was the development of a family of drugs, called ampakines, intended to enhance LTP.
Ampakines
had followed a tortuous path, but finally in the spring of 2005,
Cortex, the small biotech firm in Irvine that licensed them, had a
viable drug candidate ready for federally mandated human trials. The
drug, called CX717, had sailed through Phase I safety trials, and Cortex
was now applying to the Food and Drug Administration to conduct
separate Phase II trials for its effects on sleep deprivation,
Alzheimer’s disease and attention-deficit hyperactivity disorder (ADHD).
The trials could be do-or-die events for Cortex.
That
CX717 was to be tested in such a variety of diseases was in part a
business strategy to give Cortex as many chances at success as possible.
The strategy also reflected Lynch’s belief that LTP was a fundamental
brain process. Whatever the cause of many neurological diseases, poor
communication between neurons was almost always one of the results.
The
first CX717 results from a small sleep-deprivation trial came back in
May. They were better than could have been hoped. The only drawback to
the trial was its size — just 16 men, who were deprived of a night’s
sleep, then given the drug and tested. Without the drug and without
sleep, their test scores fell off the chart. With the drug and without
sleep, they tested the same as they had when well-rested, and had none
of the jitters commonly associated with stimulants.
It was a clear win for the company. Its stock sank.
“Try
to lift the species out of the puddle of its own crap and what do you
get?” Lynch joked. The other trials had yet to get underway, and there
was too much going on to brood. The lab’s incredible run of success in
nailing down details of the biochemical processes underlying memory and
forgetfulness, begun that March, continued through the spring and
summer.
Lynch began experimenting with more potent versions of the
ampakines. Julie Lauterborn, working in the lab of Christine Gall,
Lynch’s longtime significant other and collaborator, had earlier
discovered that some ampakine variants increased production within the
brain of compounds known as neurotrophins.
These molecules, in
particular one known as brain-derived neurotrophic factor, or BDNF, were
essential to the maintenance of brain function. So many claims had been
made for BDNF over the years that Lynch tended to disregard them. He
mocked it, referring to BDNF as the “big-deal growth factor.”
“That’s the story — everything that’s wrong with the brain, BDNF will fix it,” Lynch said. “To me, it was ludicrous.”
But
when Kramar in his lab confirmed other reports that BDNF played a
crucial role in the LTP process, Lynch began to examine the interaction
between ampakines and BDNF.
He was astonished to learn that
particular ampakines could be used almost as a switch to turn on BDNF
production and thereby boost LTP.
There had been a long history of
attempts to somehow get more BDNF delivered to the brain. A company in
San Diego had gone so far as to drill holes in heads and pump BDNF
directly in. To have found a simple, apparently painless and yet
powerful means to turn on BDNF production would be like discovering a
magic potion.
Lynch was convinced that many neurological diseases —
Alzheimer’s, Huntington’s, Parkinson’s — were in part caused by the
normal wear and tear that accompanied aging.
Brain cells, unlike most
of the cells in the body — or most of the cells in most of living
creatures in all the known world — were more or less permanent. They
did not die and get replaced by new cells. They lived, for a century if
their host did, and accumulated all the damage anything that old might
expect.
The combination of aging and specific diseases, some of them
genetic in origin, led to mental difficulties — memory loss among them,
Lynch thought. Ampakines were supposed to help ameliorate many brain
diseases by bulldozing through the problems the diseases created,
compensating for aging. The drugs wouldn’t cure the diseases, but would
relieve the most debilitating symptoms.
Tools of the trade
The
direction of science is largely determined by the tools available to
pursue it. Problems arise when a tool dictates direction — an
illustration of the axiom that if you only have a hammer, every problem
looks like a nail.
Grant-makers and journal editors demand that you
use the popular tools to gain their approval. Tenure committees make
decisions based on grants and publications. In very short order,
research becomes normative. Interlopers are shunned, and risk-taking is
constrained.
“If you’ve got a method that lets you look at something,
then the answer must be what that method allows you to see,” Lynch
said.
In Lynch’s view, an unsightly number of neuroscientists have
been swinging hammers at a problem — memory — that didn’t look to him
like a nail. The field was enthralled by the tools of molecular biology
and their ability to manipulate genes in experimental animals.
It was
routine to reverse-engineer laboratory mice or rats — knocking genes
in or out of the animals — so that they had or lacked certain
qualities. In this way, the animals could mimic specific disease states.
There were rats with Huntington’s disease, Alzheimer’s, Parkinson’s.
One
problem in using the animals for neuroscience was the complexity of the
human brain, which in many of its actions was redundant. If one gene
was knocked out, eliminating the protein that gene manufactured, other
genes might make compensating proteins. It was often impossible to
delineate precisely what caused what.
“They’re nowhere near knowing
what the machine is, so they can’t know what the machine produces,”
Lynch said. “It’s like a 747 crash-landed in the jungle. The monkeys are
crawling all over it, having a hell of a time trying to figure out what
it is.”
Lynch realized he suddenly possessed new tools of his own:
Whatever the fate of the unpublished Kramar paper (which was eventually
published), the method it described to visualize the late stage of LTP
was an important new laboratory tool; and the ampakines themselves were a
tool that could probe the inner workings of LTP.
The outline of
Lynch’s LTP hypothesis was this: When you experienced a sensation in the
outside world — seeing, smelling or touching something — the
sensation was translated by the sensory organs into an electrical signal
that was routed to the brain, where it caused the brain cells, or
neurons, that received the stimulus to release chemicals to neighboring
neurons. A cascade of chemical events inside those neighboring neurons
resulted in their interior reorganization. That reorganization
strengthened the connection between cells at the points where they meet,
called the synapses. Networks of those neurons with strengthened
connections constituted the underpinning of memory.
In a normal LTP
experiment, a slice of a rat’s brain was subjected to a precisely timed
and measured electrical stimulus, mimicking the electrical signal
produced by a real-world sensory stimulus. The experimenter measured the
strength of the electric signal as it traveled through the slice. If a
larger-than-usual signal exited the slice, that meant LTP had occurred
and connections between the neurons in the slice had been strengthened.
All you really needed for these experiments were a microscope, a
chemical catalog and a pair of electrodes. That didn’t mean they were
easy, just that the tools to do them were straightforward.
Using her
visualization method and the ampakines, Kramar examined the BDNF-LTP
interaction. Over the course of months, what emerged was a picture that
was at once immensely complicated and impossibly elegant.
“Endogenous
BDNF does everything they said. It’s all true. It’s all true,” Lynch
said. “But I was too lazy to read the papers carefully.”
BDNF, Lynch
now thought, was crucial to the physical restructuring inside a neuron
during LTP. In essence, it was on the “on” switch. Lynch thought another
molecule, adenosine, was the “off” switch. A fine balance of the two
was needed for the brain to work. Through the fall, Lynch and company
elaborated on this line of thinking.
Chris Rex, a grad student who
had discovered LTP deficits in middle-aged rats, set up an experiment to
see what would happen if he used ampakines to instigate BDNF production
in the same rats.
It worked. The ampakine turned on the BDNF, and
the BDNF promoted LTP. The age-related deficit disappeared. As Lynch put
it later: “Middle-aged aging cured.”
Lynch, at Gall’s persistent
urging, began to reintegrate himself into the wider world of science. He
accepted invitations to speak at a few conferences.
The progress was
not without drama, some of it self-inflicted. Lynch continued his war
with the university administration. Finally, feeling he wasn’t being
treated with sufficient respect, he shut down his lab at 101 Theory
Drive, dispersing his researchers to Gall’s lab. Some of the scientists
were by then — or would soon be — gone for good. Ted Yanagihara went
to med school in New York. Laura Colgin left on a postdoc to Europe. The
lab’s computer expert left for private industry.
Lynch’s health
teetered between bad and calamitous. He showed up at a conference in
Vancouver fevered and shot full of antibiotics. He gave his lecture,
listened to others and made small talk, all the while looking to be on
the verge of collapse. “Just to have one damned thing that works,” he
said.
His various maladies never seemed to go away. The problems he’d
had with balance didn’t completely resolve. His neurologist held to the
initial diagnosis — a viral infection. Lynch suspected something more
serious but never pursued another diagnosis. “I put that into the
‘Forget About It’ file,” he said. “Don’t have time to worry about it.”
Through
it all, he continued to publish. In a universe where most papers are
written in a combination of dense chemical symbolism and genre jargon,
Lynch’s work stood out for its sometimes whimsical, often literary tone.
Some examples: “Consolidation: A View From the Synapse,” “Long-Term
Potentiation in the Eocene,” “Spandrels of the Night?” and “Ampakines
and the Three-Fold Path to Cognitive Enhancement.”
Not that this
endeared Lynch to everyone in the field. In addition to the Kramar
paper, he began to have others rejected at a rate he had never
experienced. Said Lynch: “Uneducated reviewers to the left, pygmies to
right, but on came the army of science.”
The army’s accumulating
evidence was producing a rich portrait of the LTP process that seemed to
Lynch to have far-reaching implications. LTP seemed to be a fundamental
brain process, perhaps the fundamental brain process.
The end of the artifact
Adisquieting
aspect of LTP research had long been that Lynch, his colleagues and
thousands of other scientists had devoted decades of research to it,
attempting to describe its details, yet all the while there was no
assurance it had anything to do with memory. They hoped it did. Some
believed it did. None of them knew.
Kandel, the godfather himself,
said he was far from convinced LTP had real-world significance. Lynch,
in the early years of LTP research, seldom used the word memory to
describe its relevance. He said it was a presumed “substrate of
behavioral plasticity,” a phrase nearly perfect in its obfuscation.
The
great final task for Lynch would be linking LTP unequivocally to
learning and memory. There was a chance he had spent 30 years chasing a
mere curiosity, “an interesting little piece of biology,” as he put it.
This worried Lynch; it scared the hell out of him.
Lynch
determined he might be able to use the new visualization method to do
an experiment that could show without question that LTP was memory.
The
visualization technique involved staining structural molecules inside
neurons so that their reorganization could be seen. The reorganization
occurred only on the extensions of the neurons known as dendritic
spines. If particular spines had no reorganization, the dye would wash
right through without sticking. It would only stick to the spines where
LTP had occurred.
Lynch’s lab had introduced and used the technique
in normal LTP bench experiments. Lynch wanted to try it in actual
animals. Rats would be trained in a new task, something that would be
encoded into memory. Then the dye would be injected into the rat’s
brain.
If it worked, the staining pattern would illuminate a neural
trace of memory. This was not a modest undertaking. Scientists for more
than a century had been trying to find such traces, often referred to as
engrams.
For a time in the mid-20th century, the search had been a
furious chase. It had since been all but abandoned as a sort of pipe
dream.
His goal, Lynch said, was to see “spines encode memory in real
rats after real learning . . . maps of memory encoding sites in the
brain. . . . The crowd goes wild!!!”
The crowd, such as it was, would need patience. It took nearly half a year to get the experiment up and running.
By
fall, the lab had demonstrated that the technique could work in a rat
learning to navigate a new environment. The next test would be to
compare the brains of rats that had been allowed to roam free with rats
kept in cages. Presumably, the roaming rats would have learned and
remembered something of their environment and converted this learning
into memories that would have caused more spines to restructure.
After
the roaming period, the rats would be sacrificed and the brains of the
roaming and the caged rats examined. If successful, it would confirm
that the method worked in living animals, and the team could proceed to
do specific learning and memory experiments. Lynch thought that might
take a few days. Days turned to weeks, weeks to months.
Given that Lynch had been at this for three decades, a few months hardly mattered, unless you were living them.
Then,
on a Saturday evening in early spring, Kramar was at the microscope
when she saw it — an actual trace of an actual memory. She started
screaming.
“Gary was in the bathroom. I was so excited I almost ran in there to get him,” Kramar said.
Now
that he knew the method worked, Lynch gathered the group the next
Monday to prepare for the final push. He had decided the rats would
learn pairs of odors such as lemon-orange and strawberry-peppermint.
“The
olfactory advantage is that we understand where odor memory is encoded.
We know a priori that if the animal learns, it has to be here. In the
visual cortex, you don’t know where to start,” he said.
Vadim
Fedulov, a graduate student, was assigned to run the rats in a maze in
the odor-learning experiment. He had been in the lab for just a couple
months. Lynch had agreed to take him on when it looked as if he might be
tossed out of grad school altogether. He was young, very bright,
somewhat unpredictable and not punctual at all. He was, in other words, a
typical Lynch recruit.
One morning not long after, before the
markets opened, Cortex announced the results of a clinical trial in
which the ampakine CX717 had been given to adults diagnosed with ADHD.
The results were an unqualified success. The drug reduced ADHD symptoms
across the board, almost equal to existing medications — mainly
stimulants — without any of their deleterious side effects.
ADHD
affects an estimated 4% of children in the United States. More than 30
million prescriptions are written for the disorder annually.
Cortex
stock doubled in value over the next week. Chief Executive Roger Stoll
announced the company was in negotiations with at least eight big
pharmaceutical companies that wanted to license CX717. Such a deal,
Stoll said, would be worth immediately as much as $30 million to the
company and eventually several hundred million dollars.
Lynch was
ecstatic. It was the first big public demonstration of the power of the
ampakines. This was a day he had waited 15 years for.
“It’s immensely
gratifying. It really is,” he said. “It validates the principle that
you can treat neurological diseases by increasing cortical
communication.”
Lynch, feeling magnanimous, patched up his
relationship with the university and moved back to 101 Theory. Kramar
received a job offer from an Irvine biotech company. Although she hated
the timing, in tears, she took the job.
Danielle Simmons took over
her role in the engram search. Fedulov built his T-maze — which was
outfitted with sliding doors and flashing lights and apertures through
which he inserted cotton swabs soaked in various scents — and began
running rats.
Lynch was scheduled to speak at a pharmaceutical conference in San Francisco. He prepared to be welcomed as a conquering hero.
Two
days before the conference, the FDA called Cortex and said it had found
unspecified problems in preclinical results — that is, lab tests on
animals — with CX717. It ordered an immediate halt to all human trials.
Monday
morning, as Lynch was scheduled to speak, the clinical hold was
announced. Lynch had a bronchial infection and was loaded with
antibiotics and steroids; his plane was two hours late because of bad
weather. Cortex stock fell 60%. The subject of Lynch’s talk was the
failure of translation from preclinical lab work to clinical trials in
memory drugs.
Irony wasn’t quite strong enough to describe the circumstances.
Lynch
had no objection to the FDA’s action, even though he thought the hold
would be resolved painlessly. “They’re doing what they have to do,” he
said. “We’re putting stuff in people’s brains, and they should be
careful.”
Lynch returned to Irvine, the lab and the odor-learning experiment.
Fedulov
had trained three rats and was ready to inject the dye and have the
brains prepared for examination. Two of the rats, for reasons unknown,
died from the injections. The sole remaining rat was sacrificed, its
brain sliced and set on slides. Fedulov had the slides at 101 Theory.
Lynch wanted to look at them at Gall Lab, where Lauterborn, an expert
microscopist, could read and photograph the images. He called Fedulov
and asked that he bring the slides.
Tracey Shors, a neuropsychologist
from Rutgers who early in her career had written a much-discussed paper
casting a skeptical eye toward the role of LTP in memory, happened to
be on the UCI campus. She and Lynch were old friends and met to discuss
their various researches. Lynch told her about the afternoon’s
prospects; she was interested, but skeptical. Bring me an engram, she
said. Bring me an engram.
Lynch went to lunch, just about dying from
anxiety. When he returned, Fedulov was nowhere to be found. There was
Lynch, the big experimental result waiting — no rat brains, no
scientists.
“You’d think I was trying to launch the space shuttle,” he said.
Lynch
called Fedulov on his cell. He’d come and gone and left the slides in a
refrigerator, neglecting to tell anyone. To further darken the
atmosphere, a short paper Lynch had written describing the preliminary
results of this work had come back from a journal editor, declined with
scathing reviews.
Fedulov finally showed up, retrieved the slides and
mounted the first in the microscope. One of the great difficulties in
finding the markings of memory, assuming you even knew what they looked
like, was knowing where to look. The brain, at the nano-scale, was a
very big place. Even a rat has hundreds of billions of synapses. USC
neuropsychologist Thompson has estimated that the human brain, with as
many as 10 quadrillion synapses, is capable of more distinct neural
patterns — memories — than there are atoms in the universe. The reason
Lynch had chosen an odor-learning exercise for this experiment was
because the olfactory cortex was the simplest system to navigate.
Lauterborn
searched through the slides, found the olfactory cortex and moved aside
to let Lynch take a look. He immediately began oohing and ahhing. “Oh.
Oh. Yes, yes, they could.” He continued to scan across the slides. “Yes,
yes, yes, YES. I’m almost convinced. Almost.”
Lauterborn took
digital photographs of the slides through the computer and brought up
the images for more-detailed examination. The slides were a guided tour
through the brain from the olfactory bulb to the cortex to the
hippocampus, to the region thought to control emotion, the amygdala.
Lynch
scanned across the brain regions. “That’s gorgeous. . . . That’s the
way the world’s supposed to be. Look at that. . . . All this time, that
is the picture I wanted to see. Right there.”
The following week they replicated the experiment with four more rats.
Lynch
was again at the microscope. The image on the monitor showed a vast
gray field of brain matter. The gray was lighted up here and there
sparsely, but intensely.
“You see ’em. You see ’em. Look at them,” he
said. He traced on a piece of paper the path from the nose to the point
on the image in the olfactory cortex being viewed. “It’s a direct
connection between the olfactory cortex and the hippocampus [the memory
center of the brain]. Four synapses from the nose to the hippocampus.”
Back
at the microscope, he followed the path through to the hippocampus and
murmured, “Vadim, Vadim, I’m going to make you famous. . . . That’s it.
That is the first demonstration that LTP is engaged in memory.”
Is memory? I asked.
“Is memory,” he said. “It couldn’t have been much better. All those years, all those arguments — it’s all gone.”
He
went back to the microscope. After a minute or so of further scanning
and examination, he shouted, “You see that, boys and girls? Science
works.”
A theory of a lot
One day not long after, Arvid Carlsson, a
Nobel laureate pharmacologist from Sweden, visited. Lynch briefed him
on the engram experiments. Carlsson was enthralled.
“To me, it seems
so absolutely surprising and convincing,” he said. “It makes so much
sense. It seems to be a fundamental discovery.”
The approbation of a
learned and widely respected old hand like Carlsson was gratifying to
Lynch, but to a surprising extent he had moved beyond needing it. The
science had yielded. All he wanted now was more.
As the results
continued to come in, he got it. An unrelenting problem in memory
research for decades had been determining the location of long-term
memory storage. “One of the darkest areas of research,” Alcino J. Silva
of UCLA called it. “We know nothing about it, literally nothing.”
Researchers
almost unanimously agreed that the hippocampus played a fundamental
role in acquiring new memories, but they seemed to be moved later to the
cortex for storage. If that were true, who or what moved it?
Because
Lynch Lab could now see actual memory traces, the scientists were able
to plot them onto brain maps. The maps showed how complicated a
phenomenon even a simple memory was and the degree to which a single
memory trace was distributed across brain regions. When the rats learned
new smells, the trace of the learning landed in the olfactory bulb,
then in the cortex, the hippocampus, the amygdala.
In a single
stroke, Lynch’s LTP research seemed to have yielded a realistic
hypothesis for long-term storage. The memory wasn’t moved from the
hippocampus but was encoded simultaneously in both the hippocampus and
the cortex. Presumably, a signal could later be sent from the
hippocampus to keep or discard that particular memory.
Last October,
the FDA lifted its hold on ampakine clinical trials, but it imposed a
dose limit for patients, and the trials didn’t restart until the limits
were lifted this July. In the meantime, other researchers around the
globe began experimenting with ampakines in a variety of indications,
ranging from breathing disorders to mental retardation.
Rex and Lulu
Chen, a new graduate student, devised an alternative means for the
visualization experiments. The new method was easier to execute and
produced stunning, unambiguous, easily replicable results. Publication
of their work early this year provided the validation Lynch had
anticipated almost two years earlier.
Offers — pleas, even — to
collaborate rained in from around the globe. There was talk with the
National Institutes of Health about setting up an engram project, a sort
of national memory- mapping program to extend Lynch’s work.
Lynch
took gleeful satisfaction in the way in which the lab’s methods —
described disdainfully as producing results that were “simply not
credible” and contradictory to “all the assembled cell biological
knowledge” a year before — suddenly seemed poised to become standard
practice in dozens of labs.
While declaring almost daily that he was
about to quit the whole enterprise, Lynch couldn’t help himself. The
experiments seemed to provide interesting insights into Huntington’s,
retardation and, out of left field, menopause. Kramar, bored with the
pace of work in private industry, returned to the lab. Hypotheses were
being born by the dozen, and Lynch began planning the next expeditions
onto the reedy shores of the unknown.
Lynch had in his career tried
hard, he thought, even if no one shared the opinion, to remain modest,
shying from big ideas and theories. He had thought it vain to suppose he
could formulate an overarching explanation of memory and cognition. Now
he was greedy. He wanted to shake the entire apparatus and demand all
its secrets fall to the ground.
A physicist, when grandiose, will
talk about forming a TOE — a Theory of Everything. Lynch wasn’t ready
quite for that. This was still biology, after all. He would settle for a
TOAL — Theory of a Lot.
“I need to solidify the breakthrough — get
my arms around how big the thing is, how much of brain biology can be
folded into it,” he said.
The brain, he thought, represented a
devil’s bargain. You get memory storage almost beyond measure, but
because that memory required more or less permanent neurons, you could
not routinely replace them with new cells. If neurons broke, and they
do, you were stuck with the results. As he so bluntly put it: You get
stupid. And, because the brain controls so much of what the body does,
when neurons fail you lose much more than memory.
“The evolutionary
idea of stability as the cost of memory fascinates me,” Lynch said. “And
maybe, at the end, there lies the answer for how to get my broken-down
brain going again.”
Most of the work by this time was being done in
Gall’s lab, a few hundred yards away from 101 Theory, where, some days,
Lynch worked alone.
It seemed fitting, somehow. There he sat at the
end of the great, long chase, often sick as a dog, the entry locked, the
clamorous tribes of the neurosciences a low hum in the distance; no
phone, no e-mail, not even a name on the door to betray his presence.
The only way you would know he was there at all was the blue Corvette
out front. And, of course, the science, which, no matter the
circumstance, difficulty or hour, had poured out for 30 years like water
from the well. And poured still.