Friday, August 27, 2010

Don’t throw away the cane

In Chapter 1 of the 4th edition of Intermediate Physics for Medicine and Biology, Russ Hobbie and I quote the article “Don’t Throw Away the Cane” by Walter Blount (Journal of Bone and Joint Surgery, Volume 38, Pages 695–708, 1956).
The patient with a wise orthopedic surgeon walks with crutches for six months after a fracture of the neck of the femur. He uses a stick for a longer time—the wiser the doctor, the longer the time. If his medical adviser, his physical therapist, his friends, and his pride finally drive him to abandon the cane while he still needs one, he limps. He limps in a subconscious effort to reduce the strain on the weakened hip. If there is restricted motion, he cannot shift his body weight, but he hurries to remove the weight from the painful hip joint when his pride makes him reduce the limp to a minimum. The excessive force pressing on the aging hip takes its toll in producing degenerative changes. He should not have thrown away the stick.
I recently looked up the full article, which is delightful. It was listed as a JBJS classic in 2003 (Volume 85, Page 380). Here are a few more quotes. I suggest you read the entire paper.
As the causes of premature death are conquered one by one, man is given a longer life in which to grow old gracefully. In the twilight years that his forefathers rarely knew, he needs help in seeing, hearing, chewing, and walking. Gradually we are coming to look upon eye glasses, hearing devices, and dentures as welcome aids to gracious living rather than as the stigmata of senility. They should be accepted eagerly as components of a richer life. The cane, too, should be restored to favor as a means of preventing fatigue and a halting gait, rather than maligned as a sign of deterioration.

The use of the cane in order to prevent strain upon an ailing hip or knee is not generally accepted. In the patient's mind there is a nice distinction between the permissible use of a stick postoperatively and the adoption of this humble support for no other reason than the relief of a slight physical infirmity. A fat lady may waddle like a duck when she laboriously walks a few steps, but she resents the suggestion that she carry a cane. She would look much better with a stick than with the limp; and with support she could walk enough to get some exercise. More walking would help with weight reduction. But no! she is not ready for a cane yet! The patient with residual disability after poliomyelitis and with a fatiguing, unsightly lurch needs a cane. Early degenerative hip disease may require no treatment other than weight reduction and a stick in the opposite hand; however it takes an impressive orthopaedic surgeon to sell the idea…

As Pauwels has shown so well (Fig. 8), the use of a cane in the left hand reduces the pressure on the right femoral head without the need for limping. The support afforded by the stick greatly lessens the pull required of the abductor muscles in helping to support the body weight. The cane works through a long lever, so that a moderate push on the stick greatly relieves the strain on the hip [my boldface]. The relative forces are shown in Table I. Pauwels estimated that during the stance phase of walking, without the support of a cane, an average person exerts a static force of 385 pounds on the stationary hip. This weight can be reduced to 220 pounds by pushing down on a stick with the opposite hand the equivalent of 20 pounds. The cane is really an efficient mechanical device…

I should rather be remembered as a thoughtful surgeon than as a bold one. I submit that a well planned sequence of lesser operations with long intervals between, and the use of a cane as needed, may prove better for the patient and productive of a more desirable end result than some more heroic surgical procedure. There is a tendency among orthopaedic surgeons to exchange simple methods for dramatic treatment that will not require the use of the cane. The surgeon looks for a single, definitive, bridge-burning operation that will cure the patient completely for the rest of his life. Too often, this goal is not reached. The patient still needs the stick (or even crutches) after this heroic operation. If a satisfactory arthroplasty or reconstruction operation is performed, how much better it would be for most patients to urge the continued use of a cane in order to preserve the function of the reshaped bone by taking the strain off the hip for years, not for months only.
Blount was a leading physician and surgeon in orthopedics. His grandfather was a civil war surgeon, his mother was a physician and surgeon, and his sister was a pediatrician. He attended the University of Illinois and Rush Medical College. He helped develop the Milwaukee brace for spinal malalignment, was an expert on fractures in children, and introduced tibial stapling for epiphyses. In 1954 he became president of the American Academy of Orthopedic Surgeons.

Friday, August 20, 2010

The Anger Camera

In Section 17.12 of the 4th edition of Intermediate Physics for Medicine and Biology, Russ Hobbie and I discuss the gamma camera, used in to produce images during nuclear medicine procedures. The gamma camera often goes by another name, the Anger camera, after its inventor Hal Anger (1920–2005). Anger’s contributions to medical physics imaging rank him with giants of the field such as Godfrey Hounsfield, Allan Cormack, and Paul Lauterbur. Yet, it is harder to find information about Anger than about these other luminaries. His obituary in the New York Times stated
Mr. Anger was known in his field for inventing the gamma camera, which was first exhibited in 1958 at a meeting of the Society of Nuclear Medicine. The device, also known as a scintillation camera and later as the Anger camera, produced images of internal processes by tracking tiny amounts of radioactive substances, known as radiopharmaceuticals, given to patients.

The invention and later improvements represented a major advance in the diagnosis and treatment of brain tumors, bone marrow disorders and other life-threatening diseases.
Another obituary in the IEEE publication The Institute wrote
Seen by many as a quiet genius who shaped the future of nuclear medicine, Hal took a hands-on approach to science that also led to his invention of the well counter, which is used daily in nuclear medical labs worldwide to measure small quantities of radioactive substances. He also invented the whole-body scanner, the positron camera, and the multiplane tomographic scanner.

Nuclear medicine has been profoundly affected by Hal Anger. Millions of patients have benefited from diagnosis and treatment that depended on the Anger camera and the innovations made possible by its development.
Anger described his invention in a paper titled simply “Scintillation Camera” (Review of Scientific Instruments, Volume 29, Pages 27–33, 1958). The abstract is reproduced below.
A new and more sensitive gamma-ray camera for visualizing sources of radioactivity is described. It consists of a lead shield with a pinhole aperture, a scintillating crystal within the shield viewed by a bank of seven photomultiplier tubes, a signal matrix circuit, a pulse-height selector, and a cathode-ray oscilloscope. Scintillations that fall in a certain range of brightness, such as the photopeak scintillations from a gamma-ray-emitting isotope, are reproduced as point flashes of light on the cathode-ray tube screen in approximately the same relative positions as the original scintillations in the crystal. A time exposure of the screen is taken with an oscilloscope camera, during which time a gamma-ray image of the subject is formed from the flashes that occur. One of many medical and industrial uses is described, namely the visualization of the thyroid gland with I131.

Friday, August 13, 2010

The barn

Figure 15.2 of the 4th edition of Intermediate Physics for Medicine and Biology shows the cross section for the interaction of photons with carbon versus photon energy. The caption of the figure says “The cross section is given in barns: 1 b = 10−28 m2.” Where did this strange unit come from?

The July 1972 issue of Physics Today published a letter by M. G. Holloway and C. P Baker, explaining “How the Barn was Born.”
Some time in December of 1942, the authors, being hungry and deprived temporarily of domestic cooking, were eating dinner in the cafeteria of the Union Building of Purdue University. With cigarettes and coffee the conversation turned to the topic uppermost in their minds, namely cross sections. In the course of the conversation, it was lamented that there was no name for the unit of cross section of 10−24 cm2. It was natural to try to remedy this situation.

The tradition of naming a unit after some great man closely associated with the field ran into difficulties since no such person could be brought to mind. Failing in this, the names Oppenheimer and Bethe were tried, since these men had suggested and made possible the work on the problem with which the Purdue project was concerned. The “Oppenheimer” was discarded because of its length, although in retrospect an “Oppy” or “Oppie” would seem to be short enough. The “Bethe” was thought to lend itself to confusion because of the widespread use of the Greek letter. Since John Manley was directing the work at Purdue, his name was tried, but the “Manley” was through to be too long. The “John” was considered, but was discarded because of the use of the term for purposes other than as the name of a person. The rural background of one of the authors then led to the bridging of the gap between the “John” and the “barn.” This immediately seemed good and further it was pointed out that a cross section of 10−24 cm2 for nuclear processes was really as big as a barn. Such was the birth of the “barn.”

To the best knowledge of the authors, the first public (if it may be called that) use of the barn was in Report LAMS-2 (28 June, 1943) in which the barn was defined as a cross section of 1 × 10−24 cm2.

The authors would like to insist that the “barn” is spelled just that way, that no capital “b” is needed, and that the plural is “barns” with no letter “e” involved, and that the symbol be a small “b.” The meanings of “millibarn” and “kilobarn” are obvious.

Friday, August 6, 2010

Iron, Nature’s Universal Element

Iron, Nature's Universal Element:  Why People Need Iron and Animals Make Magnets,  by Eugenie Mielczarek, superimposed on Intermediate Physics for Medicine and Biology.
Iron, Nature's Universal Element:
Why People Need Iron
and Animals Make Magnets
,
by Eugenie Mielczarek.
A few months ago in this blog, I mentioned that I put the book Iron, Nature’s Universal Element: Why People Need Iron and Animals Make Magnets, by Eugenie Mielczarek and Sharon McGrayne, on my list of things to read this summer. Well, I finished this book, and I recommend it. Russ Hobbie and I cite the book in Chapter 8 of the 4th edition of Intermediate Physics for Medicine and Biology.
Magnetism is used for orientation by several organisms. A history of studies in this area is provided in a very readable book by Mielczarek and McGrayne (2000).
My favorite part of Iron, Nature’s Universal Element was Chapter 4, on magnetotactic bacteria. Russ and I discussed these interesting little creatures in Section 8.8.3
Several species of bacteria contain linear strings of up to 20 particles of magnetite, each about 50 nm on a side encased in a mambrane [Frankel et al. (1979); Moskowitz (1995)]. Over a dozen different bacteria have been indentified that synthesize these intracellular, membrane-bound particles or magnetosomes (Fig. 8.25). Bacteria in the northern hemisphere have been shown to seek the north pole. Because of the tilt of the earth’s field, they burrow deeper into the environment in which they live. Similar bacteria in the southern hemisphere burrow down by seeking the south pole. In the laboratory the bacteria align themselves with the local field.
The caption to our Figure 8.25 reads
The small black dots are magnetosomes, small particles of magnetite in the magnetotactic bacterium Aquaspirillum magnetotacticum. The vertical bar is 1 [micron] long. The photograph was taken by Y. Gorby and was supplied by N. Blakemore and R. Blakemore, University of New Hampshire.
Mielczarek and McGrayne provide a colorful description of how magnetotactic bacteria were discovered.
“I see it crystal clearly,” Richard Blakemore said, recalling the evening he discovered Earth’s smallest living magnets. “I get excited every time I look at them.”

It was already dark outside the laboratory as Blakemore, peering through his microscope, searched through mud samples for bacteria. At twenty-three, Blakemore was a second-year graduate student in microbiology at the University of Massachusetts in Amherst. In 1975, fledgling microbiologists there were often assigned such simple tasks as identifying the material between their teeth or analyzing organisms in mud. His professor had collected the mud from a Massachusetts marsh, and asked Blakemore to learn everything possible about some large spiral bacteria in it. But that night, Blakemore said, “other organisms forced their existence on me...”

So while Blakemore looked through the scope, [advanced graduate student John Bresnick] picked up a magnetic stirrer lying beside the microscope and brought it up behind the swimmers. “Fortunately,” Blakemore recalled, “he had the end of the magnet pointing toward them so that it attracted them. And—all of a sudden—en masse—this whole massive population of bacteria swims in exactly the opposite way across the microscope stage. It was incredible, just incredible, and no one even believed my response. They thought I was kidding—until they looked in...” At that point, John Bresnick said, “I think you’ve discovered something...” “From then on,” Blakemore said, still starry-eyed more than twenty years later, “it was a night of incredulity...”

Blakemore’s microbiology professor was in Italy at the time, and Blakemore was exploding with the news, so he raced home to tell his wife Nancy. Abandoning all grammar in the joy of the memory, Blakemore said “It couldn’t have been perfecter. I didn’t really—hardly—know how to take it in.”

Friday, July 30, 2010

X-ray Crystallography

Two weeks ago in this blog, when reviewing Judson’s excellent book The Eighth Day of Creation, I wrote that X-ray crystallography played a central role in the development of molecular biology. But Russ Hobbie and I do not discuss X-ray crystallography in the 4th edition of Intermediate Physics for Medicine and Biology, even though it is a classic example of physics applied in the biomedical sciences. Why? I think one of the reasons for this is that Russ and I made the conscious decision to avoid molecular biophysics. In our preface we write
Biophysics is a very broad subject. Nearly every branch of physics has something to contribute, and the boundaries between physics and engineering are blurred. Each chapter could be much longer; we have attempted to provide the essential physical tools. Molecular biophysics has been almost completely ignored: excellent texts already exist, and this is not our area of expertise. This book has become long enough.
Nevertheless, sometimes—to amuse myself—I play a little game. I say to myself “Brad, suppose someone pointed a gun to your head and demanded that you MUST include X-ray crystallography in the next edition of Intermediate Physics for Medicine and Biology. Where would you put it?”

My first inclination would be to choose Chapters 15 and 16, about how X-rays interact with tissue and their use in medicine, which seems a natural place because crystallography involves X-rays. Yet, these two chapters deal mainly with the particle properties of X-rays, whereas crystallography arises from their wave properties. Also, Chapters 15 and 16 make a coherent, self-contained story about X-rays in medical physics for imaging and therapy, and a digression on crystallography would be out of place. An alternative is Chapter 14 about Atoms and Light. This is a better choice, but the chapter is already long, and it does not discuss electromagnetic waves with wavelengths shorter than those in the ultraviolet part of the spectrum. Chapter 12 on Images is another possibility, as crystallography uses X-rays to produce an image at the molecular level based on a complicated mathematical algorithm, much like tomography uses X-rays to predict an image at the level of the whole body. Nevertheless, if that frightening gun were held to my head, I believe I would put the section on X-ray crystallography in Chapter 11, which discusses Fourier analysis. It would look something like this:
11.6 ½ X-ray Crystallography

One application of the Fourier series and power spectrum is in X-ray crystallography, where the goal is to determine the structure of a molecule. The method begins by forming a crystal of the molecule, with the crystal lattice providing the periodicity required for the Fourier series. DNA and some proteins form nice crystals, and their structures were determined decades ago.* Other proteins, such as those that are incorporated into the cell membrane, are harder to crystallize, and have been studied only more recently, if at all (for instance, see the discussion of the potassium ion channel in Sec. 9.7).

X-rays have a short wavelength (on the order of Angstroms), but not short enough to form an image of a molecule directly, like one would obtain using a light microscope to image a cell. Instead, the image is formed by diffraction. X-rays are an electromagnetic wave consisting of oscillating electric and magnetic fields (see Chapter 14). When an X-ray beam is incident on a crystal, some of these oscillations add in phase, and the resulting constructive interference produces high amplitude X-rays that are emitted (diffracted) in some discrete directions but not others. This diffraction pattern (sometimes called the structure factor, F) depends on the wavelength of the X-ray and the direction (see Prob. 19 2/3). One useful result from electromagnetic theory is that the structure factor is related to the Fourier series of the electron density of the molecule: F is just the an and bn coefficients introduced in the previous three sections, extended to account for three dimensions. Therefore, the electron density (and thus the molecular structure) can be determined if the structure factor is known.

A fundamental limitation of X-ray crystallography is that the crystallographer does not measure F, but instead detects the intensity |F|2. To understand this better, recall that the Fourier series consists of a sum of both cosines (the an coefficients) and sines (bn). You can always write the sum of a sine and cosine having the same frequency as a single sine with an amplitude cn and phase dn (See Prob. 19 1/3)

an cos(ωn t) + bn sin(ωn t) = cn sin(ωn t + dn) . (1)

The measured intensity is then cn2. In other words, an X-ray crystallography experiment allows you to determine cn, but not dn. Put in still another way, the experiment measures the power spectrum only, not the phase. Yet, in order to do the Fourier reconstruction, phase information is required. How to obtain this information is known as the “phase problem,” and is at the heart of crystallographic methods. One way to solve the phase problem is to measure the diffraction pattern with and without a heavy atom (such as mercury) attached to the molecule: some phase information can be obtained from the difference of the two patterns (Campbell and Dwek (1984)). In order for this method to work, the molecule must have the same shape with and without the attached heavy atom present.

* for a fascinating history of these developments, see Judson (1979)

Problem 19 1/3 Use the trigonometric identity sin(A+B) = sinA cosB + cosA sinB to relate an and bn in Eq. (1) to cn and dn.

Problem 19 2/3 Bragg’s law can be found by assuming that the incident X-rays (having wavelength λ) reflect off a plane formed by the regular array of points in the lattice. Assume that two adjacent planes are separated by a distance d, and that the incident X-ray bean falls on this plane at an angle θ with respect to the surface. The condition for constructive interference is that the path difference between reflections from the two planes is an integral multiple of λ. Derive Bragg’s law relating θ, λ and d.

Campbell, I. D., and R. A. Dwek (1984) Biological Spectroscopy. Menlo Park, CA, Benjamin/Cummings.

Judson, H. F. (1979) The Eighth Day of Creation. Touchstone Books
For more information on X-ray crystallography, see http://www.ruppweb.org/Xray/101index.html or http://www.xtal.iqfr.csic.es/Cristalografia/index-en.html.

Friday, July 23, 2010

AAPT Summer Meeting in Portland Oregon

On Tuesday, Russ Hobbie gave a talk about “Medical Physics in the Introductory Physics Course” at the American Association of Physics Teachers Summer Meeting in Portland Oregon. His session, with over 100 people attending, focused on Reforming the Introductory Physics Courses for Life Science Majors, a topic currently of great interest and one that I have discussed before in this blog. You can find the slides that accompanied his talk at the 4th edition of Intermediate Physics for Medicine and Biology website. His talk focused on five topics that he feels are crucial for the introductory course: 1) Exponential growth and decay, 2) Diffusion and solute transport, 3) Intracellular potentials and currents, 4) Action potentials and the electrocardiogram, and 5) Fitting exponentials and power laws to data. All these topics are covered in our book. Russ and I also compiled a list of topics for the premed physics course, and cross listed them to our book, this blog, and other sources. You can find the list on the book website, or download it here.

Our book website is a source of other important information. For instance, you can download the errata, containing a list of known errors in the 4th edition of Intermediate Physics for Medicine and Biology. You will find Russ’s American Journal of Physics paper “Physics Useful to a Medical Student” (Volume 43, Pages 121–132, 1975), and Russ and my American Journal of Physics Resource Letter MP-2: Medical Physics” (Volume 77, Pages 967–978, 2009). Other valuable items include MacDose, a computer program Russ developed to illustrate the interaction of radiation with matter, a link to a movie Russ filmed to demonstrate concepts related to the attenuation and absorption of x rays, sections from earlier editions of Intermediate Physics for Medicine and Biology that were not included in the 4th edition, and a link to the American Physical Society, Division of Biological Physics December 2006 Newsletter containing an interview with Russ upon the publication of the 4th edition of our book. You can even find a link to the Intermediate Physics for Medicine and Biology facebook group.

Russ and I hope that all this information on the book website, plus this blog, helps the reader of Intermediate Physics for Medicine and Biology keep up-to-date, and increases the usefulness of our book. If you have other suggestions about how we can make our website even more useful, please let us know. Of course, we thank all our dear readers for using our book.

Friday, July 16, 2010

The Eighth Day of Creation

The Eighth Day of Creation: The Makers of the Revolution in Biology, by Horace Freeland Judson, superimposed on Intermediate Physics for Medicine and Biology.
The Eighth Day of Creation:
The Makers of the Revolution in Biology,
by Horace Freeland Judson.
I recently finished reading The Eighth Day of Creation, a wonderful history of molecular biology by Horace Freeland Judson. The book is divided into three parts: 1) DNA—Function and Structure: the elucidation of the structure of deoxyribonucleic acid, the genetic material, 2) RNA—The Functions of the Structure: the breaking of the genetic code, the discovery of the messenger, and 3) Protein—Structure and Function: the solution of how protein molecules work. The first part centers on the story of how Watson and Crick discovered the double-helix structure of DNA, a story also told in Watson’s book The Double Helix (required reading for any would-be scientist). I was less familiar with the RNA tale in the second part, but was fascinated by the “Good Friday” meeting in which the various roles of RNA (as both a messenger taking the genetic information from DNA to the protein, and as part of ribosomes where protein synthesis takes place) was first understood by Sydney Brenner and Francois Jacob, among others. I was somewhat familiar with Kornberg’s deciphering of the genetic code from my days at the National Institutes of Health, where Kornberg worked. The last section tells how Max Perutz used X-ray crystallography to determine the structure of hemoglobin, the first protein structure known.

New to me was the story of Jacques Monod and his study of bacteria, which led to our understanding of how protein synthesis is controlled. Last week in this blog I mentioned seeing a display about Monod at the Pasteur museum in Paris. Particularly fascinating was the story of Monod’s role as a leader of the French resistance against the Nazis during World War II, and how he continued his scientific research while participating in the resistance. Judson writes
“In the autumn of 1943, a meeting was called in Geneva of representatives of all the armed groups of the French resistance, to coordinate their military actions. Just before the meeting, Philippe Monod heard from his brother that he was the delegate of the Francs-Tireurs from Paris. In November, the Gestapo arrested a minor agent of one of the main resistance networks in France, Reseau Velites, centered on the Ecole Normale Superieur. Marchal’s identity [Marchal was an alias used by Monod] and activities were known to the agent. Monod had to go underground completely, leaving his apartment, never sleeping more than a night or two at one address, staying away from the Sorbonne. On 14 February 1944, the Gestapo caught Raymond Croland, chief of the Reseau Velites, who knew Monod.

On the run from his own laboratory, Monod was given bench space by [Andre] Lwoff. “I don’t think I was ever searched for, actually,” he said. “But the possibility existed because at least one—in fact, several men had been picked up who knew what I was doing and who knew my name and where I worked. But it was known that I lived near the Sorbonne and worked at the Sorbonne, so the Gestapo would have had no reason to hunt for me at the Pasteur Institute.” In Lwoff’s laboratory, in collaboration with Alice Audureau, a graduate student, Monod that winter began a new set of experiments…
I would rank The Eighth Day of Creation second in my list of the best scientific histories I have read, just behind Richard RhodesThe Making of the Atomic Bomb, and just ahead of Bruce Hunt’s The Maxwellians. Interestingly, some of the characters who appeared in The Eighth Day of Creation also played a role in The Making of the Atomic Bomb: in particular, George Gamow and Leo Szilard (Szilard was mentioned in the very first sentence of Rhodes’ book). Readers of the 4th edition of Intermediate Physics for Medicine and Biology will be interested in learning that many of the pioneers in molecular biology were trained as physicists. Judson writes “new people came into biology, and most famously the physicists: Max Delbruck, Leo Szilard, Francis Crick, Maurice Wilkins, [and] on an eccentric orbit George Gamow.” I couldn’t help but be struck by the central role of X-ray crystallography in the history of molecular biology. Under physicist William Bragg’s leadership at the Cavendish, four Nobel prizes were awarded (in the same year, 1962) for molecular structure determination: Watson and Crick for DNA, and Perutz and Kendrew for the structure of hemoglobin and myoglobin. I highly recommend the book, especially for young biology students interested in the history of their subject.

I will end with the opening paragraphs of The Eighth Day of Creation, where Judson draws parallels between the revolutions in physics in the first decades of the 20th century and the revolution in biology in the middle of the century.
The sciences in our century, to be sure, have been marked almost wherever one looks by momentous discoveries, by extraordinary people, by upheavals of understanding—by a dynamism that deserves to be called permanent revolution. Twice, especially, since 1900, scientists and their ideas have generated a transformation so broad and so deep that it touches everyone’s most intimate sense of the nature of things. The first of these transformations was in physics, the second in biology. Between the two, we are most of us spontaneously more interested in the science of life; yet until now it is the history of the transformation of physics that has been told.

The revolution in physics came earlier. It began with quantum theory and the theory of relativity, with Max Planck and Albert Einstein, at the very opening of the century; it encompassed the interior of the atom and the structure of space and time; it ran through the settling of the modern form of quantum mechanics by about 1930. Most of what has happened in physics since then, at least until recently, has been the playing out of the great discoveries—and of the great underlying shift of view—of those three decades. The decades, that shift of view, the discoveries, and the men who made them are familiar presences, at least in the background, to most of us; after all, they built the form of the world as we now take it to be. The autobiographies of the major participants, their memoirs and philosophical reflections, have been composed, their biographies written in multiple—and they remain long in print, for these were men of intelligence, originality, and, often, eccentricity. The scientific papers have been scrutinized as historical and literary objects. The letters have been catalogued and published. The collaborations have been disentangled, the conferences reconvened on paper with vivid imaginative sympathy, the encounters, the conversations, even the accidents reconstructed.

The revolution in biology stands in contrast. Beginning in the mid thirties, its first phase, called molecular biology, came to a kind of conclusion—not an end, but a pause to regroup—by about 1970. A coherent if preliminary outline of the nature of life was put together in those decades. This science appeals to us very differently from physics. It directly informs our understanding of ourselves. Its mysteries once seemed dangerous and forbidden; its consequences promise to be practical, personal, urgent. At the same time, biology has been growing accessible to the general reader as it never was before and as the modern physics never can be. Indeed, part of the plausibility of molecular biology to the scientists themselves is that it is superbly easy to visualize. The nonspecialist can understand this science, at least in outline, as it really is—as the scientist imagines it. Yet the decades of these discoveries have hardly been touched by historians before now. The Eighth Day of Creation is a historical account of the chief discoveries of molecular biology, of how they came to be made, and of their makers—for these, also, though only two or three are yet widely known, were scientists, often of intelligence, originality, even eccentricity.

Friday, July 9, 2010

Paris

I just returned from a vacation in Paris, where my wife and I celebrated our 25th wedding anniversary. Russ Hobbie was there at the same time, although conflicting schedules did not allow us to get together. My daughter Katherine posted the blog entries for the last two weeks, when I had limited computer access. Thanks, Kathy.

Although most of our time was spent doing the usual tourist activities (for example, the Arc de Triomphe, the Notre Dame Cathedral, Versailles, and, my favorite, a dinner cruise down the Seine), I did keep my eye open for those aspects of France that might be of interest to readers of the 4th edition of Intermediate Physics for Medicine and Biology. We visited the Pantheon, where we saw the tomb of Marie Curie (a unit of nuclear decay activity, the curie, was named after her and is discussed on page 489 of Intermediate Physics for Medicine and Biology). Marie Curie lies next to her husband Pierre Curie (of the Curie temperature, page 216). Also in the Pantheon is Jean Perrin, who determined Avogadro’s number (see the footnote on page 85) and Paul Langevin, of the Langevin equation (page 87). Hanging from the top of the dome is a Foucault pendulum, in the exact place where Leon Foucault publicly demonstrated the rotation of the earth in 1851. I like it when physics takes center stage like that.

Another scientific site we visited is a museum honoring Louis Pasteur at the Pasteur Institute. Pasteur chose to be buried in his home (now the museum) rather than in the Pantheon. Readers of Intermediate Physics for Medicine and Biology will find him to be an excellent example of a researcher who bridges the physical and biological sciences. His first job was as a professor of physics, although he would probably be considered more of a chemist that a physicist. His early work was on chiral molecules and how they rotated light. He later became famous for his research on the spontaneous generation of life and a vaccine for rabies. In his book Adding A Dimension, Isaac Asimov lists Pasteur as one of the ten greatest scientists of all time. The museum is enjoyable, although it is not as accessible to English speakers as some of the larger museums such as the Louvre and the delightful Musee d’Orsay. Because I speak no French, I had a difficult time following many of the Pasteur exhibits. Also at the museum was a nice display about microbiologist Jacques Monod, who I will discuss in a future entry to this blog.

A Short History of Chemistry, by Isaac Asimov, superimposed on Intermediate Physics for Medicine and Biology.
A Short History of Chemistry,
by Isaac Asimov.
The only other French scientist on Asimov’s top-ten list was the chemist Antoine Lavoisier. Oddly, the French don’t seem to celebrate Lavoisier’s accomplishments as much as you might expect. (Beware, my conclusion is based on a brief 2-week vacation, and I may have missed something.) Perhaps his death by the guillotine during the French Revolution has something to do with it. We visited the Place de la Concorde, where Lavoisier was beheaded. In A Short History of Chemistry, Asimov writes
In 1794, then, this man [Lavoisier], one of the greatest chemists who ever lived, was needlessly and uselessly killed in the prime of life. “It required only a moment to sever that head, and perhaps a century will not be sufficient to produce another like it,” said Joseph Lagrange, the great mathematician. Lavoisier is universally remembered today as “the father of modern chemistry.”
I normally associate Leonardo da Vinci with Italy, but when touring the Chateau at Amboise in the Loire Valley, we stumbled unexpectedly upon his grave. He spent the last three years of his life in France. We toured an excellent museum dedicated to da Vinci, containing life-size reconstructions of some of his engineering inventions. Although da Vinci had many interests and may be best known for his paintings (yes, I saw the Mona Lisa while at the Louvre), at least some of his work might be called biomedical engineering, such as his work on an underwater breathing apparatus and on human flight.

Seventy-two famous French scientists and mathematicians are listed on the Eifel Tower, including Laplace (of the Laplacian, page 91), Ampere (of Ampere’s law, page 206, and the unit of current, page 145), Navier (of the Navier-Stokes equation, page 27), Legendre (of Legendre polynomials, page 184), Becquerel (of the unit of activity, page 489), Fresnel (of the Fresnel zone for diffraction, page 352), Coulomb (of the unit of charge and Coulomb’s law, both on page 137), Poisson (of Poisson’s ratio, page 27; the Poisson-Boltzmann equation, page 230; and the Poisson probability distribution, page 572), Clapeyron (of the Clausius-Clapeyron relation, page 78), and Fourier (of the Fourier series, page 290). I could not see all these names because the tower was partially covered for painting. Note that Lavoisier was included on the Eiffel Tower, but Poiseuille (of Poiseuille flow, page 17) was not. The view from the top of the tower is spectacular.

I admit, I am not the best of travelers and am glad to be home in Michigan. But I believe there is much in France that readers of Intermediate Physics for Medicine and Biology will find interesting.

Friday, July 2, 2010

Reynolds Number

The Reynolds number is a key concept for anyone interested in biofluid dynamics. Russ Hobbie and I discuss the Reynolds number in Section 1.18 (Turbulant Flow and the Reynolds Number) of the 4th edition of Intermediate Physics for Medicine and Biology.
The importance of turbulence (nonlarminar) flow is determined by a dimensionless number characteristic of the system called the Reynolds number NR. It is defined by

NR = L V ρ/η

where L is a length characteristic of the problem, V a velocity characteristic of the problem, ρ the density, and η the viscosity of the fluid. When NR is greater than a few thousand, turbulence usually occurs…

When NR is large, inertial effects are important. External forces accelerate the fluid. This happens when the mass is large and the viscosity is small. As the viscosity increases (for fixed L, V, and ρ) the Reynolds number decreases. When the Reynolds number is small, viscous effects are important. The fluid is not accelerated, and external forces that cause the flow are balanced by viscous forces… The low-Reynolds-number regime is so different from our everyday experience that the effects often seem counterintuitive.”
Steven Vogel, in his fascinating book Life in Moving Fluids, describes the importance of the Reynolds number more elegantly.
The peculiarly powerful Reynolds number [is] the center piece of biological (and even nonbiological) fluid mechanics. The utility of the Reynolds number extends far beyond mere problems of drag; it’s the nearest thing we have to a completely general guide to what’s likely to happen when solid and fluid move with respect to each other. For a biologist, dealing with systems that span an enormous size range, the Reynolds number is the central scaling parameter that makes order of a diverse set of physical phenomena. It plays a role comparable to that of the surface-to-volume ratio in physiology.
The Reynolds number is named after the British engineer Osborne Reynolds (1842–1912). He developed the Reynolds number as a simple way to understand the transition from laminar to turbulent flow of fluids in a pipe. Perhaps it is fitting to let Reynolds have the last word. Below he describes experiments in which he added a filament of dye to the fluid (as quoted by Vogel in Life in Moving Fluids).
When the velocities were sufficiently low, the streak of colour extended in a beautiful straight line across the tube. If the water in the tank had not quite settled to rest, as sufficiently low velocities, the streak would shift about the tube, but there was no appearance of sinuosity. As the velocity was increased by small stages, at some point in the tube, always at a considerable distance from the trumpet or intake, the colour band would all at once mix up with the surrounding water. Any increase in the velocity caused the point of break-down to approach the trumpet, but with no velocities that were tried did it reach this. On viewing the tube by the light of an electric spark, the mass of colour resolved itself into a mass of more or less distinct curls showing eddies.

Friday, June 25, 2010

Adolf Fick

Russ Hobbie and I discuss Fick’s laws of diffusion in Chapter 4 of the 4th edition of Intermediate Physics for Medicine and Biology. The German scientist Adolf Fick (1829–1901) was a classic example of a researcher who was comfortable in both physics and physiology. He enrolled at the University of Marburg with the goal of studying mathematics and physics, but eventually switched to medicine, and earned an MD in 1852. Of particular interest to me is that he wrote a classic textbook titled Medical Physics (1856), which was one of the first books on this topic. I have not read this book, which almost certainly is written in German (although I am half German through my father’s side, I cannot speak or read the language). Nevertheless, I wonder if Intermediate Physics for Medicine and Biology might be a descendant of this text.

Fick was only 26 when he proposed his two laws of diffusion. The first law (Eq. 4.18a in our book)—similar to Ohm’s law for electrical current or Fourier’s law for heat conduction—states that the diffusive flux is proportional to the concentration gradient. The constant of proportionality is the diffusion constant, which Fick first introduced. Fick’s second law (Eq. 4.24) arises by combining his first law with the equation of continuity (Eq. 4.2) and is what we generally refer to as the diffusion equation. He tested his two laws by measuring the diffusion of salt in water. He even noticed the strong temperature dependence of the diffusion constant.

Fick contributed to physiology and medicine in several ways. He made the first successful contact lens, and he developed a method to measure cardiac output based on oxygen consumption and blood oxygen concentration. You can find more information about his life at http://www.corrosion-doctors.org/Biographies/FickBio.htm.