Teaching Biological Physics

8
University of Pennsylvania University of Pennsylvania ScholarlyCommons ScholarlyCommons Department of Physics Papers Department of Physics 3-2005 Teaching Biological Physics Teaching Biological Physics Raymond E. Goldstein Philip C. Nelson University of Pennsylvania, [email protected] Thomas R. Powers Follow this and additional works at: https://repository.upenn.edu/physics_papers Part of the Physics Commons Recommended Citation Recommended Citation Goldstein, R. E., Nelson, P. C., & Powers, T. R. (2005). Teaching Biological Physics. Physics Today, 46-51. Retrieved from https://repository.upenn.edu/physics_papers/490 This paper is posted at ScholarlyCommons. https://repository.upenn.edu/physics_papers/490 For more information, please contact [email protected].

Transcript of Teaching Biological Physics

University of Pennsylvania University of Pennsylvania

ScholarlyCommons ScholarlyCommons

Department of Physics Papers Department of Physics

3-2005

Teaching Biological Physics Teaching Biological Physics

Raymond E. Goldstein

Philip C. Nelson University of Pennsylvania, [email protected]

Thomas R. Powers

Follow this and additional works at: https://repository.upenn.edu/physics_papers

Part of the Physics Commons

Recommended Citation Recommended Citation Goldstein, R. E., Nelson, P. C., & Powers, T. R. (2005). Teaching Biological Physics. Physics Today, 46-51. Retrieved from https://repository.upenn.edu/physics_papers/490

This paper is posted at ScholarlyCommons. https://repository.upenn.edu/physics_papers/490 For more information, please contact [email protected].

Teaching Biological Physics Teaching Biological Physics

Disciplines Disciplines Physical Sciences and Mathematics | Physics

This journal article is available at ScholarlyCommons: https://repository.upenn.edu/physics_papers/490

46 March 2005 Physics Today © 2005 American Institute of Physics, S-0031-9228-0503-030-9

Over the past few years, people trained in physics andworking in physics departments have taken an un-

precedented interest in biological problems. A host of newexperimental and theoretical techniques has opened upthe quantitative study of systems ranging from single mol-ecules to networks of simple agents performing complexcollective tasks. Many departments have begun aggressiveprograms to hire faculty into the emerging field of biolog-ical physics. Engineering departments, too, are investingin the interface of the life and physical sciences, both inbioengineering and in related areas such as chemical en-gineering, solid mechanics, and materials.

Not surprisingly, the new faculty members, like theircolleagues, are interested in teaching the subjects that ex-cite them. Meanwhile, physical-science students are be-ginning to demand courses relevant to the life sciences.And high-level reports such as the National ResearchCouncil’s Bio2010 have emerged to stress the importanceof quantitative, physics-based thinking for future life sci-entists.1 (See also reference 2 and the Reference Frameessay by John J. Hopfield, PHYSICS TODAY, November2002, page 10.)

With all this momentum, it may come as a surprise tofind that many people, particularly junior faculty, tell ushow difficult they have found it to create and sustain newcourses in biological physics. We examine some of the rea-sons for those barriers and offer a few ideas for suchcourses that we have gleaned from our own and ourfriends’ experiences.

It mattersThe stakes are high. Many physics departments feel thatthey must struggle to maintain their student enrollments.It often seems that scientifically talented students come totheir universities already convinced that their best careeroptions lie in the life sciences. That perception can be frus-trating—after all, many exciting advances in molecularand cell biology rest upon physical techniques and ideas.

Why don’t students see this and studymore physics?

Perhaps it is because when stu-dents read course catalogs, they oftenget no hint of the great ferment goingon in our laboratories. Many physicsdepartments teach a collection of un-dergraduate courses whose outlines

are similar to the menu of 30 years ago, despite the factthat the research interests of physics faculty have changeddramatically. The curriculum must change. In particular,we believe that biological physics must become a main-stream course in all physics departments, offered as regu-larly as, for example, courses on solid-state or high-energyphysics. We also believe that many engineering programswould welcome the introduction of such a course, either in-ternally or as offered by physics departments.

Many departments already offer special-topicscourses, graduate courses, and seminars relating to bio-logical physics. Unfortunately, such courses often lack thestability enjoyed by a modern physics course, for example.They are frequently electives that busy students can’t fitinto their already crowded schedules. Often, the courseshave a limited scope, unlike the overviews of a subject thatphysics departments provide in condensed matter or par-ticle physics. They frequently evaporate when a particularfaculty member goes on sabbatical or is needed to teach amore “critical” course. Moreover, they are frequently cre-ated from scratch by junior faculty members, who are al-ready heavily burdened.

We believe it is important to offer a consistent, stan-dardized course that is a requirement for at least some fla-vors of the physics degree, and perhaps for the bioengi-neering degree. An intermediate-level course can be a goodelective for other engineering majors, as well as for majorsin biochemistry and even chemistry. We describe ideas forsuch a course, which can also serve graduate students. Asin other intermediate-level courses, students in biologicalphysics have much to learn from a hands-on laboratory,and we describe some of our experiences implementinglabs at various levels.

Many of the concepts central to biological physics canbe taught at a lower level, and hence in larger servicecourses. The Bio2010 report urges biology departments tobegin requiring rigorous and relevant physics experience.Premedical students are frequently poorly served by tra-ditional physics courses based on the Medical College Ad-missions Test. We describe some ideas for meeting thechallenge of the service course and outline some experi-ences teaching biological physics in the context of an en-gineering program.

Now is a good time to look beyond individual coursesand consider offering an entire degree program in biolog-ical physics, or at least a specialized concentration withinthe physics major. A recent article discusses ideas alongthese lines.3

Ray Goldstein ([email protected]) is a professor ofphysics and applied mathematics and codirector of the quantitativebiology initiative of the BIO5 Institute at the University of Arizona inTucson. Phil Nelson ([email protected]) is a professorof physics at the University of Pennsylvania in Philadelphia andauthor of the textbook Biological Physics: Energy, Information, Life(W. H. Freeman, 2004). Tom Powers ([email protected])is the James R. Rice Assistant Professor of Solid Mechanics andassistant professor of engineering at Brown University in Provi-dence, Rhode Island.

Biological physics courses can serve a variety of students,among them life-sciences students who need to understandthe role of physical principles in the world of biology.

Raymond E. Goldstein, Philip C. Nelson, and Thomas R. Powers

Teaching Biological Physics

Six obstaclesSurely many of the assertions we havemade are not very controversial. So whydo many faculty members find obstaclesto creating biological physics courses?We hear professors say that their col-leagues offer objections along the fol-lowing lines.! “We don’t have enough people; we’rebarely able to teach a complete curricu-lum of the old stuff.” It’s true. Physics de-partments have shrunk relative to thehalcyon days of yore and have forced fac-ulty members to eliminate many ad-vanced courses they think are importantfor their students. The only answer isthat we physicists must relentlesslyprune our offerings to make room for thenew, exciting subjects. Do we really needa two-semester sequence in X, when onemight do? Do we really need to offer Yevery year? Does the enrollment in Z re-ally justify its existence as a regularcourse, or could it be offered as a read-ing course? All of those questionsamount to: Would we really rather retainX, Y, and Z and miss out on biologicalphysics? They’re not easy questions, butthey do admit rational discussion.! “We can’t justify a new elective coursefor our few majors.” This question over-laps the preceding one. We would suggestthat the small number of physics majorsmight result, in part, from a lack of cur-riculum options that students imagine asrelevant to their own future careers.What’s more, we find that biologicalphysics attracts students from severalmajors, not just a small, captive audienceof physics majors.! “It’s already offered in another de-partment.” Other departments do offercourses with names like molecular bio-physics, physiology, medical imaging, and neuroscience.Closer inspection, however, reveals that the material stu-dents need is thinly distributed through several differentcourses—for example, the several semesters of generalchemistry, physical chemistry, organic chemistry, and bio-chemistry. Students don’t have enough time to take allthose courses!

Even more significant is that many courses in other de-partments turn out to be very different from what we inphysics think of as physics education. Our ambition is togive students simple, general-purpose intellectual tools thattie together many kinds of apparently disparate phenom-ena. Many of those tools are quantitative and involve themodeling of an unfamiliar situation to confirm or falsifysome physical hypothesis. We’d like our students to be ableto face a problem they have never seen, pull the right quan-titative tool out of their bag, and use it to solve the problem.Do you really have a course like that in your medical school?! “That’s not really physics.” A course in biologicalphysics can tell a story whose protagonists include MaxPlanck, Albert Einstein, Marian Smoluchowski, LinusPauling, Max Delbrück, Hendrik Kramers, Richard Feyn-man, and a corresponding list of contemporary names. Arethey really physicists? Such a course can also serve as anentry to the conceptual foundations of nanotechnology andsoft condensed matter. It also can supply a context in

which to teach ideas in continuum mechanics, practicallybanished from many physics curricula (see the ReferenceFrame essay by Jerry Gollub, PHYSICS TODAY, December2003, page 10). Is that really physics?! “Our students can’t handle it.” We hear this commentsurprisingly often. The implication seems to be that hardcourses will drive away even the few physics majors left.But students work hard when they think they’re gettingsomething they need. Physical chemistry is a lot of work,it looks good on a medical-school application, and studentstake it. A course in biological physics can play the samekind of role.

Finally, a worry that comes from within:! “I’m nervous! I don’t know all that stuff, and I don’thave time to teach myself and create a whole new course.And what would I cover, anyway? The field is too huge.”We still feel this! It never goes away. But high-energy par-ticle physics is huge, and somehow physicists teach that.In part, we’re comfortable teaching particle physics be-cause we took it when we were students, but more impor-tant, it has a fairly stable canon that’s clearly described intextbooks. Standard texts are now starting to emerge inbiological physics too,4 together with the usual ancillarymaterials like problem sets and solutions. Such curricularmaterials will make it unnecessary for the overburdenedinstructor to invent the whole course from scratch.

http://www.physicstoday.org March 2005 Physics Today 47

f

f

Makes no net progress:a

Makes net progress:b

Figure 1. Mechanical analogies can illuminate physical ideas sometimeshidden in a mass of molecular details. Naively, it would seem that thermalkicks would drive the ratchet in (a) to the right and do work against a load f.Once the students have worked through the microscopic details of why itdoesn’t work, they are ready to study the modified ratchet in (b), where amechanism releases each pawl only after it emerges on the right side of thewall. The device may seem fanciful, but it contains the essence of an ideacurrently believed to underlie real molecular motors. (Adapted from P. Nel-son, ref. 4; used with permission of the publisher.)

48 March 2005 Physics Today http://www.physicstoday.org

We have found it surprisingly easy to get a coopera-tive atmosphere in an interdisciplinary class. The bio-chemistry majors do catch us making biochemical slips.But if we’re respectful of their own uneasiness with di-mensional analysis, then they end up going the extra mileto teach us things we need. Even better, a little hint is oftenall our students need to team up among themselves in in-terdisciplinary study groups.

An intermediate-level courseOne way to plan a course is to fix the starting and endingpoints, then find a least-action trajectory connecting them.For example, one of us constructed a course that picks upthe story where first-year physics courses typically end andfinishes with meaningful discussions of single-molecule ma-nipulation, molecular motors (see figure 1), and the mecha-nism of nerve impulses, all topics of current research inter-est. What points must students visit along the way?

In a nutshell, the answer was to begin with simple es-timates and dimensional analysis, then review a little cellbiology to get the players on the stage. Next, studentsneeded a little kinetic theory, leading to the key concep-tual foundation of random walks. From that vantage point,one can see many of the phenomena that make the sub-micrometer realm so different from our own—for example,the strange world of low Reynolds number.5 Students canalso get a feeling for the tendency of entropy to increase.

Only when students had a feeling for entropy did thecourse introduce its abstract definition. But the abstractviewpoint has its virtues, as it leads to a compact and gen-eral analysis of entropic forces. Viewing chemical forces asa particular case of entropic forces (see figure 2) was a newexperience for most students and one that led naturally tothe study of single-molecule devices like enzymes and mo-lecular motors. Students found it satisfying to analyze acase in which an entropic force can be rigorously computedand compared to experiment, as in the mechanical exten-sion of long molecules of DNA. The course culminated witha discussion of the collective behavior that emerges whenmany similar molecular agents act in concert to create trav-eling nonlinear waves of excitation—the nerve impulse.

The course also assigned reading selections from amainstream cell-biology textbook. There is no substitutefor learning the language, and the iconography, from anindustry-standard source, although some beautiful andtechnically accurate popularizations are also quite use-ful.6 Some course assignments also involved the use ofcomputers. Students need to get into the habit of turning

to mathematical software to ex-amine the behavior of a mod-el; they also need the habit of turning to a molecular-visualization package andallied online databases to ex-amine a macromolecule ofinterest.

The syllabus just out-lined will not suit everyone, of course. But it does havesome claim as a conceptual framework into which manymore advanced ideas can be fitted. We think that studentsneed the core ideas of several disciplines before they moveon to more abstract topics like neural and metabolic net-work analysis. Whatever topics you choose to cover, wesuggest that every concept be rooted in some quantitativeexperimental data, the same rigorous standard that is ap-plied when framing a course in, say, quantum mechanics.As physicists, we were taught that if we can write andsolve a simple model that draws a curve through some realdata and that is rooted in concepts that explained otherkinds of experiments, then we may have learned some-thing. Of course, plenty of life scientists concur with suchan ethos. We would like to instill it in students, even if itmeans skipping much of the voluminous factual materialof cell biology.

Laboratory coursesPerhaps the simplest way to incorporate experiments inbiological physics and soft condensed matter physics intothe undergraduate curriculum is to introduce them intostandard junior- or senior-level physics lab classes. Con-sider the current menu of classic experiments—the speedof light, the electron charge to mass ratio e/m, Mössbauerspectroscopy, and so forth. That lineup exists not becauseeach experiment teaches a particular skill that is so im-portant later in a student’s professional life, but because,taken as a whole, it exposes students to the interplay be-tween theory and experiment and incorporates a healthydose of error analysis. Viewed this way, a lab in which stu-dents use an inexpensive laser-tweezer setup7 to learnabout optical trapping and Stokes drag is surely as im-portant as one in which they use a Michelson interferom-eter to measure the refractive index of gases. Likewise, ex-amining the Brownian motion of micron-size spheresthrough a microscope and determining Avogadro’s numberas Einstein proposed and Jean Perrin carried out must beas important as measuring e/m.

At the graduate level the possibilities are muchgreater, particularly as departments embark on interdis-ciplinary education and research efforts such as the Inte-grative Graduate Education and Research Traineeship(IGERT) program supported by NSF. The experience of oneof us with such an effort has provided the resources andstudents to experiment with how a graduate course can bestructured.8

A centerpiece of that effort is a dedicated laboratory—home to a course taken by all students in the IGERT pro-

Figure 2. Chemical forces are partly entropic in character. If initially the number of buffalo in the lower-energy state

exceeds that state’s equilibrium value, there will be a net flux of buffalo into the higher-energy state.

Reasoning in this way gives students a physical intuition for the concentration-

dependent part of the chemical poten-tial. (Cartoon by Larry Gonick from P.

Nelson, ref. 4; used with permission.)

http://www.physicstoday.org March 2005 Physics Today 49

gram—that combines faculty from applied mathematics,physics, neuroscience, physiology, and molecular and cellbiology. The students come from the above departmentsand others such as biomedical engineering and even as-tronomy. In view of the diverse clientele and a desire togive all students a meaningful laboratory experience, thenotion of a methods course that would simply survey cur-rent laboratory techniques in biological physics was re-jected. Instead, the course emphasizes the interplay of the-ory and experiment through a case-study format. Thatserves the students well, for each student comes with def-inite strengths and weaknesses. Graduate students in ap-plied mathematics, for example, have no problem with themore theoretical underpinnings of the various lab experi-ments. But they must learn to ask questions with an ex-periment, a skill that is usually underemphasized in theircurriculum. Students in biochemistry are very familiarwith biological protocols but are weak on dimensionalanalysis and the study of the partial differential equationsthat describe diffusion or fluid flow. Thrown together in alab course where cooperative learning is emphasized, thestudents can produce remarkable results.

A key issue is determining what are the right experi-ments. Based in part on the research expertise of the fac-ulty running the course, the decision was made to organ-ize the experiments by length scale, as illustrated in figure3. After several weeks of introductory lecture material onthe essentials of dimensional analysis and estimation,Stokes drag and the Stokes–Einstein relation, electro-physiology, pattern formation, microscopy, micromanipu-lation, and time-series analysis, each pair of studentsspent the remainder of the semester on one experiment.For example, at the molecular scale of nanometers, stu-dents investigated the properties of the motor protein ki-nesin as it carries microspheres along microtubules. Re-lated to that study was an investigation of the Kramersproblem of thermally assisted hopping over a potentialbarrier created by two nearby optical traps. At the cellu-lar scale of microns, students measured action potentialsin neurons of the moth Manduca sexta. Particle-trackingstudies of bacterial chemotaxis and related fluid motionsexplored scales of hundreds of microns. Experiments at thecentimeter scale investigated pattern-forming processesincluding bacterial bioconvection and the Belousov–

Zhabotinsky reaction known for its rotating spiral waves. Each experiment has a well-established underlying

theoretical description. Like the experiments themselves,some of those theories are quite involved. Yet, by spendingmuch of a semester focused on both theory and experi-ment, surrounded by fellow students from a broad rangeof departments in the atmosphere of a research group, theIGERT students mastered them.

The laboratory layout includes common space for thestudents to gather for theoretical lectures and to make oralpresentations. That space also doubles as the home for anundergraduate biological physics course in which the ex-perimental setups serve as sophisticated demonstrations.What better way to lecture about bacterial chemotaxisthan with a suitable microscope setup three meters awayso the students can see it with their own eyes?

A no-prerequisites courseA course for students with no university-level physics ex-perience can be structured in many ways. For example,one could add a half-credit seminar that runs parallel toan existing version of first-year physics. One such coursewas recently taught by Jané Kondev and Robert Meyer atBrandeis University.9

The course is primarily intended for first-year stu-dents who are considering a major in biological physics, al-though enrollment of premedical students and studentsfrom other majors is not uncommon. As a survey, the sem-inar plays the dual role of attracting students to the majorand providing motivation to labor through the dauntingmainstream introductory courses in physics, chemistry,and biology. The course meets once a week over two se-mesters. Thus, the time commitment required of the stu-dents is not too big, which is particularly important forfreshmen, who are very busy fulfilling other university re-quirements. The course has no exams, but the studentshave weekly homework and reading. The homework con-sists mostly of estimation; the readings are taken from el-ementary physics and biology texts.

The physics content of the course follows the first-yearphysics sequence, with mechanics and thermodynamics inthe fall semester and electricity and magnetism and opticsin the spring. The corresponding coverage has the advan-tage of letting the students see how physics concepts

a c d

Figure 3. Experiments at several length scales are the cornerstone of an interdisciplinary laboratory course for gradu-ate students. The photographs here illustrate (a) a microsphere (large black-and-white structure) pulled along a mi-crotubule by the motor protein kinesin, (b) fluorescently labeled ganglia in the moth Manduca sexta, (c) bioconvec-tion patterns in a drop of bacterial suspension, and (d) spiral waves from a Belousov–Zhabotinsky reaction that takesplace in a petri dish. (Images a, b, and d courtesy of Koen Visscher and Robert Reinking, University of Arizona.)

3 mm 125 mm 1 mm 4 mm

b

50 March 2005 Physics Today http://www.physicstoday.org

treated in other courses are applied practically to problemsin biology. For example, a simple model of springs con-nected in parallel leads to an estimate for the stretch mod-ulus of single-stranded DNA. The modulus, in turn, can be read from force–extension graphs obtained in single-molecule DNA stretching experiments.

In addition to applying principles of mechanics andthermodynamics, the course introduces new concepts suchas diffusion. The mathematics of diffusion is developedfrom a one-dimensional random-walk model that alsogives insight into the size of polymers in solution and al-lows for simple estimates of the size of the E. coli genome.For example, students study micrographs that show thebacterium’s DNA spilled out of the cell; those images yielda crude estimate of the radius of gyration. In the springsemester, simple concepts from electricity and magnetismand optics are applied to such topics as conduction by ionchannels and the compound lens of a bee’s eye.

The goal of the course is to introduce the field of bio-logical physics by examining the living world quantitatively.A corresponding course theme is the idea that simple quan-titative estimates can reveal all sorts of interesting infor-mation. An example is given in figure 4, which shows howthe cell division time for E. coli can be used to estimate therate of amino-acid addition on the ribosome.

Biological physics for engineersPhysics undergraduates and engineering undergraduatestypically have different motivations for their courses ofstudy, and those differences must be respected when teach-ing biological physics to engineers. Broadly speaking,physics students study science to understand nature,whereas engineering students study science because it is

the basis for technology. In particular, although engineer-ing students appreciate the importance of fundamentalunderstanding and are capable of the same level of math-ematical sophistication as physics students, they are eagerto know the application of a model before they are readyto give their full attention to its development. Engineer-ing students want to see the connection of their course-work to the “real world” of industry. But, having studiedtopics such as beam theory, plate theory, and transport the-ory in their traditional courses, they are eager to see howthat material applies to biology. Bioengineering studentsare especially excited to see material familiar from theirbiology courses treated from the physical point of view.

What does the engineering student’s perspectivemean for a syllabus? One approach we have found to besuccessful is to organize the syllabus around biological orbiotechnological questions. Engineering students are par-ticularly receptive to questions that relate to optimizationor physical constraints. For example, at the beginning ofthe course, one might discuss the packing of DNA into thehead of a virus. A simple question is, How much work doesit take to stuff the DNA into the head? A natural place tostart is with the random-walk model of a polymer. The stu-dents are then led directly to the ideas of entropy and en-tropic forces. That path has the side benefit of helping tocement the students’ understanding of entropy, which, inthe engineering curriculum, is typically first encounteredfrom the abstract perspective of thermodynamics. Oncethe idea of an entropic force is in place, it is natural to in-troduce single-molecule experiments and experimentalforce–extension curves, and then point out the limitationsof the random-walk model as an explanation for those

0 0 0 00 0 0 00 0 3 60 5 0 0. . . .

a Bacterial cell division

b Protein synthesis

Minutes

Deciseconds

Ribosome

MessengerRNA

Figure 4. The rate of protein synthesis on ribosomes in E. coli can be estimated from the cell division time. As shown in (a),the bacterium divides in about half an hour. During that time, the mass of all the proteins roughly doubles. To estimate the totalprotein mass, note that about 20% of the cell volume is taken up by proteins, the volume of E. coli is approximately 1 mm3,and the density of proteins is close to that of water. Thus, the protein mass is about 0.2 × 1 mm3 × 1 g/cm3 " 1011 daltons; a Dais equal to an atomic mass unit. An examination of the 20 amino acids found in living matter reveals that their mass is typically100 Da. The number of amino acids that need to be strung together to make all the proteins for the two daughter cells is there-fore 1011 Da/100 Da ⊂ 109. One can estimate from microscopy that an E. coli bacterium has on the order of 2 × 104 ribo-somes, so the average rate of amino-acid addition on any one of them is 109/(2 × 104 × 2000 s) " 20 s⊗1, as shown in (b), andwhich roughly corresponds to the experimentally established value. (Adapted from R. P. Phillips, J. Kondev, Physical Biology ofthe Cell, to be published by Garland Press. Used with permission.)

March 2005 Physics Today 51

curves. Students begin to see how accounting for DNA’selastic resistance to bending and twisting explains the ex-periments, and are then in a position to evaluate the rel-ative importance of elastic and entropic effects in the orig-inal question of DNA packing.

The idea of organizing the syllabus around biologicalquestions can be readily continued for a host of topics. Forexample, membrane and vesicle mechanics can be moti-vated by a discussion of encapsulated drug delivery, or thephysics of cell adhesion can be motivated by a discussionof the challenges of tissue engineering. The unity of me-chanics and statistical mechanics allows those disparatetopics to be treated with a limited number of fairly simpleideas.

Go for itCreating a sustained curriculum change is a big job. Butit’s important to meet the challenge: Over the past decade,many physics departments have seen significant erosionin their service teaching and hence their institutional sup-port. We need to explore all possible ways to offer newcourses that other departments (and their students) actu-ally want. Moreover, many undergraduate students arebasing their choice of graduate school, in part, on the avail-ability of biological physics programs. Soon, undergradu-ates will choose their majors and high-school students willselect their colleges on similar grounds.

But it’s not all about gloomy trends! It’s also excitingto open new doors for students, and gratifying when theyrespond as they have to our courses. What’s more, we havefound teaching biological physics—a course we never tookas undergraduates—to be a valuable and fun part of ourown professional growth. Some of that excitement hasrubbed off on our students, then back onto us, and so on.The prospect of such stimulation is a big part of why wewent into academic careers in the first place. So if you findyourself teaching biological physics, remember to enjoy it.And good luck.

We thank Jané Kondev and Rob Phillips for some of the ma-terial summarized in this article.

References1. National Research Council, Bio2010: Transforming Under-

graduate Education for Future Research Biologists, NationalAcademies Press, Washington, DC (2003), available athttp://www.nap.edu/books/0309085357/html.

2. W. Bialek, D. Botstein, Science 303, 788 (2004); B. Alberts,Cell 92, 291 (1998).

3. S. Amador Kane, Am. J. Phys. 70, 581 (2002).4. See, for example, B. Ahlborn, Zoological Physics, Springer,

New York (2004); B. Cotterill, Biophysics, Wiley, New York(2002); K. A. Dill, S. Bromberg, Molecular Driving Forces, Gar-land, New York (2002); P. Nelson, Biological Physics: Energy,Information, Life, W. H. Freeman, New York (2004); B. Nolt-ing, Methods in Modern Biophysics, Springer, New York(2003).

5. E. Purcell, Am. J. Phys. 45, 3 (1977); H. Berg, Random Walksin Biology, 2nd ed., Princeton U. Press, Princeton, NJ (1993).

6. See, for example, M. Hoagland, B. Dodson, The Way LifeWorks, Random House, New York (1995).

7. S. P. Smith et al., Am. J. Phys. 67, 26 (1999); J. Bechhoefer, S.Wilson, Am. J. Phys. 70, 393 (2002).

8. R. E. Goldstein et al., http://www.physics.arizona.edu/%gold/ajp.pdf. The Integrative Graduate Education and Re-search Traineeship program at the University of Arizona is de-scribed at http://bmpi.web.arizona.edu.

9. See http://www.brandeis.edu/programs/biophysics. !

www.pt.ims.ca/4626-27 or Circle #27