<<

Polymer : a Geometric Introduction

T. A. Witten James Franck Institute University of Chicago Chicago Illinois 60637

Typeset May 25, 1998 http://mrsec.uchicago.edu/tten/Geometric

Long chain polymers dissolved in a qualitatively alter the way the liquid moves and transmits forces. The basic origins of this behavior can be understood geometrically by recognizing that a polymer chain resembles a . The spatial distribution of may be described by scaling properties and quantied using the notion of fractal dimension and dilation invariance. The strong thermodynamic and hydrodynamic interactions of polymers may be accounted for in terms of the intersection properties of fractal objects. These intersection properties show why polymers exclude ow as well as one another from their interiors, despite their arbitrarily small interior concentration. Self avoidance decreases the fractal dimension of a polymer. The origin of this decrease and conditions for its occurance are explained. From these geometric properties, scaling laws describing how osmotic pressure, diusion and relaxation depend on molecular weight and concentration are explained.

I. INTRODUCTION mers. This enables us to provide an economical and uni- ed overview of the phenomena treated in greater depth in these books. As much as possible we shall attempt to Why should a physicist be interested in polymers? account for the interaction of polymers with their sur- They do not hold the key to vast sources of energy as roundings by the mathematical laws describing the in- atomic nuclei do. They do not defy the intuition with ul- tersections between two fractal structures. We begin by trasmall dissipation as superconductors and superuids recalling the scaling properties of any exible chain of do. They do not reveal subtle new nonabelian symme- randomly-oriented links, noting that such a random-walk tries as do subatomic particles. Nor do they hold secrets structure has the spatial scaling properties of a fractal about the origin or fate of the universe. Polymers are object as dened by Mandelbrot (Mandelbrot 1982). We just ordinary matter—just insulating organic . shall then survey the important ways in which fractal These molecules are merely larger than usual and are in structures, including polymers, interact with their envi- the form of chains of small subunits. Yet these chain ronments. This leads to a discussion of the thermody- molecules have ways of interacting unmatched in other namics and hydrodynamics of a polymer . With forms of matter. The study of polymers over the last the fractal properties in mind we discuss the interac- few decades has forced us to broaden our notion of how tion of a random-walk polymer with itself, nding that matter can behave—how it can organize itself in space, these interactions change the spatial arrangement of the how it can ow, and how it can transmit forces. These considerably. Having described the behavior of new behaviors arise from a few qualitative features of the individual polymer molecules, we can discuss the behav- polymer molecule’s structure. The potential for shap- ior of solutions, notably those where the polymer chains ing these phenomena is just beginning to be realized as interpenetrate strongly. The spatial, energetic and dy- the power of synthetic chemists to control the molecular namic behavior of these solutions can be understood in structure increases. its major outlines from the fractal properties of isolated The purpose of this introductory essay is to convey an polymers treated earlier. We collect in Table I the main understanding of how polymer dier from other quantities used for this discussion. liquids and from other forms of condensed matter. Ex- cellent treatments exist already (Jannink et al. 1992), notably Scaling Concepts in by P. G. deGennes (DeGennes 1979). Here we adopt a geometric approach, exploiting the “fractal” structure of the poly-

This article is adapted from the author’s chapter in La Juste Argile, M. Daoud and C. Williams, eds. (les editions de physique, Les Ulis France 1995), to be published in English as M. Daoud and C.E. Williams (Eds.): Physics (Springer-Verlag, Berlin Heidelberg 1999)

1 II. RANDOM-WALK POLYMER intricacies we note two simple features. First, the - carbon bonds along the chain backbone can rotate. Sec- ond, the successive bond angles of the are bent, so that the chain is exible. This means that the vector ~a linking the rst and last atoms of the seg- ment shown is nearly unaected by the direction of the preceding segment along the chain. This means that the overall shape of the polymer can be represented as a se- quence of n vectors ~ai that vary randomly and are statis- tically independent of each other1. A typical molecule might have a hundred segments of this size; one can make polystyrene molecules containing 10,000 or more such segments. Thus the overall structure of the 1 molecule is apparently that of a long, random walk in 2 space. 3 a To gauge the overall size of a large polymer, we recall 4 familiar properties of random walks. We expect longer polymers to extend over larger distances in space. For random walks we can nd this relationshipp in a quanti- h 2i tative way. The root-mean squaredp distance√ rn gives h 2i the typical size of a walk: rn = na. The typical H H size is proportional to the square root of the chainlength C C C C H H n. Since a random-walk polymer of n segments is like a

C C random walk of n steps, the same scaling law applies to H H H H C C C C these polymers2.

C C C C H H H H The random-walk structure of exible polymers has C C H H an important consequence that accounts for many of the distinctive properties of polymers: they are tenuous.To FIG. 1. Top: Detail of a polystyrene molecule ((CH– explain this notion, we consider a sphere of radius R C6H5)–CH2)n) as it might appear in a good solvent . much larger than our elementary segments a but much Each sphere represents a carbon and one or two smaller than the entire polymer coil. The sphere contains small hydrogen atoms attached to it. The distance be- and solvent; since the liquid has a well-dened tween connected atoms is about 1.4 Angstroms. The density, the amount of matter inside is proportional to 3 chemical bonds are superimposed on one repeating unit R —i.e., the volume of the sphere. But the fraction i or , and on a section of the chain backbone. of this volume occupied by monomers is very small. The The backbone bonds may rotate freely. A few successive piece of chain in the sphere varies in length as the chain backbone repeat units are labeled 0, 1, 2, ... . The vec- uctuates. But the average number n(R)p of segments h 2i tor ~a for a four-monomer segment is shown. This struc- is roughly that number whose typical size rn is the ture was generated by a Monte Carlo computer simula- sphere radius R. Thus, n(R) ' (R/a)2. Knowing how n tion(Mondello et al. 1991), which simulates the random varies with R, we can readily deduce how the volume frac- rotations of the bonds as they might occur in solution. tion i depends on R. If each of the n polymer segments The simulated molecule has about 1/20 the mass of a displaces volume v in the solution then ' vn/R3. Since polymer in a typical styrofoam cup. Bottom: Chemical n grows as R2, the volume fraction scales like 1/R:as diagram of two repeating units of polystyrene. the sphere is made larger, the density of polymer within it decreases indenitely. If the sphere is several hundred Figure 1 shows a close-up of polystyrene, the best Angstroms in size, the proportion of polymer inside may studied polymer molecule and one of the most common. be only a fraction of a percent. We describe objects with The picture suggests the many intricacies controlling the such arbitrarily low density as “tenuous”. Large poly- structure and interactions of these molecules with them- mers in a solution often amount to no more than a dilute selves and with their solvent environment. From these impurity, whose volume fraction is as small as in this

1Though the statistical independence of the segments is not quite complete, it becomes rapidly more so as one considers longer segments. For the time being we shall suppose that the segments uctuate completely independently. 2In what follows, we shall see that real polymers are often not exactly random-walk polymers, so that this proportionality is somewhat altered. For the time being, we’ll conne our discussion to random-walk polymers.

2 sphere. But as we shall see below, this vanishingly small main interest is the case where these objects are poly- amount of polymer can exert a major inuence on certain mers. But we shall try to make our reasoning applicable liquid properties. for any connected fractal structures, with general dimen- These tenuous polymers t the denition of a fractal sion D. In this way our conclusions will be applicable object. A fractal object is an assembly of particles dis- to several types of fractal structure that impart special tributed in space in such a way that the average number properties to liquids: random-walk and rigid linear poly- of particles n(R) within a sphere of radius R surrounding mers (Elias 1984), branched polymers (Daoud 1995), and an arbitrary particle varies as R to some xed power D colloidal aggregates (Witten et al. 1986). 3 called the “fractal dimension” (Mandelbrot 1982) Any We shall suppose that the constituent particles have fractal object is tenuous: the average volume fraction size a, and that all the fractals have the same size R. 3 D3 (R) in the sphere is vn(R)/R , which varies as R . For deniteness we can imagine that these had been Whenever D is smaller than 3, as it is in a fractal, the made by cutting out spherical volumes from a fractal power is negative and becomes arbitrarily small. Our much larger than R. The region within each sphere reasoning from the last paragraph implies that random- is often called the pervaded volume of the object. For walk polymers are fractal objects with fractal dimension the time being we shall treat each fractal as a frozen D = 2. One may deduce several important features of object with no internal motion. We may apply the how polymers behave in a liquid using only their fractal fractal law n(r)=(constant)rD for r = a to deduce property. n(R) ' (R/a)D. Each type of fractal has its own value of the fractal dimension D, so that the n(R) for dier- ent types of object may be quite dierent. We imagine III. FRACTALS IN SOLUTION that the concentration of objects is roughly as suggested in Figure 2: the distance between objects is somewhat larger than their size R. Since each of these objects is tenuous, the overall volume fraction of particles is very small. In what follows we shall focus on the limiting be- havior as R becomes indenitely large compared to the particle size a. Our problem is to gauge how much these few fractals inuence the macroscopic properties of the liquid. One basic inuence is their thermodynamic eects, seen. e.g. , through the osmotic pressure (Reif 1965). If the objects are conned in part of the solution by some membrane that allows the solvent but not the objects to pass, the objects exert an outward average force on the membrane. R The force per unit area is the osmotic pressure. Osmotic pressure measurements of polymer solutions are a routine procedure. Even if the objects exerted no in- uence on each other, there would be osmotic pressure, FIG. 2. A liquid containing fractals. Represented are just as there is pressure in an ideal of noninteract- two linear polymers, a branched polymer and a colloidal ing atoms. But just as the pressure in a real gas diers . Each has been trimmed to t into a spherical from the ideal gas presure, so the osmotic pressure in our pervaded volume of radius R, shown as shaded regions. solution of objects is not quite the ideal pressure. The For comparison a cloud of dispersed monomers from one dierence reects the inuence of the objects upon each of the polymers is also shown. other. These objects inuence each other because no two par- We now imagine many fractal objects oating in so- ticles can be in the same place at the same time. That lution as sketched in Figure 2 . In this section we shall means that the objects cannot be positioned arbitrarily nd that depending on their D values, two fractals in in the solution. If we did position them arbitrarily, we this solution show one of two qualitative dierent types would sometimes create illegal congurations in which of interaction. They may be “mutually transparent” and two or more constituent particles intersected. One way readily interpenetrate each other, or they may be “mu- to construct a valid state of the solution is to start by po- tually opaque” and strongly resist interpenetration. Our sitioning the objects completely at random within it and

3The sphere must be much larger than the size of a particle and much smaller than the overall size of the object. One cannot hope to meet both of these conditions arbitrarily well unless the object contains arbitrarily many particles.

3 then simply removing all the objects that make illegal small internal volume fraction i as the original fractal. intersections. When the illegal congurations have been If the separation r between the clouds is near zero, the discarded, each object has in eect been pushed out of the two clouds overlap completely. The probability that a excluded volume around the other objects. This reduc- given particle of the rst cloud intersects is essentially 5 tion of accessible volume increases the osmotic pressure the volume fraction of the second cloud i. This prob- exerted by the objects. The fraction of objects removed ability is much less than unity in a fractal object, as we for this reason is just the fractional increase in osmotic have seen. But the average number of particles in the pressure due to interactions. Thus to gauge the impor- rst cloud that intersect is the sum of this small proba- tance of interactions, we may look at an arbitrary object bility for each of the n particles in the rst cloud. The in the solution, and ask what is the probability that it average number I is thus n times the single-intersection 2 3 will be discarded because it intersects another. probability, i.e., I ' ni ' n (a/R) . Using the fractal law: n ' (R/a)D, the average number of intersections is A convenient way to treat the discarding process is via roughly I ' (R/a)2D3.IfD>3/2, I becomes arbitrar- the “pair distribution function” g(r). To dene g(r)we ily large for suciently large R. Random-walk polymers, arbitrarily designate a “home particle” near the center with D = 2, are in this category. of each fractal object. Then g(r) means the probability Naturally the probability g(r)ofno intersection is very that two objects whose home particles are r apart are in small for such clouds6. The same reasoning holds if the a legal, nonintersecting conguration, which survives the two clouds are displaced by some distance r<2R: the discarding process. If two solid spheres of radius R were average number of intersections is still arbitrarily large placed at random in a very large volume , all separa- compared to 1. Thus the survival probability g(r) re- tions r<2R would be illegal and all larger separations mains much smaller than 1 for all r<2R. These would be legal. That is, g(r)=0forr<2R and g(r)=1 clouds must create osmotic pressure in the solution just for r>2R. The probability of an illegal intersection is as though they were spheres of radius R. This simply the illegal volume 4 (2R)3 divided by the total 3 solid-sphere behavior emerges no matter how tenuous the volume . This all-or-nothing picture is not true for clouds are, provided R and thence n are large enough fractals. When they are placed close to each other they and provided D>3/2. Even if we halve the size of the may nevertheless avoid each other: g(r) is greater than constituent particles of a large cloud, there is no cor- zero even for r<2R. The overall discarding probability responding decrease in the excluded volume V . This is simply that for the solid spheres times the average4of t volume changes arbitrarily little. There are so many ille- (1 g(r)), i.e., [h1 g(r)i ( 4 (2R)3)]/. The nu- r<2R 3 gal intersections that even when their number is cut by merator in [...] is called the mutual excluded volume V t a large factor, the chance of escaping from all of them of the two objects in question. If the solution contains remains negligible. a large number of objects N, the fractional increase in osmotic pressure is simply half the fractional volume ex- This solid-sphere behavior can be greatly weakened by 1 cluded to each one: viz. , 2 NVt/. Two neighboring frac- re-arranging the n particles of our original fractal ob- tals have nearly empty solvent in the overlapping regions. jects. To show this, we now arrange our particles into Yet the probability of intersection (1 g(r)) can still be solid square rods. Each rod must thus have a width w important. Whether this probability is high or low de- such that the volume of the rod is equal to that of all the pends on how the particles are arranged. To illustrate particles: 2Rw2 ' a3n. Naturally, this width is much this point, let us consider two extreme arrangements with smaller than the length 2R. We may take the home par- contrasting intersection behavior. In the rst arrange- ticle to be at the center of each rod. If the two rods are ment, the n particles of each object are distributed com- spaced at separation r

4Explicitly, we mean the average over all separations R: R d3r(1 g(r)) h1 g(r)i r<2RR . r<2R d3r 1 r<2R

5 More precisely, i is the probability that a given point is occupied by a particle. The probability that some particle intersects a given particle is 8i. 6 Specically, the probability of no intersection for a given particle of the rst cloud is (1 i). The probability that all n n particles do not intersect is the product of the (independent) probabilities for each of the n particles: i.e., (1 i) . For large n and small i this product becomes exp(ni) ' exp(I). When the average number of intersections I is large, the survival probability g(r) is very small.

4 two rods with r = R/2. rated by a distance r of order R. Since they have been placed completely at random, the average number of in- tersections I(r) is no dierent than if they were uncorre- lated clouds. But since these objects are fractals, these intersections occur in large groups, and there is some probability—namely g(r)— that there is no intersection at all. The average number of intersections can be ex- pressed as the probability of an intersection (1 g(r)) times the number intersections, given that there is at least one. We call this M(r); thus I(r)=(1g(r))M(r). This scheme allows us to understand the g(r)ofthe r clouds and the rods discussed above. For the clouds there is no dierence between I(r) and M(r). The av- erage number of intersections M if there is known to be one is virtually the same as the overall average I.Thus I(r)=(1 g(r))M(r) becomes I(r)/M (r)=1 g(r). Since the left side is nearly one, g(r) must be much less than one, as found above. For the rods, I(r)isun- changed. But the number of extra intersections M is FIG. 3. Two rods of length 2R whose centers are sepa- dierent. For example if the two rods cross at right an- rated by R/2. The sphere indicates the possible orienta- gles, the number of intersections is (w/a)3. Since in- tions for the rod on the right. The shaded area suggests tersecting rods typically intersect at some nite angle, forbidden orientations which would intersect the other the average M is also of order (w/a)3. Now, we recall rod. that w2 ' na3/R, and I ' n2/R3 ' w4/(a3R). Thus (1 g)=I/M ' w/R, in agreement with our previ- From the picture it is clear that the rods may eas- ous conclusion. Intersections must be rare, since any ily avoid each other. For any typical orientation of the intersection that does occur implies a number of extra rst rod, the second rod may have almost any orientation intersections that is much larger than the average. without hitting it. Since all orientations are equally likely a priori, the probability that the second rod would have a For two general fractals M and I are of the same order of magnitude, but not identical. Extending the reasoning forbidden orientation is proportional to the shaded frac- ' 3 ' D1+D2 3 tion of the sphere in the Figure. This fraction is small; above, I n1n2/R (R/a) . If this exponent thus the probability g(r) that two randomly oriented rods is negative, then the average number of intersections de- would not intersect is near unity. It is much larger than creases indenitely as R increases. Then the probability it was for the two clouds. Accordingly, the excluded vol- of intersection 1 g(r) is also indenitely small. The ume is much smaller than for the clouds. To work out fractals interfere with each other hardly at all for most the allowed orientations and thus the excluded volume r 3, the average number of intersections 3 I is large. The probability of intersection, though smaller average h1 g(r)i is roughly w/R.ThusVt ' (w/R)R , 3 than for two random clouds, is still substantial. For ex- compared with R for the clouds. 1 The above examples show that it is a subtle task to ample if r = 2 R, the legal probability g(r)isaxed →∞ gauge the strength of osmotic interactions of our fractal number that does not go to zero or to unity as R . polymers. Even simple rearrangements of the particles Thus non-transparent fractals can interpenetrate much → can have a big eect on the interaction strength. Hap- less than transparent fractals, where g(r) 1. But they pily, though, the intersections of fractals (without any re- can interpenetrate much more than if their particles were → arrangements) can be readily analyzed. Thus we consider rearranged into random clouds; there g(r) 0. the g(r) of two fractals with dimensions D1 and D2, both The mutual excluded volume of two objects of size R made from particles of radius a. One way to infer this is the solid-sphere excluded volume times the average g(r) is by analyzing its eect on the number of intersec- h1 g(r)i, as noted above. Since g(r) neither approaches tions I(r). In fractals, unlike clouds, these intersections 0 or 1 for most of the sphere volume, its average must are not statistically independent. If one intersection oc- also be some nite, nonzero fraction. Thus these fractals curs at some point, many others must occur near it. We have an excluded volume that is a nite, nonzero fraction denote the number of extra intersections by M. That is, of the solid-sphere excluded volume, despite their tenu- M is the number of intersections for those objects which ous structure. We term such fractals mutually opaque. have at least one intersection. We conclude that two fractals may interact in one of two This M can be used to nd information about our qualitatively dierent ways. If D1 + D2 < 3 they are desired probability g(r). We consider two fractals sepa- transparent, and have a mutual excluded volume much

5 smaller than their pervaded volume. If D1 +D2 > 3, they IV. DIFFUSION AND FLOW NEAR A FRACTAL are opaque and have a mutual excluded volume that is a xed fraction of their pervaded volume7. One fundamental way a polymer interacts with its uid surroundings is by modifying the ow of this uid. This Whenever two fractals are mutually opaque, their spe- modication is important. It leads to the thickening of cic fractal dimensions D1 and D2 are not important in uids like , or salad dressing. The determining their excluded volume Vt. The size of their same principles control a simpler phenomenon: the ab- constituent particles is also unimportant. Only their sorption of a diusing substance by the polymers or other 3 overall size R is important. The ratio Vt/R must ap- fractal absorbers. The diusing substance might be a proach a nite limit as R goes to innity. Another way small-molecule solute that sticks to the fractal on con- to express this scaling is to consider the solid spheres tact. It might be some electronically excited molecule that would produce the same osmotic pressure as the that diuses through the liquid and de-excites when it fractals in question. Thus every population of fractals of encounters the fractal (DeGennes 1982). All of these dif- size R also has a “thermodynamic radius” Rt. This Rt fusers can be described as a substance with concentration is dened to be radius of solid spheres that have exactly u(r) initially distributed uniformly throughout the solu- the same excluded volume as the fractals—by denitiion, tion at some concentration u . Whenever a particle of 4 3 ' 3 0 Vt = 3 (2Rt) . Since for opaque fractals Vt R , this this substance encounters the fractal, it disappears. We means that their “thermodynamic size” Rt is compara- wish to know how the fractal structure inuences the rate ble to their geometric size R. The ratio of Rt to R is an of disappearance. important characteristic of any given type of mutually The motion of a diusing particle is a random walk. opaque fractals, including polymers; we shall return to it Thus the track that such a particle follows has the same below. fractal properties as a random-walk polymer: both have D = 2. The encounters between a diusing particle and With these properties of fractal intersection in mind, a fractal may thus be viewed as the intersection of two we may readily determine the order of magnitude of the fractals. In this view, the tracks of the diusing particles interactions in a , as measured, e.g. ,by are a tangle of intersecting random walks whose lengths osmotic pressure. We have noted that the excluded vol- increase steadily with time. These walks would ll space ume Vt directly inuences this pressure. Specically, we uniformly were it not for the fractal absorber. But with may express the pressure described above in terms of the this adsorber, any particle that touches the fractal must number of fractals per unit volume, cp = N/. As noted disappear. Thus the random walk representing its fu- above, the osmotic pressure for small cp is given by ture motion must be removed. The remaining random ' 1 walks are no longer uniform; instead, their density near cpkT[1 + 2 Vtcp]. The front factor, proportional to the absolute temperature T , is simply the osmotic pres- the fractal is depleted. By analyzing this density we may sure of an ideal solution of noninteracting objects. Since understand the rate of adsorption. If there is little modi- the coecient Vt is roughly the volume within a polymer cation of the diusing density, there is little adsorption. coil, R3, the interaction term becomes comparable with The most ecient adsorber that can t within a distance the ideal term when the volume per coil, 1/cp, is compa- R of a given center is a perfectly absorbing sphere of 4 3 radius R. rable to the pervaded volume within a coil, 3 R . This behavior is well documented in practice. The concen- As before, it is convenient to designate a home parti- tration at which the interaction term becomes equal to cle on the fractal adsorber. We now consider an arbitrary the ideal pressure is often called the “overlap concentra- point at a distance r from the home point. Despite the tion” cp. Qualitatively, our dilute solution of polymers absorption we expect some of the walker’s tracks to re- generates much more osmotic pressure than if the chains main at r, as shown in Figure 4. The density of such were collapsed into a compact mass of monomers. Our walkers relative to the initial density is the probability random-walk polymers produce osmotic pressure qualita- that the walker at r has not been removed by adsorp- tively as though the large pervaded volume within each tion. This probability is the probability that the were completely lled with monomers. Gram for walk representing its past has not intersected the frac- gram, the fractal structure of a polymer makes it inter- tal. This probability is precisely the g(r) discussed in act qualitatively more strongly with its neighbors than it the last section. The fractals in this case are the ad- would in a compact, nonfractal form. Analogous strong sorbing fractal, with dimension D, and the random walk, interaction occurs with the uid solvent itself, as we now with dimension 2. As we saw above, this g(r) depends explore. on the fractal dimensions. If D + 2 is less than 3, the

7 “Borderline” fractals, with D1 +D2 = 3, are neither transparent nor opaque in general and must be treated on a case-by-case basis.

6 two are mutually transparent: g(r) ' 1. virtually all xed fraction of R. the walkers in the pervaded volume of the fractal never The interaction of a fractal with a gentle ow in the touch the fractal. Their density u is thus virtually unaf- surrounding liquid is similar to its interaction with a dif- ' fected: u(r) u0. But if D + 2 is greater than 3, the two fusing density. If the ow is suciently gentle, it has a are mutually opaque, and g(r) is substantially smaller negligible eect on the shape of even a exible polymer. than 1 for most points r within the absorbing fractal. A fractal can generate such ows to produce hydrody- All connected fractals have D 1 and thus show opaque namic drag when an external force like gravity is exerted behavior. For all such fractals, including random-walk on the fractal. More importantly, the fractal perturbs polymers, the diusing density is substantially depleted an imposed ow, thus increasing the dissipation and in- throughout the volume of the absorber. creasing the eective of the uid. The cases considered above lead us suspect that fractals are espe- cially eective at modifying ow. We shall see below that this is true. We imagine a fractal held stationary as the surround- ing uid gently ows upward at a speed v0. The fractal inuences the uid, because at any point on the (xed) fractal the adjacent ow velocity must be zero. At some given distance r from the fractal’s home particle there is some average velocity v(r). In general, the velocity at r need not be upward as the ow at large distances is. But for our qualitative purposes it is sucient to consider only the vertical part of the velocity. The velocity eld around the fractal is controlled by the laws of hydrody- namics; these express the balance of forces and accelera- tions on each uid element. But in the present situation, the meaning of the hydrodynamic laws can be summa- rized in a phrase: momentum moves through the uid FIG. 4. An absorbing sphere in a uniform sea of random by diusion. The uid innitely far from a fractal has walkers. The density of walkers in the central region a given xed momentum per unit volume, proportional was initially uniform. Then all walkers that intersected to its speed v0. Each fractal particle absorbs all the mo- the absorber were removed. This results in a depletion of mentum density adjacent to it, so that the uid next to density in the vicinity of the absorber. A fractal absorber it is motionless. In steady ow, each particle absorbs mo- with dimension D>1 produces a similar depletion re- mentum at some constant rate, and thus experiences a gion. steady force. Momentum absorbed from the fractal parti- cles is replaced by diusion of momentum from the uid This depletion implies a high rate of adsorption. To outside. Thus the velocity eld plays exactly the role 1 played by the diusing eld u treated above. We may see this, we focus on that distance r such that g(r)= 2 . The fractal must absorb at least as fast as a sphere of view the vertical momentum as being carried by a tangle radius r that absorbs half the particles that hit it8. But of random walkers, as with any other diusing substance this sphere adsorbs at a rate comparable to that of a per- (Roux 1987). fectly absorbing sphere of radius r. Moreover, this r is Any fractal that is transparent to a diusing substance some xed fraction of the fractal’s radius R. These com- is also transparent to ow. Though the uid is stationary parisons lead to the conclusion that our fractal absorbs at near each particle of the fractal, still v(r) ' v0 through- a rate comparable to that of a perfectly absorbing sphere out most of the fractal volume. The uid passes right of the same size as the fractal. As with the mutual inter- through the fractal. On the other hand, any fractal that actions discussed previously, this strong absorption must is opaque to a diusing substance is also opaque to a ow. emerge for a large enough fractal regardless of how small The speed v(r) throughout the interior is substantially (or how weakly absorbing) its constituent particles are. reduced relative to the distant speed v0. Since the uid The indenitely large number of intersections between cannot ow through the fractal, it mostly ows around it. mutually opaque objects compensates for any such weak- The force on our fractal due to the ow can be expressed ness. A fractal of sucient size R absorbs at the same in terms of an equivalent hard sphere—which would feel rate as if it were a perfectly absorbing sphere of some the same force. The radius of this equivalent sphere is

8The actual rate is proportional to the sphere’s radius, the distant concentration and the diusion constant of the diusing substance.

7 called the “hydrodynamic radius” Rh of the object. Since the velocity far outside the fractal is comparable to that the fractal is opaque to ow, this Rh is comparable to the of a solid sphere whose radius is a xed fraction of R. geometric radius R, as seen above for the thermodynamic radius Rt. This means that the fractal absorbs momen- tum at a rate comparable to that of a solid sphere with the same size as the fractal9. The force on the fractal is thus similar to the force on the sphere. Any object moved through a uid exerts a force depending on its size. By reducing the radius of our solid sphere by a - a nite factor, we can achieve the same force that the fractal exerts. The radius of this equivalent sphere is dened as the hydrodynamic radius Rh. Evidently, if the object is a large fractal, the hydrodynamic radius is a xed fraction of its geometric radius R. Hydrodynamic drag on an object also controls the Brownian motion of that object in the solution. Brow- nian motion is caused by random thermal forces in the surrounding solvent. These forces, like imposed forces, produce a proportionate velocity v0, given by the hydro- b dynamic drag ratio (F/v0) characteristic of the object and solvent. Knowing the velocity resulting from the ran- dom thermal forces allows one to calculate the diusion 2 constant . This gives the mean-squared distance x 2 covered in a given time t: x = t. Einstein ( 1905) FIG. 5. Two types of ow: a) elongational ow; b) shear discovered that the diusion constant of any object at 10 ow. The elongational ow shown amounts to a vertical absolute temperature T is determined by its drag ratio : stretching of the upper left picture, as indicated by the = kT(v0/F ). Given the drag ratio of our fractals, this arrows. Shear distortion, indicated by opposed arrows means = kT/(6Rh). Thus opaque fractals diuse in at center, can also be eected by a diagonal stretching a solution at about the same rate as solid spheres of the (lower-left picture) followed by a rotation, to obtain the same overall size. lower-right picture. This opacity holds for other ow conditions, notably shear ow. In a shear ow, like that in a pipe or a of One consequence of the disturbed ow is an increased swirling water, the velocity ~v(r) near the wall runs paral- dissipation of the uid’s kinetic energy as . The lel to it and is proportional to the distance from it. In a power dissipated per unit volume in a small region is shear ow a small cubical region of the uid distorts into equal to the viscosity times the square11 of the shear a rhomboidal box as shown in Figure 5. The fractional rate, ˙ 2. A solid sphere in a uid has none of this dissi- movement of the top of the box relative to the bottom is pation within the sphere. But there is extra dissipation called the shear ; this shear increases at a constant rate outside it: the uid must ow faster in order to pass in time; the shear rate is called ˙ . From the point of view around the sphere. The overall eect is to increase the of a fractal moving with the uid, there is right-moving dissipation moderately in a volume comparable to the uid above it and left-moving uid below it. The fractal sphere volume (Happel et al. 1973). The extra power P particles disturb this ow much as they did the uniform is 2.5 times that which would have occurred in the sphere uid treated above. And again the disturbance is sub- volume V if the sphere had been absent: P =2.5˙ 2V . stantial. The ow within an opaque fractal is reduced by An opaque fractal produces dissipation much like a a nite factor regardless of how large (and tenuous) the solid sphere of the same size. It cannot dissipate more, fractal becomes. And as with uniform ow, the change in since no solid object conned within a spherical volume

9 The actual force F is proportional to the sphere radius Rh, the distant speed v0 and the viscosity of the uid. Specically F =6v0Rh. 10The Boltzmann constant k is a conversion factor from conventional temperature units to energy units. At room temperature, kT is roughly 1/40 of an electron volt—a hundred times smaller than the energy needed to break a polymer chain. 11The work done per unit time on a cubical volume element of side L like that of Figure 5b is the (lateral) force on the top (or bottom) of the element, times the speed of the top face relative to the bottom face. In a viscous uid this force is the product of the viscosity , the shear rate ˙ and the area L2. The speed is the shear rate ˙ times the height L. Combining, the power P dissipated in the cube is proportional to its volume L3: ˙ 2L3.

8 can produce more dissipation than a sphere lling the In this section we have seen how fractals can modify volume. It cannot dissipate much less since the exterior ow and diusion in a liquid. In our discussion we have ow and dissipation are like that of a solid sphere. By idealized the fractals. We have treated them as com- shrinking our solid sphere by some moderate factor, we pletely rigid objects of a xed size R. Polymers, though may nd a radius Rv such that the sphere and the fractal they are fractals, are not rigid and do not have the same have the same dissipation. This Rv is called the “visco- size. Each polymer, as it moves through a liquid by Brow- metric” radius. Regardless of how large and tenuous an nian motion, changes its size and shape continually. It is opaque fractal becomes, Rv/R remains a xed number, only its average size and shape that stay xed. Still the as with Rt/R and Rh/R above. We shall return to dis- laws of ow and diusion discussed above are applicable. cuss the value of this fraction and other related ones. One must simply replace the size R by an appropriate Figure 5a depicts an elongational ow, in which a small average size, such as the average distance between two cube is elongated into a rectangular box. The eect of a arbitrary monomers. Another eect of a polymer’s exi- fractal on a weak elongational ow is the same as in shear bility is that the polymer can be readily distorted by ow. ow. Indeed, a shear ow is equivalent to an elongational In order to have the properties described above the uids ow, as the gure shows. Thus each fractal produces dis- must ow gently enough that this type of distortion can sipation like a solid sphere with the same radius Rv in be neglected. elongational or in shear ow. We have seen that polymers, like all opaque fractals, Since fractals increase dissipation, they increase the ef- have a number of characteristic eects on a uid. They fective viscosity of the uid: more external work—more produce osmotic pressure and viscosity in the uid, and force or pressure—is required to maintain a given ow they individually experience hydrodynamic drag. Each when the fractals are added to the liquid. The macro- of these eects can be quantied by a length. For the scopic viscosity is that required to account for the power osmotic pressure this was the thermodynamic radius Rt, dissipated in a volume . This power P is ˙ 2. On the radius of a sphere with the same excluded volume the other hand, this power can be expressed in terms of as the polymer. For the viscosity it was the viscometric radius Rv. For the drag coecient it was the hydrody- the power P0 without the fractals and the extra power dissipated by the N fractals added. If the viscosity of namic radius Rh. All these radii have roughly the same 2 size, and like the geometric size R they grow with the the pure solvent is s, then P0 = s˙ . Each fractal 1/D 2 4 3 molecular weight M as M , where D is the fractal di- causes a dissipation s˙ (2.5Vv), where Vv = 3 Rv is the volume of the equivalent sphere. Adding the dissi- mension. In the next section, we shall encounter eects pation from the N fractals in the solution volume we that alter the fractal dimension of a polymer from the 2 random-walk dimension of 2. But since our conclusions have P = s˙ [1+2.5Vv(N/)]. Here we imagine a solution dilute enough that the ows around the fractals apply to arbitrary opaque fractals, they apply also to don’t interfere with one another. We recognize the N/ these altered polymers. as the concentration of fractals cp. The macroscopic vis- cosity is increased by the same factor as the power P : V. HOW SELF-INTERACTION CHANGES THE = s[1+2.5Vvcp]. FRACTAL DIMENSION This increase in viscosity is quite similar to the increase in osmotic pressure discussed in the previous section: The conceptual random-walk polymers we’ve consid- =kTc [1 + 1 V c ]. Both osmotic pressure and vis- p 2 t p ered up to now are dierent from real polymers in an cosity are increased in proportion to the concentration important way. A random-walk polymer typically inter- of fractals. For both quantities the amount of increase sects itself many times, while real polymers cannot inter- per fractal is expressed as a volume, Vt or Vv. And for sect themselves. A real polymer is “self-avoiding”: two of opaque fractals both of these volumes are of the order of its atoms cannot be in the same place at the same time! the pervaded volume 4 R3. 3 Thus instead of treating polymers as random walks, it A small amount of fractal increases the vis- would seem more realistic to treat them as self-avoiding cosity noticeably just as it increases osmotic pressure no- random walks. The subtleties of self avoidance can lead ticeably. The same concentration cp that doubles the os- to large changes in a polymer’s size and hence viscosity motic pressure is roughly the concentration that doubles with a few-degree shift in the temperature, as explained the viscosity. As we noted above, this is the “overlap con- below. The dierences between random walks and self- 1/3 centration” at which the distance between fractals cp avoiding random walks also form an important subject in is comparable to their size R, and the overall volume mathematical polymer physics. Our goal here is to give fraction in the solution is comparable to the (small) some account and some feeling for these dierences. internal volume fraction i within a fractal. Fractals, If we consider all the possible bond angles of a polymer, including polymers, are potent viscosiers. Many thick we will generate a large set of congurations. Many of uids in everyday such as motor oil, or bottled sauces these congurations have monomers that intersect each owe their thick consistency to a small proportion of poly- other, and are hence unrealistic. One way to arrive at mers. the proper set of realistic congurations is to start with

9 the full set of random-walk polymers and then discard the self-intersecting ones. (This is the same discarding process we used above in our discussion of interactions between fractals.) The fraction of the original congu- rations remaining is the probability that the polymer is self-avoiding. We wish to know how much the remaining self-avoiding polymers dier in their properties from the full set of random-walk polymers.

FIG. 6. A two- chain with no avoidance between the two dimers (left) and with avoidance (right). Some possible positions of the right half are sketched. The average end-to-end distance is a factor 2 larger in the self-avoiding case.

After all the intersections from the various stages above We may begin with the full set of random-walk con- have been discarded, we have a completely self-avoiding gurations and then start examining them for self in- chain. On the other hand, each stage has expanded the tersections. For convenience we imagine that our poly- overall size by a factor . This expansion factor is (ex- mer contains a number of monomers n that is an integer cept perhaps for the earliest stages) independent of the power of two: n =2K . We group the n monomers of stage k. Thus the overall expansion factor is K , where the chain into n/2 pairs. We group each two adjacent K is the total number of stages. We conclude that the pairs into a quartet, thus forming n/4 quartets. Then nal size R is related to the original random-walk size K 1/2 we group each two adjacent quartets into an octet, to R0 by R = R0 . Using R0 =(constant)n and form n/8 octets. We continue in this way (K times) K = log n/ log 2, we have R =(constant)n1/2+log / log 2. until we’ve made one nal pairing containing the rst Since a fractal obeys R = constant n1/D, our self- half and the last half of the monomers. Now we discard avoiding chain acts like a fractal of dimension D = self-intersecting congurations in stages. First we dis- 1/(1/2 + log / log 2). This argument leads us to ex- card those in which a monomer intersects with its paired pect that self-avoidance maintains the fractal nature of a monomer. Non-intersecting pairs are somewhat larger polymer, but decreases its fractal dimension. These ar- on the average than unrestricted pairs. But since inter- guments are borne out by more detailed work (Jannink sections are still freely allowed for other monomers, the et al. 1992). overall structure of our chain is still a random walk. It The fractal dimension D must not decrease too much. is merely expanded by some factor because all the ele- If it became smaller than the dimension of mutual opac- mentary pairs within it have been expanded by the same ity, D =3/2, the consistency of our arguments would factor as a dimer. We call this factor 1. Next we dis- break down. For then, when we imposed self avoidance card congurations in which a dimer intersects with its between the two halves of a segment, these two halves paired dimer. Figure 6 shows the result for two dimers would not be mutually opaque. Their probability of in- making a quartet. The eect on an isolated quartet is to tersection would be indenitely small. Accordingly, the expand it by some modest factor 2. The overall chain fraction discarded would be small, and the discarding is expanded by the same factor, leading to a cumulative process would have virtually no eect on the size of the expansion from both stages of 12. After many such segment. The expansion factor would approach 1. This stages, we must consider chain-sections with m mono- would be true for virtually all the stages k, so that the mers. Each half of the section already avoids itself, but above formula would lead to an unmodied D. D would the two halves do not avoid each other. We now dis- remain 2 as in a simple random-walk polymer. That is, card all congurations in which the rst half intersects the assumption that D<3/2 leads to the conclusion the last half. We have seen above that the probability that D = 2: a contradiction. We are thus forced to the of intersection is substantial if the two halves are fractals conclusion, rst noted by Des Cloizeaux (Des Cloizeaux with D>3/2. Congurations with a smaller size will be 1970), that 2 >D>3/2. The actual properties of self- more likely to intersect than those of a larger size. Thus avoiding polymers have been studied by detailed calcula- when the intersecting ones are discarded, the remaining tions, by computer simulations, and by measurements on ones have a larger average size. The expansion factor k real polymer solutions. All these approaches conrm that at the k-th stage is nite and limited even for arbitrar- these polymers, like random-walk polymers, are fractals. ily large segments. i.e., it is insensitive to the stage k. They have a D very close to 5/3—a value consistent with Accordingly, we replace it by with no subscript. Des Cloizeaux’s inequality.

10 Since real polymers are fractals just as random-walk pictured in Figure 1, the solvent cyclohexane near 351 polymers are, the properties we deduced for general frac- degrees Centigrade is a well-known theta solvent (Jan- tals apply to real polymers. Since D>3/2, they are nink et al. 1992). In practice, an increase of temperature opaque to one another. Thus they increase osmotic pres- by only 10 degrees from the theta temperature can dou- sure as though they were hard spheres of radius Rt com- ble the volume of a long polymer and thus increase the parable to their average geometric size R. Since real solution viscosity signicantly. Solvents with attraction polymers have D>1, they are also opaque to dius- weaker than this compensating amount are called “good ing substances and to ow. They experience hydrody- solvents.” All good solvents show some net self repul- namic drag and enhance viscosity as though they were sion. Since the self repelling pieces are mutually opaque, solid spheres roughly as large as their geometric size R. the probability of discarding becomes independent of the Large real polymers have hydrodynamic radii Rh and repulsion strength. The expansion factors k reach the viscosimetric radii Rv that is a xed multiple of the geo- same value for large segments (large k) even when some metric size R. These multiples depend only on the frac- attraction is present. Thus for all “good solvents” the tal structure, and not on the specic polymer: they are large-scale behavior13 is that of a self-avoiding walk, with the same for all suciently large polymers. Measure- D ' 5/3. ments on a number of polymers and solvents have es- If this cyclohexane is cooler than 35 degrees, the at- tablished the “Graessley ratios”12 (Davidson et al. 1987) traction becomes stronger than the compensating value. Rt : Rh : Rv : R =1.01:1:1.03:1.9. Polymers in such a solvent become progressively smaller A real polymer must always avoid itself in any solvent. than the random-walk size: they have a D that is larger Yet a real polymer need not always behave as a self- than 2. Such polymers attract one another in solution as avoiding walk with D ' 5/3. This is because the basic well as attracting themselves. Thus it is dicult to iso- eect of self avoidance can be modied by other eects. late them and measure their D denitively. Such solvents In some solvents the monomers feel an eective attrac- are called “poor solvents”. tion for each other: they are more likely to be near each other than if they were placed in the solution at random. We must take account of this attraction as well as the self VI. SOLUTIONS OF MANY POLYMERS avoidance when we consider the expansion of a polymer. For example, in the tetramer of Figure 6, we saw that throwing out the self-intersecting congurations makes The properties of individual polymers are strongly de- the average size expand. But because of the attraction, termined by their fractal structure, as we have seen monomers which are close together like the shaded ones above. This is even more true as we increase the con- in Figure 6 should be given extra weight compared to centration of our polymer solution from the dilute limit monomers that are far apart. This extra weight tends studied above. As we shall see, the fractal structure ac- to reduce the average size of the pair. It can happen counts for the oozing, springy ow of substances like mu- that the reduction in size caused by the attraction just cus and . For simplicity in what follows we compensates for the increase in size caused by the self- shall consider only good solvents, containing polymers avoidance, so that the expansion factor is 1. When with D =5/3. As the number of polymers per unit vol- this happens for many stages k, the attraction completely ume cp increases, so does the osmotic pressure . As we 1 compensates for the repulsion and the fractal dimension have seen, /(kTcp) ' 1+ Vtcp. We saw that the con- 2 D remains 2 like a random-walk polymer. Naturally, this centration cp at which the second term becomes equal compensation requires a very specic amount of attrac- to the rst is the concentration at which the pervaded tion between the monomers. Thus even a small change volumes of the dierent coils begin to overlap. It was in temperature can destroy the compensation. A solvent thus called the overlap concentration. The associated and temperature for which the compensation is perfect is volume fraction is roughly the internal volume frac- 4/3 called a “theta solvent”. For example, for the polystyrene tion i ' (R/a) within a coil. The work done to

12The geometric measure R used here is the radius of a solid sphere that would have the same root-mean-square distance between arbitrary monomers that the polymer has. The values reported are for high-molecular-weight polystyrene in . Reported values for other polymers in good solvents dier from these by several percent; this variation is consistent with experimental error. Several of these lengths are equal within experimental error, but there is no theoretical reason to believe they are exactly equal. 13Naturally, chains in solvents close to the theta state must become very long in order to attain the asymptotic self-avoiding behavior. The chain size required for self-avoidance to become signicant is called the “thermal blob” length t (Daoud et al. 1976). Chains smaller than a thermal blob size are approximately random-walk chains. Even chains much larger than this size behave as simple random walk fractals with D = 2 on length scales r < t. This t goes to innity as theta-solvent conditions are approached.

11 compress a solution to is roughly the osmotic pres- more than doubled, in accordance with self-avoiding walk sure times the volume . Thus a sizeable fraction of behavior. this work is due to the interactions between polymers. Though our interpenetrating chains have the overall ' This work is called the free energy. Since kT cp, structure of a simple random walk, the original -chains the free energy is roughly a thermal energy kT for each that were joined to create the nal chains have their orig- chain. For typical polymers, with a molecular weight of inal self-avoiding structure. Semidilute chains thus obey 105, this characteristic pressure is of the order of 10 3 two dierent fractal laws. The amount of chain n(r) en- atmosphere. closed in a radius r grows as r5/3 if a . But for 2 The overlap concentration cp is an important divid- r R, n(r) grows as r . The -chains are of- ing line in polymer solutions. Solutions much less con- ten called “blobs” (Daoud et al. 1975). The density of centrated than cp, are eectively dilute and interactions monomers around a given monomer is shown in Figure 7 between the polymers are unimportant. Solutions much . The chains of a semidilute solution each come in con- more concentrated than cp have strong interactions be- tact with arbitrarily many others. The various chains are tween the polymers. They are strong enough to distort strongly entangled with one another. The result resem- the polymer coils substantially and to bring about coop- bles a network of random strands. erative motions of many chains. As the volume fraction increases above , the chains interpenetrate. Concentrations much larger than but still much smaller than unity are called “semidilute.” A 1 chain in the semidilute regime interacts strongly with many other chains. Accordingly, its structure is much al- tered from the self-avoiding walk structure of a dilute chain. To understand this structure, it is convenient log(φ(r)) -4/3 to build the semidilute solution by joining chains orig- inally at the overlap concentration . These primary chains have an internal volume fraction i roughly equal to the overall volume fraction of the solution. This φ means they have a particular size satisfying ' i ' -.8 (/a) 4/3. We choose an initial -chain at random and -1 then connect one of its ends to the nearest available end from another chain. Since the initial chains were at , ξ the two ends to be joined are no farther apart than about a log(r) a chain-size R. Thus the joining does not distort the two chains very much. After doubling all the chains in this way, we repeat the process so that each chain is quadru- pled in length. With sucient repetitions we may in- crease the chain length as much as desired, while keeping FIG. 7. Local volume fractions of monomer (r) at dis- the volume per chain constant. This joining process thus tance r from an arbitrary monomer in a semidilute solu- takes us further and further above . tion of very low concentration . Solid curve, monomers These chains interpenetrate strongly, with many chains from the chain going through the origin. Dashed curve, occupying the pervaded volume of each. All these chains volume fraction of other chains. Scales are logarithmic, must avoid each other and not intersect. But unlike di- so that power laws appear as straight lines. Within a lute chains, these interpenetrating chains gain nothing blob size the total volume fraction is dominated by the by expanding into a self-avoiding-walk structure. To see chain passing through the origin. Beyond the distance this, we consider doubling one of the chains by joining two the total volume fraction is dominated by other chains. nearby ends together. The two pieces joined have no spe- cial alignment with each other; their end-to-end vectors From the joining construction above we can also under- are unrelated to each other—they must each avoid many stand the osmotic pressure in a semidilute solution. As other chains as well as avoiding each other. Thus when noted above, the pressure is important because it counts the two are joined, the two halves remain uncorrelated, the amount of free energy stored in the solution. Be- as in a random walk discussed in the rst section. The fore the joining process, we have seen that this energy mean-squared end-to-end distance is just double that of is about kT per chain. When the rst two chains are each half. This argument holds for all subsequent dou- joined together, the distance to the surrounding chains blings, as well. Thus the long, interpenetrating chains is not much changed. Thus the interaction energy with are like simple random walks, with D = 2. This con- these chains is about the same before and after joining. trasts with the situation in a pure solvent. There, if This continues to be true as the joining process continues. two chains are joined, the two halves tend to be aligned, Thus after this process, the interaction energy remains and the mean-squared end-to-end distance is signicantly about kT per joined segment, as it was originally. That

12 is, the interaction free energy and osmotic pressure are of FIG. 8. Osmotic compressibility vs. volume fraction for order kT per blob14. Expressed in terms of the volume various polymers and solvents. Osmotic compressibility fraction , we may write ' kT/3 ' (kT/a3)3/(3D). means change of osmotic pressure with volume fraction. This prediction that 9/4 and is independent of Horizontal scale is volume fraction , relative to . Ver- chainlength is well veried experimentally, as shown in tical scale is osmotic compressibility relative to its value Figure 8. We may readily remove all the solvent from in the dilute limit. The scales are logarithmic. Points our solution and increase the volume fraction to unity. are combined data on three samples of poly -methyl Sometimes removing the solvent causes crystallization of in toluene, diering by a factor of ve in molecu- the polymers, or it immobilizes them in a glassy struc- lar weight(Noda et al. 1981), and one sample of polyiso- ture. But at high enough temperature the chains in this prene () in cyclohexane(Adam 1988). The solvent-free state remain in the liquid state called the superposition of this data suggests that all long polymers melt. In this melt state the blob size has been reduced in good solvents have the same osmotic-pressure behav- to a monomer size a. The density at all distances from a ior. The line indicates the expected 5/4 behavior in the given monomer is as uniform and constant as in a small- semidilute regime. molecule liquid. The chains are then simple random walks at all length scales much larger than a monomer. The osmotic pressure in this state is dicult to dene, VII. DYNAMIC RELAXATION IN POLYMER since all the solvent has been removed. It is progressively SOLUTIONS more dicult to remove the last traces of solvent; in this sense the osmotic pressure becomes large on the scale of kT per monomer. The most important and obvious property of a poly- mer solution is the way it ows. A polymer solution ows in a springy, syrupy way quite dierent from the ow of a simple liquid like water or . To understand these distinctive ow properties, we must understand how the polymers move through the liquid. In a quiescent dilute solution a polymer, like any other molecule, moves via diusion: the motion is a random walk, whose mean squared distance x2 covered is the diusion constant times the elapsed time t, as noted in Section IV. The various chains are far apart and thus they do their random motions independently. As the concentration approaches , this ceases to be true. The random currents moving one polymer are also felt by its neighbors, so that nearby chains have similar motions. To characterize the motion we must specify two diu- sion coecients: the self-diusion coecient and the co- operative diusion coecient. The self diusion coe- cient s is dened by the motion of an individual chain. 2 100. normalized s x /t, as introduced above. Neighboring chains osmotic impede the random currents near a given chain; thus 50. compressibility s decreases as the concentration increases. The other important aspect of diusion is the spreading of extra 10. local concentration in the solution. The extra material spreads over a distance x whose square is proportional to 5 time. The “cooperative diusion coecient” c gives the 2 constant of proportionality: c = x /t. A small region 1 of extra concentration is under extra osmotic pressure. 0.1 0.5 1 5 10. This extra pressure, due to the repulsive interaction be- φ/φ∗ tween the polymers, tends to spread the chains in the con- centrated region apart faster than they would otherwise

14In this argument we ignored the interactions between the segments being joined. It is correct to ignore these, because the osmotic pressure and the work to compress the solution to the given concentration are both expressed relative to the dilute state. In the dilute state the nal chains have already been joined together. Thus it does not contribute to the work of compression and should not be counted.

13 spread. Thus the interactions increase c. And thus as connections generate internal forces amongst the blobs. the concentration increases from zero, c increases from But such forces don’t aect the center-of-mass motion. its limiting value of , just as s decreases. Thus they don’t aect the diusion constant. The self At semidilute concentrations, these two diusion con- diusion along its contour is K times smaller than that stants become qualitatively dierent. If the chains in a of a disconnected blob. semidilute solution are more concentrated in a given re- From this contour diusion constant, we may deduce gion, the extra concentration can spread by having the the time for a chain to diuse its own contour length solvent move into the region through the entangled net- K. This time is called the “ time” rep.We 2 work of polymers. This ow is impeded by the polymers may express it as rep ' (K) K/blob. This rep is evi- just as it is near an isolated polymer. The ow must go dently proportional to the cube of the chainlength; it can around the chains rather than going uniformly through easily reach tens of seconds, even without especially large them, as we saw above. This impeded ow depends on molecular weights. From this rep we can deduce the self- how the monomers collectively produce drag. This drag diusion constant.√ In the time rep the chain moves its is about the same in the semidilute solution or in a sim- own size R ' K as it moves its own length along the ilar one obtained by cutting the blobs apart from one contour. The self-diusion constant can thus be written 2 2 another. But with the blobs cut apart in this way, the s ' R /rep ' blob/K . It is much smaller than the solution is now at the overlap concentration . Here cooperative diusion constant c, which is roughly the we can readily estimate the cooperative diusion coe- size of blob. It is also much smaller than the diusion cient. It is not greatly dierent from the dilute diusion constant in the dilute state where the chain has size constant of an isolated blob. We saw in Section IV that R0: = blob(/R0). this diusion constant is inversely proportional to the ra- With these dynamical responses in mind, we can ac- dius : blob = kT/(6s). We have deduced in the last count for the most important type of relaxation in a section that the blob size decreases as the concentra- polymer solution: the relaxation of mechanical stress. ' 3/4 tion increases: a . Combining, we infer that This stress arises, e.g. , when the solution is stretched ' 3/4 c kT/(sa) . The cooperative diusion becomes along some axis as it would be if poured from a vessel. faster as the concentration increases. But joining the When this stretching is done more rapidly than rep, the chains to make them longer has no eect on the cooper- polymer chains cannot disentangle during the stretching. ative diusion. Each chain is obliged to elongate with the sample as a To deduce the self-diusion coecient s we must ask whole, and thus becomes distorted. With each successive how a given polymer coil moves in the semidilute envi- blob, a given chain encounters other chains that constrain ronment. Its motion is strongly impeded because it is it. Thus if the sample is stretched by a factor of two, each entangled with many other chains. To understand the blob is obliged to elongate by roughly a factor of two as diusion, we must understand the disentanglement pro- well. The force exerted to cause the stretching supplies cess. There is one way for a chain to move that does not the energy to distort these blobs. The work required to require the motion of other chains. It may simply move pull the ends of any free polymer a distance compara- along its own contour as a worm does. This “reptation” ble to the chain size is of the order of the thermal energy motion arises from the independent Brownian motions of kT. Since each blob is elongated to this degree, each blob the dierent blobs composing it. It is simplest to visual- stores about kT of energy under a unit elongation. If our ize this motion of a chain along its contour by imagining solution is rapidly released from this stretched position, that the sequence of K blobs making up the chain lie in the restoring forces on the chains bring them back to their a straight line. Each blob is subjected to independent undistorted shape; the sample springs back elastically to random forces. Considering the combined eect of all of its original shape, like a solid. This is the origin of the these, there is some net tendency at a given moment for of rubber. It is also the cause of the springy be- the chain to move to the left or to the right. It is sim- havior of a raw egg white, of viscous uids in our body, plest to consider the motion of the center of mass, since and of many other liquids in daily life. this center can only depend on the external forces on the Now if the sample is not released but is held in its chain. The total external force is the sum of K inde- stretched position, the chains move by the same repta- pendent random forces, one for each blob. The external tion responsible for self diusion. As they move, they forces are the same whether the blobs are connected or are able to disengage from each other. The entangle- not. If they are not, then each blob diuses indepen- ments that initially constrained the chains unravel and dently with diusionP constant blob. The center-of-mass the chains become undistorted again. The time required has position x = k xk/K. The mean-squared center- for this is the time for a chain to reptate out of its initial 2 of-mass position x is the average of those of its blobs: constraints: i.e., the reptation time rep. Thus the initial 2 2 x = xk /K. Thus the diusion constant of the cen- applied force relaxes on this time scale. Now if the sam- ter of mass is reduced from that of a blob by a factor of ple is released, there is no restoring force and it keeps its the number of blobs in a chain K. Now we may read- distorted shape. Over this time scale it has owed like a ily account for the eect of connecting the blobs. The liquid rather than springing back like an elastic solid.

14 VIII. CONCLUSION Volume fraction: the fraction of the uid’s volume occupied by polymers or other solutes. In their atomic makeup the polymer solutions we have D Fractal dimension from the formula for the distri- considered dier very little from a simple liquid like wa- bution of matter in a fractal object: n(R) ∝ RD ter. Water consists of a mass of small molecules that may move about freely, subject to the lenient constraints of g(r) Pair distribution function describing the number the liquid state. In a polymer solution almost all of this of particles at a displacement r from a given parti- freedom remains. Only a small fraction of the atoms are cle. linked in sequence so that their mutual separations are xed. These sequences are the exible polymer chains. Volume of the solution being studied. We have seen that even a small fraction of these chains changes the macroscopic properties of the liquid dramat- N Number of solute objects in the solution being stud- ically. It creates large spatial structures which are very ied. eective at modifying ow, transmitting stress and in- creasing viscosity. In addition these structures store me- Vt Thermodynamic volume or excluded volume: the chanical energy in a convenient form over macroscopic volume rendered inaccessible to the solute objects times. in a solution by the addition of one further solute The polymer phenomena sketched here now constitute object. a mature eld of science. Most of the phenomena are well Rt Thermodynamic radius: radius of a sphere whose documented experimentally and are understood theoret- volume is Vt. ically. They reveal a type of matter that is distinctive, controllable, and understandable from a few basic princi- cp Number concentration of solute: N/. ples. This understanding now forms the basis for a wealth of new phenomena involving polymers. Polymer chains Osmotic pressure in a solution. in new architectures such as branches, combs, and loops change the nature of entanglement and disentanglement. kT Boltzman constant times temperature. Polymers in conned geometries—adsorbed or grafted to cp Overlap concentration, at which the osmotic pres- surfaces, in thin layers, or in layers— sure rises to twice its ideal-solution value. produce new types of ow and interaction with bulk uid. Heterogeneous polymers with mutually immiscible parts Viscosity of a solution. separate on a molecular scale, forming controllable, periodic patterns. The elements of fractal structure, in- Rh Hydrodynamic radius: radius of a sphere with the terpenetration, and deformability that make simple poly- same drag coecient as a given solute object. mers interesting and useful promise also to create further surprising behavior in these new situations. Diusion constant. Shear strain or shear.

ACKNOWLEDGMENTS ˙ Shear rate: rate if increase of shear per unit time.

This research was supported in part by the Na- Rv Viscometric radius: radius of here that increases tional Science through award number DMR the uid visosity the same amount as given solute 9528957 object does.

t Size of a segment of a polymer above which self- IX. KEY TO SYMBOLS (IN ORDER OF avoidance eects become important. APPEARANCE) Overlap volume fraction: Volume fraction of a so- lution whose solute particles are at a concentration n Number of elementary segments in a polymer. cp. h i ... Means average over some uctuating quantity: Blob size: geometric size of polymers that would be a Length of a step in a random-walk polymer. at their overlap concentration in a solution of given volume fraction. R Radius of imaginary sphere used in dening fractal properties. Radius of the truncated fractal objects (r) Local volume fraction at a distance r from an dened in the text. arbitrary monomer.

n(R) Average number of chain segments within dis- s Self-diusion constant dened from measurements tance R of an arbitrary segment. of a single object’s displacement over time.

15 c Co-operative diusion constant dened from mea- Davidson, N.S., L. J. Fetters, W.G. Funk, N. Hadjichristidis, surements of the density of solute objects over time. and W. W. Graessley, 1987, 20 2614 Des Cloizeaux, J., 1970, J. de Physique 31 715 Reptation time: time for a polymer to move a Einstein, A., 1906, Ann. Physik 17 549 (1905) ; 19 371 rep Elias, Hans-Georg, 1984, Macromolecules 2nd edition, New distance equal to its radius through the tangle of York: Plenum Press p124. other polymers constraining it. Happel, John and Howard Brenner, 1973, Low Reynolds num- ber hydrodynamics, with special applications to particulate media 2d rev. ed. Leiden, Noordho International Publish- ing Jannink, G. and J. Des Cloizeaux, 1992, Polymers in Solution (Oxford). Mandelbrot, B. B., 1982, The Fractal Geometry of Nature (Freeman, San Fransisco). Adam, Mireille, 1988, private communication Mondello, M. , Hyung-Jin Yang and R. J. Roe, 1991, Simu- Daoud M., J. P. , B. Farnoux, G. Jannink, G. Sarma, lation at the University of Cincinnati using the Cray Y-MP H. Benoit, R. DuPlessix, C. Picot, P. G. DeGennes, 1975, at the Ohio Supercomputer Center, private communication. Macromolecules 8 804 Noda, I., N. Dato, T. Kitano, and M. Nagasawa, 1981, Macro- Daoud, M., and G. Jannink, 1976, J. Physique 37 973 molecules 14 668 Daoud, M. 1995, in La Juste Argile, M. Daoud and C. Onsager, L., 1949, Ann. N. Y. Acad. Sci. 51 627. Williams, eds. (les editions de physique, Les Ulis France) Reif, Fredrick, 1965, Fundamentals of statistical and thermal DeGennes, P. G., 1979, Scaling Concepts in Polymer Physics, physics, New York, McGraw-Hill. Cornell, Ithaca. Roux, S., 1987, J. Stat. Phys. (USA) 48 201. DeGennes, P. G., 1982, C. R. Seances Acad. Sci. 295 1061. Witten, T. A. and M. E. Cates, 1986, Science, 232 1607.

16