1:
2: \documentclass[12pt]{iopart}
3: %\newcommand{\gguide}{{\it Preparing graphics for IOP journals}}
4: %Uncomment next line if AMS fonts required
5: %\usepackage{iopams}
6: \usepackage{graphicx}
7: \begin{document}
8:
9: \title[Polymer solutions]{Polymer solutions: from hard monomers to soft polymers}
10:
11: \author{Jean-Pierre Hansen, Chris I. Addison and Ard A. Louis}
12:
13: \address{Dept. Chemistry, Lensfield Rd, Cambridge, CB2 1EW (UK)}
14: \begin{abstract}
15: A coarse-graining strategy for dilute and semi-dilute solutions of
16: interacting polymers, and of colloid polymer mixtures is briefly
17: described. Monomer degrees of freedom are traced out to derive an
18: effective, state dependent pair potential between the polymer centres
19: of mass. The cross-over between good and poor solvent conditions is
20: discussed within a scaling analysis. The method is extended to block
21: copolymers represented as ``necklaces'' of soft ``blobs'', and its
22: success is illustrated here in the case of a symmetric diblock
23: copolymer which exhibits microphase separation.
24: \end{abstract}
25:
26: %Uncomment for PACS numbers title message
27: %\pacs{00.00, 20.00, 42.10}
28: % Keywords required only for MST, PB, PMB, PM, JOA, JOB?
29: %\vspace{2pc}
30: %\noindent{\it Keywords}: Article preparation, IOP journals
31: % Uncomment for Submitted to journal title message
32: %\submitto{\JPA}
33: % Comment out if separate title page not required
34: \maketitle
35:
36: \section{Introduction: a coarse-graining strategy}
37:
38: The present paper summarises a collective effort by the Cambridge
39: group and others to bridge the ``cultural divide'' between statistical
40: mechanics of polymer solutions and melts on one hand and of ``simple
41: liquids'' on the other. While the former is dominated by
42: field-theoretic methods and scaling concepts pioneered by S.F. Edwards
43: and P.G. de Gennes, the description of simple liquids, which lack the
44: scale invariance of polymers, requires a more atomistic approach. To
45: bridge the gap we have developed a systematic coarse-graining strategy
46: first suggested by Flory and Krigbaum\cite{flory}, whereby individual
47: monomer degrees of freedom are traced out for fixed centre-of-mass
48: (CM) positions of the polymer coils, thus defining an effective
49: interaction between the CMs. Consider a system of N identical
50: polymer, each consisting of M monomers (or segments) at positions
51: $\vec r_{i \alpha}$ $(1\leq i \leq N ; 1\leq \alpha \leq M)$. If
52: $H(\{\vec r_{i \alpha}\})$ is the interaction hamiltonian of the
53: system, the probability distribution of the N CMs at positions
54: $\vec R_i$ is
55: \begin{eqnarray}
56: \label{eq1}
57: P_N(\{\vec R_i\})&
58: =\frac{1}{Q_N} \int e^{-\beta H(\{\vec r_{i \alpha}\})} \prod_{i=1}^{N}\delta(\vec R_i - \frac{1}{M}\sum_\alpha \vec r_{i \alpha}) \prod_{i=1}^{N} \prod_{\alpha=1}^{M} d \vec r_{i\alpha}\\
59: &=\frac{e^{-\beta V_{eff}(\{\vec R_i\})}} {\int e^{-\beta V_{eff}(\{\vec R_i\})} \prod_{i=1}^{N} d \vec R_{i}}
60: \end{eqnarray}
61: where $\beta=\frac{1}{k_BT}$, $Q_N$ is the partition function for the
62: $N\times M$ monomers, and the total effective interaction energy of
63: the N CMs is rigorously defined by:
64: \begin{eqnarray}
65: \label{eq2}
66: V_{eff}(\{\vec R_i\})=-k_BT \ln[C \times P_N(\{\vec R_i\})]
67: \end{eqnarray}
68: This effective energy is a free energy, and hence state-dependent,
69: and, in general, many-body in nature. In the low concentration limit,
70: (\ref{eq2}) reduces to a sum of effective pair interactions between
71: two isolated polymer coils:
72: \begin{eqnarray}
73: \label{eq3}
74: v_2(|\vec R_1 -\vec R_2|)=-k_BT \ln[C \times P_2 (\vec R_1,\vec R_2)]
75: \end{eqnarray}
76: Since global properties of polymeric systems are independent of
77: chemical detail in the scaling ($L\rightarrow \infty$) limit, we adopt
78: henceforth a simple lattice model of polymer solutions namely that of
79: mutually and self-avoiding walks (SAW) of length $L=M-1$ on a cubic
80: lattice of lattice spacing $b$ (equal to the segment length);
81: non-connected nearest neighbour monomers of the same or different
82: polymer coils have an attractive energy $-\epsilon$. This model
83: accounts for the key polymer features, namely connectivity, excluded
84: volume and solvent quality (through the value of $\epsilon$).
85:
86: Convenient thermodynamic variables are the polymer density
87: $\rho=N/(\Omega b^3)$ ($\Omega$ being the size of the lattice), the
88: monomer density $c=M\rho$ (equivalently the monomer packing fraction
89: $\phi=cb^3$ equal to the number of lattice sites occupied by monomers)
90: and the temperature T ($\beta^*=\epsilon/k_BT)$. A key
91: characteristic is the overlap density $\rho^*=3/4\pi R{_g}{^3}$
92: (where $R_g$ is the radius of gyration, $\sim bL^\nu$) which corresponds to the
93: cross-over from the dilute ($\rho <\rho^*$) to the semi-dilute regimes
94: ($\rho >\rho^*$). The semi-dilute regime differs from the melt in
95: that the monomer packing fraction remains negligible; for any given
96: $\rho/\rho^*$ this is only achieved for sufficiently long polymers,
97: since $\phi=(\rho/\rho^*) L^{(1-3\nu)} \sim L^{-4/5}$ for SAW polymers.
98:
99: \section{Polymers in good solvent}
100: Consider first the athermal limit of SAW polymers ($\epsilon=0$). In
101: the low ($\rho\rightarrow 0$) density limit $P_2 (\vec
102: R_1,\vec R_2)$ is then simply the probability that there is
103: no monomer-monomer overlap for a fixed distance $r=|\vec
104: R_1 -\vec R_2|$ between the CMs. This is well adapted to
105: Monte-Carlo (MC) sampling. Scaling theory predicts that the resulting
106: effective interaction at full overlap, $v_2(r=0)$ is independent of
107: chain length L\cite{gros}. Early simulations with rather short
108: chains pointed to $v_2(r=0) \approx 2k_BT$\cite{daut}. An in-depth
109: investigation of the L-dependence shows that $v_2(r)$ is well
110: represented by a Gaussian $v_2(r) \approx u
111: exp(-\alpha(r/R_g)^2)$, where $\alpha$ is of order 1, while
112: $u/k_BT \approx 1.80$ \cite{bolh}\cite{pelis}.\\ At finite
113: polymer concentration $\rho$, three and more-body effective
114: interactions come into play\cite{bolh2}. A more efficient strategy is
115: to determine an effective density-dependent pair potential
116: $v_2(r;\rho)$ by inverting the CM-CM pair distribution $g(r)$ from full
117: monomer-level MC simulations\cite{bolh2}\cite{louis}. It was proven
118: that this inverse problem has a unique solution\cite{hend}. The
119: inversion procedure is implemented using the HNC-integral
120: equation\cite{hansen}. The Ornstein-Zernike relation\cite{hansen}
121: allows the extraction of the direct correlation function $c(r)$ from
122: the MC data for $h(r)=g(r)-1$ at any given density. The HNC closure then expresses $v_2(r)$ as \cite{hansen}:
123: \begin{eqnarray}
124: \label{eq4}
125: \beta v_2(r)=-\ln[g(r)]+h(r)-c(r).
126: \end{eqnarray}
127: The first term on the r.h.s. is the potential of mean force, while
128: $h(r)-c(r)$ describes the effect of correlations.\\
129:
130: The MC generated $g(r)$ show that correlations {\it decrease} as $\rho$
131: increases, contrary to the more familiar behaviour observed for hard
132: core systems. The overlap value $g(r=0)$ increases steadily toward 1
133: (the ideal gas value) confirming that in the high density limit of a
134: melt, polymer chains indeed behave as non-interacting
135: polymers\cite{rubin}. The resulting effective pair potential is only
136: moderately density dependent and is well fitted by a sum of gaussians.
137: The range of $v_2(r)$ tends to increase with $\rho$, and the potential
138: develops a small amplitude negative tail for $r$ significantly larger
139: than $R_g$\cite{bolh}\cite{bolh3}.\\
140: The link with thermodynamics is via the
141: compressibility relation\cite{hansen}, which allows the osmotic
142: pressure $P$ to be expressed as:
143: \begin{eqnarray}
144: \label{eq5}
145: \beta P(\rho)= \int_{0}^{\rho}[1-\rho'\hat{c}(k=0;\rho')]d\rho'
146: \end{eqnarray}
147: where $\hat{c}(k)$ is the Fourier transform of $c(r)$. Use of the
148: effective pair potential in conjunction with the virial and energy
149: equations is meaningless\cite{louis5}. The equations of state
150: calculated from MC simulations of the full monomer level polymer
151: representation and from simulations based on the much less
152: CPU-intensive effective potential representation agree within
153: numerical uncertainties, underlining the adequacy if the HNC inversion
154: procedure for such ``soft'' effective particles. Well into the
155: semi-dilute regime ($\rho \gg \rho^*$) the slopes of the calculated
156: equation of state agrees with the des Cloizeaux scaling prediction
157: $\beta P \approx \rho^{3\nu/(3\nu -1)} \approx \rho^{9/4}$, where
158: $\nu=0.588 \approx 3/5$ is the Flory exponent for the radius of
159: gyration in good solvent($R_g \sim bL^\nu)$\cite{rubin}.\\
160: Neglecting the density-dependence of $v_2(r)$, {\it i.e.} extending
161: the low density gaussian form to all densities, brings us back to the
162: ``Gaussian core model'' (GCM) first introduced by
163: Stillinger\cite{stil}, which exhibits interesting behaviour at low
164: temperatures ($\beta^* \gg 1)$\cite{lang}. In the regime relevant for
165: polymer solutions ($\beta^* \approx 1)$ the model leads to ``mean
166: field fluid'' behaviour at sufficiently high density, where the random
167: phase approximation, $c(r)=-\beta v_2(r)$ leads to a quadratic
168: equation-of-state for $\rho \gg \rho^*$\cite{lang}\cite{louis2}:
169: \begin{eqnarray}
170: \label{eq6}
171: \beta P=\rho+ \frac{1}{2}\beta \hat{v}_2(k=0)\rho^2
172: \end{eqnarray}
173: Incorporating the $\rho$-dependence of $v_2$ is thus seen to change
174: the asymptotic $\rho^2$ behaviour into des Cloizeaux scaling
175: $\rho^{9/4}$. As a by-product of the GCM, we have developed a
176: multiple occupancy lattice model, which also gives rise to interesting
177: microphase separation\cite{finken}.
178:
179: \section{From good to poor solvent conditions}
180: We now turn our attention to the case where adjacent monomers attract,
181: {\it i.e.} $\epsilon\neq 0$ or $\beta^*>0$. This
182: attraction is solvent induced, and the quality of the solvent
183: deteriorates as $\beta^*$ increases, leading to contraction of the
184: polymer coils. At the $\theta$ temperature ($\beta^*_\theta =
185: \epsilon/k_BT_\theta$), repulsion and attraction between polymers
186: cancel, at least in the low density limit, so that polymers exhibit
187: the scaling behaviour of ideal polymers($R_g \sim L^{1/2}$). Below
188: $T_\theta$, polymer coils collapse into globules ($R_g \sim L^{1/3}$),
189: and phase separation occurs into polymer-rich and polymer-poor
190: solutions. \\
191: The most convenient diagnostic for locating $T_\theta$
192: from simulations is to calculate the second virial coefficient
193: $B_2(L;T)$ as a function of temperature
194: and polymer length. The L-dependent Boyle temperature $T_B(L)$ is
195: that at which $B_2(L;T)$ vanishes for a fixed L, then:
196: \begin{eqnarray}
197: \label{eq7}
198: T_\theta=\lim_{L \rightarrow \infty}T_B(L)
199: \end{eqnarray}
200: This leads to the estimate $\beta^*_\theta=0.2690 \pm 0.0002$
201: \cite{grass}. Note that $B_2(L;T)$ can be directly expressed in
202: terms of the low density limit of the effective CM pair potential:
203: \begin{eqnarray}
204: \label{eq8}
205: B_2(L;T)=2\pi \int_0^\infty[1-e^{-\beta v_2(r, L, T)}]r^2dr
206: \end{eqnarray}
207: Extensive MC simulations were used to determine the effective pair
208: potential for fixed length $L=100$ over a wide range of
209: temperatures($0\leq\beta^* \leq 0.3$) and densities $\rho$\cite{krak}.
210: As $\beta^*$ increases, $v_2(r=0)$ decreases and $v_2(r)$ develops an
211: attractive tail for $r > R_g$. Eventually $v_2(r)$ violates Ruelle's
212: necessary condition for the existence of a thermodynamic limit,
213: namely\cite{ruel}:
214: \begin{eqnarray}
215: \label{eq9}
216: I_2=\int v_2(r) d \vec{r} >0
217: \end{eqnarray}
218: Since $v_2$ depends on $L$, $\rho$ and $T$, so does $I_2$, and for any given $\rho$ and $L$, the limit of stability temperature $T_s$ is determined by
219: \begin{eqnarray}
220: \label{eq10}
221: I_2(L,\rho, T=T_s)=0
222: \end{eqnarray}
223: Below $T_s$, single polymer coils will collapse and we conjecture that
224: $T_\theta=\lim_{\rho \rightarrow 0} \lim_{L \rightarrow \infty}
225: T_s(\rho, L) $\cite{krak}. However, if $\rho$ is increased at a given
226: temperature $T\leq T_\theta$ the effective pair potential increasingly
227: reverts to being repulsive until the Ruelle criterion\cite{hansen} is
228: satisfied, and the polymer solution becomes thermodynamically stable
229: again. This ``restabilisation'' reflects a phase separation scenario
230: under poor solvent conditions($T < T_\theta$). \\ The variation of
231: $v_2(r)$ with temperature and density, as extracted from the HNC
232: inversion of MC data, is semi-quantitatively reproduced by solutions
233: of the PRISM integral equation\cite{schw} for the thread
234: model($b\rightarrow0, L\rightarrow\infty$ at fixed $R_g$) of polymer
235: solutions\cite{fuchs}\cite{krak2}. PRISM assumes all monomers to be
236: equivalent ({\it i.e.} neglects ends effects) and yields the monomer
237: pair distribution function $g_{mm}(r)$ of a polymer solution. The CM
238: distribution function $g(r)$ required to extract the effective pair
239: potential $v_2(r)$ may be related to $g_{mm}(r)$ by an approximate,
240: but accurate relation involving internal form factors of a single
241: coil\cite{krak3}. Further progress along these lines might eventually
242: bypass the need for time-consuming simulations of full monomer level
243: models required to determine $g(r)$. \\ The equation of state may be
244: computed as a function of $\rho/\rho^*$ and $\beta^*$, using a
245: somewhat cumbersome method based on the contact
246: theorem\cite{dick}\cite{add}, or much more efficiently by subjecting
247: the polymers to a gravitational field and invoking hydrostatic
248: equilibrium\cite{add2}. If $\rho(z)$ denotes the CM(or monomer)
249: density profile of the polymers in a vertical field, which is easily
250: measured in simulations, the osmotic pressure at an altitude $z$ is
251: simply
252: \begin{eqnarray}
253: \label{eq11}
254: \beta P(z)= \frac{1}{\zeta}\int_z^\infty \rho(z')dz'
255: \end{eqnarray}
256: where $\zeta=k_BT/Mg$ is the gravitational length, which must be
257: chosen (by tuning the product $Mg$) significantly larger than $R_g$
258: for the macroscopic description (which follows from the local density
259: approximation within density functional theory of inhomogeneous
260: fluids) to hold\cite{barrat}. Elimination of the altitude $z$ between
261: $P(z)$ and $\rho(z)$ then leads to the bulk equation of state
262: $P(\rho)$. Examples from simulations of L=500 chains are shown, for
263: four temperatures in figure 1.
264: \begin{figure}[!htp]
265: \begin{center}
266: \includegraphics[width=8cm]{fig1.eps}.
267: \caption{Equation of state on a log-log scale, calculated using Dickman's(contact theorem) method, (symbols) and hydrostatic equilibrium(solid lines).}
268: \label{fig1}
269: \end{center}
270: \end{figure}
271: \section{Long polymers: corrections to scaling}
272: Polymers are critical objects in the universality class of the zero
273: component(n=0) limit of the n-vector model of critical
274: phenomena\cite{genne}. The $L\rightarrow \infty$ limit is equivalent
275: to the limit of divergent correlation length $\xi$ at the critical
276: point of a second-order phase transition. The properties of very long
277: chains can hence be investigated using the powerful method of the
278: renormalization group (RG) and field theory\cite{freed}. These
279: predict two scaling regimes, an athermal one corresponding to the good
280: solvent (SAW) limit, and the second corresponding to the
281: $\theta$-solvent regime where ideal polymer statistics hold. The
282: cross-over between the two regimes is discontinuous in the scaling
283: limit $L \rightarrow \infty$. The objective is to predict the
284: behaviour for large but finite L from a finite size scaling analysis
285: of MC data\cite{li}\cite{pelis}. We restrict the discussion to the
286: good solvent regime. For $\beta<\beta_\theta$ any universal
287: (dimensionless) ratio $R$ may be represented as
288: \begin{eqnarray}
289: \label{eq12}
290: R(L,\beta)=R^* +\frac{a_R(\beta)}{L^\Delta} + ...
291: \end{eqnarray}
292: where $R^*$ is the temperature independent scaling limit
293: ($L\rightarrow \infty$) and the exponent $\Delta=0.517$\cite{beloh}.
294: Higher order terms involve exponents of the order of 1 or larger. The
295: temperature-dependent coefficient $a_R(\beta)$ is non-universal, but
296: ratios of such coefficients for two different dimensionless qualities
297: are again universal, {\it i.e.} model-independent. A much studied
298: example of a universal ratio is
299: $A_2(L,\beta)=B_2(L,\beta)/R^3_g(L,\beta)$. A similar ratio involving
300: the third virial coefficient
301: $A_3(L,\beta)=B_3(L,\beta)/R^6_g(L,\beta)$ has been examined in detail
302: in reference\cite{pelis}. $B_3$ is found to be always positive, but
303: to go through a sharp minimum near $T=T_\theta$. The effective pair
304: potential between the CMs of two isolated polymers ($\rho \ll \rho^*$
305: limit), divided by $k_BT$, is another universal ratio, which has been investigated by a similar scaling analysis in \cite{pelis}:
306: \begin{eqnarray}
307: \label{eq13}
308: \beta v_2(r,L,\beta)=v_\infty(x) +\frac{a_v(\beta)}{L^\Delta}v_c(x) + ...
309: \end{eqnarray}
310: where $x=r/R_g$. $v_\infty$, $v_c$ and $a_v(\beta)$ are extracted
311: from a careful analysis of MC data for several lengths L and inverse
312: temperatures $0 \leq \beta < \beta_\theta$. These may then be used to
313: predict $\beta v_2$ for any length and temperature, and agreement with
314: MC data is excellent for $\beta^* \leq 0.2$ and $L\geq 500$. Closer
315: to the $\theta$-temperature the convergence of (\ref{eq13}) with L is
316: found, not surprisingly, to be much slower, and one must then switch
317: to the scaling analysis appropriate for the $\theta$-regime. The
318: scaling analysis has been recently extended to binary mixtures of
319: polymers of different lengths and to star polymers\cite{pelis3}
320:
321: \section{Soft polymers and hard colloids}
322: Mixtures of colloidal particles and non-adsorbing polymers have
323: attracted considerable experimental and theoretical attention over the
324: last two decades, because of interesting phase behaviour induced by
325: the familiar depletion mechanism\cite{asaku}. The effective depletion
326: interaction between spherical colloidal particles may be tuned by
327: varying the size ration $q=R_g/R_c$ (where $R_c$ is the colloid
328: radius), and the polymer concentration $\rho$. Consider first
329: the case of polymers between 2 plates($q=0$) separated by z. The
330: depletion potential per unit area $W(z)$ is defined by the difference
331: in polymer grand potentials:
332: \begin{eqnarray}
333: \label{eq14}
334: W(z)&=\frac{1}{A}[\Omega(z)-\Omega(z=\infty)]\\
335: &=\int_z^\infty[P(z')-P(z=\infty)]dz'
336: \end{eqnarray}
337: Clearly since at contact 2 depletion zones are destroyed
338: $W(z=\infty)=-2\gamma_W(\rho)$, where $\gamma_W$ is the polymer -wall
339: surface tension. The simple ansatz:
340: \begin{eqnarray}
341: \label{eq15}
342: W(z)&=W(0) + P(\rho)z &\;\;\;\;\;\; ; z<D_w(\rho)=-\frac{W(0)}{P(\rho)}\\
343: & =0 & \;\;\;\;\;\; ; z>D_w(\rho)
344: \end{eqnarray}
345: reproduces direct simulation data well. The result for SAW
346: polymers\cite{louis3} differs considerably from that for ideal
347: polymers\cite{asaku}: the range $D_W$ is shorter for interacting
348: polymers, and decreases with increasing density, while it is
349: independent of density for ideal polymers; the contact value $W(0)$
350: decreases faster with $\rho$ for interacting polymers since the
351: surface tension scales as $\rho^{3/2}$ in the semi-dilute
352: regime\cite{louis4} while it scales as $\rho$ for ideal polymers. For
353: finite $q\leq 1$, the depletion force between two spheres can be
354: approximately related to the depletion potential between 2 planes via
355: the Derjaguin approximation, and by correcting for the decreasing
356: range of the force due to partial wrapping of the polymer coils around
357: the spherical colloid\cite{louis3}.\\ Such pair depletion interactions
358: do not, however, account for effective many-body interactions due to
359: finite colloid concentrations. To that purpose the coarse-graining
360: strategy described in sections 1 to 3 may be extended to the two
361: component colloid-polymer system. An effective state-dependent
362: colloid-polymer pair interaction $v_{cp}(r)$ may be extracted by an
363: HNC inversion of the polymer density profile around a
364: sphere\cite{bolh}\cite{bolh3}\cite{louis3}, similar to the inversion
365: procedure used to determine the effective polymer-polymer pair
366: potential $v_{pp}(r)$. The colloid-colloid pair potential $v_{cc}(r)$
367: is well approximated by a simple hard sphere interaction. MC
368: simulations of this effective two-component system were used to
369: calculate the phase diagram of the mixture for $q \leq 1$
370: \cite{bolh4}. The calculated binodal agrees very well with
371: experimental data\cite{rama} for interacting polymers. Significant
372: qualitative differences arise between phase diagrams for
373: ideal\cite{lekker} and interacting polymers, particularly for larger
374: $q$: the range of the concentrated (``liquid'') colloidal phase in
375: the colloid density-polymer density plane is considerably reduced when
376: polymer interactions are included, and the critical point occurs at
377: significantly higher packing fraction of the two species\cite{bolh4}.
378: These trends become more pronounced in the ``protein limit'' of large
379: polymers and small colloids ($q\gg 1$)\cite{bolh5}. Solvent quality
380: has a strong influence on the induced depletion interaction between
381: colloids, with polymers under $\theta$ conditions leading, not
382: surprisingly to a pair interaction close (but not identical) to that
383: induced by ideal polymers\cite{add}, at least for $\rho/\rho*<1$.
384:
385: \section{Diblock copolymers and beyond}
386: The coarse-graining strategy may be extended to polymers other than
387: linear homopolymers considered so far. Star polymers, for instance,
388: have been investigated along similar lines\cite{likos}; the CM is
389: replaced by the midpoint where the $f$ arms of a star polymer meet;
390: the resulting effective pair interaction diverges logarithmically as
391: $r\rightarrow0$ and hardens as $f$ increases. \\ The case of symmetric
392: diblock copolymers AB has been examined very recently\cite{add3}. The
393: A and B strands are represented as soft ``blobs'', the CMs of which
394: are tethered by an ``entropic spring'' characterised by the
395: intramolecular potential $\phi_{AB}(r)$ which is derived from the
396: MC-generated distribution function of relative distances of the A and
397: B CMs on the same copolymer. There are now three intermolecular CM
398: potentials $v_{AA}(r)$, $v_{BB}(r)$ $v_{AB}(r)$, in addition to the
399: intramolecular potential. The inversion procedure to go from the
400: partial distribution functions $g_{\alpha \beta}(r)$ to the pair
401: potentials $v_{\alpha \beta}(r)$ is now more involved\cite{add3}.
402: Even in the low density limit one already faces a four-body problem.
403: The inversion procedure has been carried out in that limit for a
404: simple athermal model, the ISI model, where A-A and B-B pairs behave
405: like ideal polymers, {\it i.e.} freely interpenetrate, while A-B pairs
406: behave like mutually avoiding walks. This is the block copolymer
407: equivalent of the familiar Widom-Rowlinson model\cite{widom} which
408: drives phase separation of simple atomic fluids (where A and B are
409: untethered). Because the two strands of the AB block copolymer system
410: are tethered, macroscopic phase separation due to A-B incompatibility
411: is suppressed and reduces to microphase separation: the symmetric
412: block copolymers form a lammelar phase, which was indeed observed in
413: MC simulations of both the full monomeric and the coarse-grained
414: representations\cite{add3}. The resulting equation of state calculated
415: by the hydrostatic equilibrium method is shown in Figure 2.
416: \begin{figure}[!htp]
417: \begin{center}
418: \includegraphics[width=8cm,angle=0]{fig2.eps}
419: \caption{Equation of state generated from MC simulations for diblock copolymers
420: using full monomer level and coarse-grained ``blob'' models.}
421: \label{fig2}
422: \end{center}
423: \end{figure}
424: $Z=\beta P/\rho$ is seen to first increase linearly up to
425: $\rho/\rho^*\approx2$, where it flattens out, and thereafter decreases
426: slowly to give an asymptotic value $>1$. This may be understood by
427: noting that in the lamellar phase(which develops for $\rho/\rho^*>2$),
428: the repulsive A-B contacts are greatly reduced. Figure 2 also shows
429: the equation of state calculated (with much less computational effort)
430: for the coarse-grained ``soft-dumbbell'' model with effective intra
431: and inter-molecular pair potentials $\phi_{AB}(r)$ and $v_{\alpha
432: \beta}(r)$ determined in the zero density limit. The agreement is
433: excellent up to $\rho\approx\rho^*$, and remains semi-quantitative
434: thereafter, despite the fact that the density dependence of the
435: effective potential has not been taken into account.
436:
437: The ``soft dumbbell'' representation of diblock copolymers provides a
438: hint of how to extend the coarse-graining strategy of linear homo or
439: hetropolymers over a wide range of polymer concentrations. As the
440: ratio $\rho/\rho^*$ increases, the fundamental length scale gradually
441: crosses over from $R_g$ (for $\rho/\rho^*<1$) to the correlation
442: length $\xi \sim \rho^{-3/4}$ deep in the semi-dilute regime, to the
443: segment length $b$ in the melt. The coarse-graining strategy put
444: forward in this paper applies to dilute and initial semi-dilute
445: regime, where polymer coils are well represented by a single
446: ``soft-core'' particle with a radius of the order $R_g$. Deeper into
447: the semi-dilute regime, the blob picture\cite{genne2} applies for
448: polymers confined by other polymers or in a pore; each polymer reduces
449: to a ``necklace'' of blobs of radius $\xi$, tethered by entropic
450: springs; different blobs on the same or neighbouring chains interact
451: via a quasi-gaussian soft-core potential, as introduced in earlier
452: sections. The radius of each blob decreases, and hence their number
453: increases (for a given overall length L) as the ratio $\rho/\rho^*$
454: increases, until the melt regime is reached where the blob size
455: reduces essentially to the segment length $b$, and the coarse-graining
456: strategy is no longer of any use. Over the whole semi-dilute regime a
457: polymer may thus be pictured as a necklace of blobs, and the present
458: coarse-graining strategy allows, in principle for an unequivocal
459: determination of intra and inter-molecular effective interactions
460: between these blobs. Work along these lines is in progress.
461:
462: \section{Acknowledgements}
463: The authors are grateful to their collaborators on this project over
464: the years; the work presented in this overview owes much to Peter
465: Bolhuis, Andrea Pelissetto, Vincent Krakoviack, Reimar Finken,
466: Evert-Jan Meijer and Benjamin Rotenberg. CIA acknowledges the support
467: of the EPSRC and AAL is grateful to the Royal Society of London for their support.
468:
469: \section{References}
470: \begin{thebibliography}{99}
471: \bibitem{flory} P.J. Flory and W.R. Krigbaum, {\it J.Chem.Phys.} {\bf 18}, 1086 (1950)
472: \bibitem{gros} A.V. Grosberg, P.G. Khalatur and A.R. Khokhlov, {\it Macromol.Chem.RapidCommun.} {\bf 3}, 709 (1982)
473: \bibitem{daut} J. Dautenhahn and C.K. Hall, {\it Macromolecules} {\bf 27}, 5399 (1994)
474: \bibitem{bolh} P.G. Bolhuis, A.A. Louis, J.-P. Hansen and E.J. Meijer {\it J.Chem.Phys.} {\bf 114}, 4296 (2001)
475: \bibitem{pelis} A. Pelissetto and J.-P. Hansen, {\it J.Chem.Phys.} {\bf 112}, 134904 (2005)
476: \bibitem{bolh2} P.G. Bolhuis, A.A. Louis and J.-P. Hansen, {\it Phys.Rev.E} {\bf 64}, 021801 (2001)
477: \bibitem{louis} A.A. Louis, P.G. Bolhuis, J.-P. Hansen and E.J. Meijer, {\it Phys.Rev.Letters} {\bf 85}, 2522 (2000)
478: \bibitem{hend} R.L. Henderson, {\it Phys.Lett.A} {\bf 49}, 197 (1974)
479: \bibitem{hansen} J.-P. Hansen and I.R. McDonald, ``Theory of Simple Liquids'' $2^{nd}$ ed, (Academic Press, London, 1986)
480: \bibitem{rubin} see {\it e.g.} M. Rubinstein and R.H. Colby, ``Polymer Physics'', (Oxford University Press, 2003)
481: \bibitem{bolh3} P.G. Bolhuis and A.A. Louis, {\it Macromolecules} {\bf 35}, 1860 (2002)
482: \bibitem{louis5} A.A. Louis, {\it J. Phys.Cond.Matt} {\bf 14} 9187 (2002)
483: \bibitem{stil} F.H. Stillinger, {\it J.Chem.Phys.} {\bf 65}, 3968 (1976)
484: \bibitem{lang} A. Lang, C.N. Likos, M Watzlawek and H. L\"owen, {\it J.Phys.Cond.Matt} {\bf 12}, 5087 (2000)
485: \bibitem{louis2} A.A. Louis, P.G. Bolhuis and J.-P. Hansen, {\it Phys.Rev.E} {\bf 62}, 7961 (2000)
486: \bibitem{finken} R. Finken, J.-P. Hansen and A.A. Louis, {\it J.Phys.A} {\bf 37}, 577 (2004)
487: \bibitem{grass} P. Grassberger and R. Hegger, {\it J. Chem. Phys.} {\bf 102}, 6881 (1995)
488: \bibitem{krak} V. Krakoviack, J.-P. Hansen, and A.A. Louis, {\it Phys.Rev.E} {\bf 67}, 041801 (2003)
489: \bibitem{ruel} D. Ruelle ``Statistical Mechanics: Rigorous Results'', (Benjamin, London, 1969)
490: \bibitem{schw} For a review see K.S. Schweizer and J.G. Curro, {\it Adv.Chem.Phys.} {\bf 97}, 1 (1997)
491: \bibitem{fuchs} M. Fuchs and K.S. Schweizer, {\it Phys.Rev.E} {\bf 64}, 021514 (2001)
492: \bibitem{krak2} V. Krakoviack, B. Rotenburg and J.-P. Hansen {\it J.Phys.Chem.B} {\bf 108}, 6697 (2004)
493: \bibitem{krak3} V. Krakoviack, J.-P. Hansen, and A.A. Louis, {\it Europhys.Lett.} {\bf 58}, 53 (2002)
494: \bibitem{dick} R. Dickman, {\it J. Chem. Phys.} {\bf 87}, 2246 (1987)
495: \bibitem{add} C.I. Addison, A.A. Louis and J.-P. Hansen, {\it J.Chem.Phys.} {\bf 121}, 612 (2004)
496: \bibitem{add2} C.I. Addison, J.-P. Hansen and A.A. Louis, {\it Chem.Phys.Chem.} (in press 2005)
497: \bibitem{barrat} J.-L. Barrat, T. Biben and J.-P. Hansen, {\it J.Chem.Phys.} {\bf 102}, 6881 (1995)
498: \bibitem{genne} P.G. de Gennes, {\it Phys.Lett} {\bf 38A}, 339 (1972)
499: \bibitem{freed} see {\it e.g.} K.F. Freed, ``Renormalization Group Theory of Macromolecules'', (Wiley, New York, 1987)
500: \bibitem{li} B. Li, N. Madras and A.D. Sokal, {\it J.Stat.Phys.} {\bf 80}, 661 (1995)
501: \bibitem{beloh} P. Belohorec and B. Nickel, Guelph University Report (1997) (unpublished)
502: \bibitem{pelis3} A. Pelissetto, {\it Private communication}
503: \bibitem{asaku} S. Asakura and F. Oosawa, {\it J.Chem.Phys.} {\bf 22}, 1255 (1954)
504: \bibitem{louis3} A.A. Louis, P.G. Bolhuis, E.J. Meijer and J.-P. Hansen, {\it J.Chem.Phys.} {\bf 117}, 1893 (2002)
505: \bibitem{louis4} A.A. Louis, P.G. Bolhuis, E.J. Meijer and J.-P. Hansen, {\it J.Chem.Phys.} {\bf 116}, 10547 (2002)
506: \bibitem{bolh4} P.G. Bolhuis, A.A. Louis and J.-P. Hansen, {\it Phys.Rev.Lett} {\bf 89}, 128302 (2002)
507: \bibitem{lekker} H.N.W. Lekkerkerker {\it et al.}, {\it Europhys.Lett} {\bf 20}, 559 (1992); E.J. Meijer and D. Frenkel, {\it J.Chem.Phys.} {\bf 100}, 6873 (1994)
508: \bibitem{rama} S. Ramakrishman, M. Fuchs, K.S. Schweizer and C.F. Zukoski, {\it J.Chem.Phys.} {\bf 116}, 2201 (2002)
509: \bibitem{bolh5} P.G. Bolhuis, E.J. Meijer and A.A. Louis, {\it Phys.Rev.Lett} {\bf 90}, 068304 (2003)
510: \bibitem{likos} For a review, see C.N. Likos, {\it Phys.Rep} {\bf 348}, 267 (2001)
511: \bibitem{add3} C.I. Addison, J.-P. Hansen, V. Krakoviack and A.A. Louis, {\it Mol.Phys.} (in press 2005)
512: \bibitem{widom} B. Widom and J.S. Rowlinson, {\it J.Chem.Phys.} {\bf 52}, 1670 (1970)
513: \bibitem{genne2} P.G. de Gennes, "Scaling Concepts in Polymer Physics" (Cornell University Press, Ithaca 1979)
514:
515: \end{thebibliography}
516:
517: \end{document}
518:
519: