• Keine Ergebnisse gefunden

Mode-Coupling Theory for Structural and Conformational Dynamics of Polymer Melts

N/A
N/A
Protected

Academic year: 2022

Aktie "Mode-Coupling Theory for Structural and Conformational Dynamics of Polymer Melts"

Copied!
4
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

VOLUME88, NUMBER18 P H Y S I C A L R E V I E W L E T T E R S 6 MAY2002

Mode-Coupling Theory for Structural and Conformational Dynamics of Polymer Melts

S.-H. Chong1,* and M. Fuchs2

1Physik-Department, Technische Universität München, 85747 Garching, Germany

2Institut Charles Sadron, 6 rue Boussingault, 67083 Strasbourg, France (Received 23 January 2002; published 22 April 2002)

A mode-coupling theory for dense polymeric systems is developed which unifyingly incorporates the segmental cage effect relevant for structural slowing down and polymer chain conformational degrees of freedom. An ideal glass transition of polymer melts is predicted which becomes molecular-weight independent for large molecules. The theory provides a microscopic justification for the use of the Rouse theory in polymer melts, and the results for Rouse-mode correlators and mean-squared displacements are in good agreement with computer simulation results.

DOI: 10.1103/PhysRevLett.88.185702 PACS numbers: 64.70.Pf, 61.20.Lc, 61.25.Hq

Polymeric materials often find applications as amor- phous solids or glasses. During their processing, the structural relaxation in polymeric melts plays an important role because it affects transport and viscoelastic proper- ties. An understanding of conformational dynamics is also required for single-chain diffusional processes. Whereas a number of phenomenological concepts have been in- vented, no microscopic theory exists which successfully explains the slow structural and conformational relaxations of polymeric melts close to the glass transition, as, for example, studied in detailed computer simulations of model systems [1].

We present a first-principles theory for structural and conformational dynamics of unentangled polymer chains using the site formalism which has already provided a good description of their static packing [2]. Building upon the mode-coupling theory (MCT) for simple and low- molecular systems [3– 5], our macromolecular exten- sion captures the segmental “cage effect” which causes a (asymptotically) molecular-weight-independent ideal glass transition driven by steric hindrance on microscopic length scales. A decoupling of local collective structural re- laxation from single-chain conformational fluctuations is obtained, leading to a first-principles derivation of the Rouse model [6]. This widely used model assumes a chain to be in a Markovian heat bath, and therefore its applicability requires a time-scale decoupling which our theory achieves. Entanglement effects [6] are neglected in our work because we do not ensure uncrossability of chains and consider only local isotropic forces [7].

A system of n chains, consisting of N identical monomers or segments, distributed with densityris con- sidered. In the site formalism [8], the structural variables are the monomer-density fluctuations for wave vector q,៬ rqa 苷Pn

i1exp共iq៬ ?ria兲, where ria denotes the position of the ath monomer in the ith chain. Structural dynam- ics shall be described by coherent density correlators, Fqabt兲 苷具rqatrbq典兾n. Here 具· · ·典 denotes canonical averaging for temperature T. A single (or tagged) polymer (labeled s) exhibits density fluctuations rq,sa

exp共iq៬ ? rsa兲, and correlators, Fq,sabt兲苷具rq,satrq,sb 典, characterize the single-chain dynamics.

We apply MCT equations in the site representation [5]

to flexible macromolecules, expecting them to capture in- termolecular caging in polymeric melts. The required in- puts are the static structure factorsSqabFqab共0兲,wqabFq,sab共0兲, and the direct correlation functionscqab[8]. There are severe difficulties to solve the (N 3 N)-matrix MCT equations for polymers because the degrees of polymer- izationNof interest are large. The simplifications adopted here are to neglect chain-end effects for cabq [2], cabqcq, and to consider the site-averaged correlator, Fq共t兲 苷 共1兾N兲PN

a,b1Fqabt兲, which deals with the total monomer- density fluctuations. This mean-field-like approximation replaces the site-specific surroundings of a segment by an averaged one. It is supported by the observation that SqFq共0兲captures the static correlations on the segmen- tal length scale [2].

We find a set of scalar equations for the normalized coherent correlatorfqt兲苷Fqt兲兾Sq:

2tfqt兲1 Vq2fqt兲1 V2qZ t

0

dt0mq共t 2t0兲≠t0fq共t0兲苷 0 , (1) mqt兲 苷 1

2 Z

dk V៬ 共៬q;k,p៬兲fkt兲fpt兲. (2) Here Vq2q2y2Sq and V 苷rmSqSkSp兵៬q?关kck 1

pcp兴其2兾共2p兲3q4 with y denoting the monomer thermal velocity,rmNr, andp៬ 苷 ៬q2k. These equations are៬ formally identical to the ones for simple systems [3]. On the other hand, one finds for the single-chain dynamics:

2tFq,st兲1 V2q,sFq,st兲 1 Vq,s2 Z t

0

dt0mq,st 2 t0兲≠t0Fq,st0兲 苷 0, (3) mabq,st兲苷 X

c

wqac q2

Z

dk Vsq;k,p៬兲Fk,scbt兲fpt兲. (4)

185702-1 0031-9007兾02兾88(18)兾185702(4)$20.00 © 2002 The American Physical Society 185702-1 First publ. in: Physical Review Letters 88 (2002), 18, pp. 185702

Konstanzer Online-Publikations-System (KOPS) URL: http://www.ub.uni-konstanz.de/kops/volltexte/2007/3781/

URN: http://nbn-resolving.de/urn:nbn:de:bsz:352-opus-37819

(2)

VOLUME88, NUMBER18 P H Y S I C A L R E V I E W L E T T E R S 6 MAY2002 Here V2q,sq2y2wq21 and Vs 苷rmq៬ ? p៬兾q2Spc2p

共2p兲3.

A traditional description of the single-chain dynamics is in terms of Rouse modes [6]. For discrete chains, Rouse-mode correlators are defined byCpp0t兲 苷具Xpt兲? Xp0典兾3N with Xp 苷p

2兾NPN

a1rsacos关共a 21兾2兲ppN兴. Sincerq,sa 艐 11iq៬ ? rsafor smallq,Cpp0t兲can be expressed as a linear combination of Fq,sabt兲 for q !0.

Here, it is neither assumed thatCpp0共t兲are diagonal nor as- sumed that the decay is exponential. Therefore, our micro- scopic theory can test the validity of the Rouse theory.

The polymers considered here shall be modeled as fol- lows. First, chains are assumed to be Gaussian, for which wqac 苷exp关2q2ja2cjb2兾6兴with the statistical segment length b. Second, each monomer is modeled as a hard sphere of diameter d and mass m, and we set bd.

All equilibrium properties are then specified by the pack- ing fractionw 苷prmd3兾6and the degree of polymeriza- tion N. Third, Sq andcq are evaluated from the polymer reference-interaction-site-model theory [2]. Let us add that the value of w can become larger than 1 for very dense systems since corrections due to nonphysical intrapoly- mer monomer overlap [9] are not taken into account in this Letter. From here on, the units will be chosen so that d 苷y 苷m 苷 1.

The derived MCT equations (1) and (2) exhibit an ideal liquid-glass transition upon increasingwat a critical value wcas discussed for simple systems [4,10]. The solid line in Fig. 1 exhibits the resulting glass-form factors fqc 苷 fqt !`兲atwcforN 苷10, and the dashed line exhibits Sq multiplied by 0.15. The fqc also measure the ampli- tude of the slow structural relaxation in the liquid state.

Inspection of Eq. (2) shows that the liquid-glass transition is driven mainly by the changes ofSqforqmax 艐7, where fqc is maximal, i.e., by local fluctuations connected to the

0 5 10 15 20 q

0.2 0.0 0.4 0.6 0.8 1.0

1 2 3 4 5 log10 N 0.6

0.8

N=10 1.0

Sq

fqc

qmax

ϕ (Ν)/ϕ c c inf

FIG. 1. The solid line exhibits the glass-form factors fqc at w 苷wc forN 苷10 along with the simulation result (circles) for a similar system taken from Ref. [14]. The dashed line denotesSqmultiplied by 0.15. The inset showswc共N兲rescaled bywcinf苷wc共N !`兲 艐1.309as a function of log10N.

average nearest-neighbor monomer distance (the cage ef- fect). With varying q, fqc oscillates in phase with Sq as found in simple systems [4,10]. The circles denote the result of molecular-dynamics simulation performed for a similar system [11 – 14]. The agreement is semiquantita- tive especially forqqmax, which is the relevantqrange for the ideal glass transition [15]. The inset demonstrates thatwcN兲becomes independent ofNfor largeN. This is because the glassy arrest is driven by the local cage effect, and the global chain size plays a gradually smaller role for larger N [16].

For packing fractionswclose to but belowwc, correla- torfAt兲of any variableAcoupling to density fluctuations exhibits slow structural relaxation and decays in two steps:

the decay towards the plateaufAc, followed by the decay to zero (so-calledadecay) [3,4]. Detailed analysis offqt兲 and Fq,sabt兲will be presented in later publications. Here, only the Rouse-mode correlatorsCpp0t兲 shall be studied.

Typically, their off-diagonal elements are found to be much smaller (at most only a few percent) compared to the di- agonal ones at all times. Therefore, the normalized diago- nal elements,cpt兲苷 Cppt兲兾Cpp共0兲, shall be considered.

Figure 2 exhibits representative results forN 苷 10for a reduced packing fraction共w 2 wc兲兾wc 苷21022.

The curves shown in Fig. 2 do not clearly exhibit the two-step relaxation scenario. This is because the plateaus of cpt兲 are so large, fpc .0.9, that only less than 10% of the decay is left for the relaxation towards the plateau. Thus, most of the relaxation of cp共t兲 occurs in thea regime. Let us characterize the a time scale by cp共tp兲 苷fpc兾20. The correspondinga time scale tqmax shall be introduced for the coherent correlator fq共t兲

-2 0 2 4 6 1.0

0.2 0.0 0.4 0.6 0.8

1 2 3 4 5 6 7 8 9

p=9 5 1

log t

10

c (t)p

N=10

βp

τ /τp qmax p 10

100

1

1.0 0.96 0.92

f1c f9cf5c

FIG. 2. Rouse-mode correlators cp共t兲 of p苷1 (solid line), p 苷5 (dashed line), andp苷9(dotted line) for N 苷10 and 共w 2 wc兲兾wc 苷21022. The horizontal lines mark the plateaus fpc. The inset shows the ratiotp兾tqmaxof thea-relaxation times (circles) and the stretching exponentbp (squares) as a function of p. The dotted line denotes a fitting based on the formula tp 苷z兾12sin2pp兾2N兲.

185702-2 185702-2

(3)

VOLUME88, NUMBER18 P H Y S I C A L R E V I E W L E T T E R S 6 MAY2002 for qqmax, characterizing the local dynamics of the

surrounding medium. The ratiotp兾tqmax is shown in the inset as a function of p. For small p, the scales tp are much larger than the local scaletqmax.

The Rouse theory assumes that all dynamical correla- tions in the surroundings are much faster than the single- chain dynamics [6]. Since a polymer is surrounded by identical polymers, the assumption of the time-scale sepa- ration cannot be justifieda priori. Our microscopic theory verifies this central assumption. The Rouse theory pre- dicts, within our units,tp 苷 z兾12sin2pp兾2N兲, wherez denotes the monomer friction. This formula, usingz as a fit parameter, is shown as the dotted line, and we found z兾tqmax 苷16.5.

The shape ofcpt兲in thearegime is often characterized by the stretching exponent bp of the Kohlrausch-law fit:

cpt兲 ~exp关2共t兾tp0bp兴. We foundbp . 0.9for allp, as shown in the inset. For smallp,bp is close to 1 due to the large separation of the scales tp andtqmax, as discussed above. bp decreases as p increases since for large p the scalestp become comparable totqmax. However,bp remains close to 1 even for large p. This is because the plateaus fpc are high, as explained in Ref. [17]. Thus, all ourcpt兲 exhibit nearly Debye relaxation as assumed by the Rouse theory. Let us note that the found features for cpt兲 hold also for N 苷100, and are in agreement with simulation results [12].

Of interest are the monomer mean-squared displacement (MSD)gat兲苷 具关៬rsat兲 2rsa共0兲兴2典and the center-of-mass MSD, gCt兲. Let us also introduce the monomer- averaged MSD, gMt兲苷 共1兾N兲PN

a1gat兲. Figure 3 ex- hibitsgMt兲andgCt兲forN 苷10and 100 for共w 2 wc兲兾 wc 苷21022. For short times, the MSD exhibits ballistic motion:gXt兲 ~t2(X 苷MorC). As the time increases, the MSD begins to be suppressed due to the cage effect and there appears the so-called b regime, where gX共t兲 is close to the square of the critical localization length, gX共t兲 艐6共rcX2. For this regime, there holds [3,4,18]

gXt兲 苷6共rXc2 26hXGt兲, XM,C, (5) where G共t兲 denotes the b correlator and hX denotes the critical amplitude. Thus, gM共t兲 and gC共t兲 cross their plateaus6共rXc2 at the same time. The valuerMc 艐0.077, quantifying the monomer localization, for both N 苷 10 and 100 is consistent with Lindemann’s melting criterion.

rCc is reduced by about1兾p

Ncompared torMc, as expected for independent motions of constituent monomers. The asymptotic law (5) for each MSD is drawn as a dashed line, and its range of validity is indicated by diamonds.

The trends in gXt兲 up to theb regime are qualitatively similar to the ones for a sphere in a simple system [18].

The increase ofgXt兲above the plateau6共rXc2towards the diffusion asymptote,gXt兲 艐6Dtwith the diffusivity D, is theaprocess of the MSD. In contrast togCt兲,gMt兲 in this regime is significantly affected by chain connectiv- ity since the segments participate in the conformational

-2 0 2 4 6 8 10 -2 0 2 4 6 8

log 10t

t0.5

~ t0.6

~

N=100 N=10

log g(t) 10log g(t) 10

4 2

6(r )Cc 2 6(r Mc 2) R2g

R2

6(r )Cc 2 6(r Mc 2) R2g

R2

-6 -4 -2 0 2 -4 -2 0

-2 0 2 4 6 8 10 1.0

1.5 2.0

-2 0 2 4 6 8

1.0 1.5 2.0

6Dt 6Dt

g (t)

M

g (t)

C

g (t)

M

g (t)

C

g (t)/g (t)1

50 g (t)/g 1 (t)

5

log t 10

log t 10

FIG. 3. Double logarithmic presentation of the MSDs gM共t兲 andgC共t兲(solid lines) forN 苷10(upper panel) and 100 (lower panel) for共w 2 wc兲兾wc 苷21022. The dashed lines denote the asymptotic formula given in Eq. (5). The open diamonds mark the points where the dashed lines differ from the solid ones by 10%. The dash-dotted lines exhibit the diffusion law,6Dt. The dotted lines show the power laws,⬃tx, withxspecified in each panel. The horizontal dotted lines from bottom to top succes- sively mark the plateau6共rCc2forgC共t兲, the plateau6共rMc2 for gM共t兲, the radius of gyration Rg2, and the end-to-end distance R2. The insets exhibit the ratiog1t兲兾gN兾2t兲.

motion and most of the relaxation ofcp共t兲occurs here, as explained above. As a result, there appears a subdiffusive (⬃tx) regime in gMt兲. For N 苷10, the exponent x is 0.60, which is close to the value (0.63) found in the men- tioned simulations [12]. For N 苷100, we findx 苷0.5, as predicted by the Rouse model and by the asymptotic evaluation of our theory. The different value forN 苷10 can be attributed to finite N effects. Thus, we find a strong polymer-specific effect for the beginning of thea process, while no such effect is reflected ingMt兲up to the b process.

The insets of Fig. 3 show the ratio of the MSD for end and central monomers,g1t兲兾gN2t兲. The ratio starts from

185702-3 185702-3

(4)

VOLUME88, NUMBER18 P H Y S I C A L R E V I E W L E T T E R S 6 MAY2002 1 in the ballistic regime, exhibits a maximum for interme-

diate times, and tends to 1 in the diffusion regime. It is seen that the ratio remains close to 1 also within theb regime. This is because the dynamics here does not reflect the chain connectivity, as explained above. The Rouse the- ory predicts the maximum to be 2 within the time regime where the monomer MSD exhibits the t12 law, i.e., in the a regime. The result for N 苷100 clearly indicates this behavior, while the maximum forN 苷 10is slightly smaller (1.8) due to finiteNeffects. The shape of the ratio for N 苷 10 in theb and a regimes is in semiquantita- tive agreement with the mentioned simulation results [19].

Thus, the chain-end effect is properly taken into account in our theory even though neglecting that effect for cabq . This is because the matrix structure of Eqs. (3) and (4) is preserved for describing the single-polymer dynamics.

In summary, a first-principles theory for structural slow- ing down of dense polymer systems has been presented which also provides a microscopic derivation of the Rouse model for unentangled chain melts. Chain connectivity is seen to cause polymer-specific long-time anomalies of the aprocess. We use the concept of an ideal MCT glass tran- sition, familiar for simple liquids and colloidal suspensions [20], and our results agree well with simulation studies of simple (coarse-grained) polymer models. Our theory thus provides insights into dynamical aspects typical for poly- mer melts which are also observed in models [21,22] with a more realistic local chemistry.

We thank J. Baschnagel, M. Aichele, and W. Götze for discussions. M. F. was supported by the Deutsche Forschungsgemeinschaft, Grant No. Fu 309/3.

*Present address: Dipartimento di Fisica, INFM and INFM Center for Statistical Mechanics and Complexity, Uni- versità di Roma “La Sapienza,” Piazzale Aldo Moro 2, I-00185, Roma, Italy.

[1] K. Binder, J. Baschnagel, and W. Paul, Prog. Polym. Sci.

(to be published).

[2] K. S. Schweizer and J. G. Curro, Adv. Chem. Phys.98,1 (1997).

[3] W. Götze, inLiquids, Freezing and Glass Transition,edited by J.-P. Hansen, D. Levesque, and J. Zinn-Justin (North- Holland, Amsterdam, 1991), p. 287.

[4] W. Götze and L. Sjögren, Rep. Prog. Phys.55,241 (1992).

[5] S.-H. Chong and W. Götze, Phys. Rev. E 65, 041503 (2002).

[6] M. Doi and S. F. Edwards,The Theory of Polymer Dynam- ics(Oxford University, Oxford, 1986).

[7] K. S. Schweizer, J. Chem. Phys.91,5802 (1989);91,5822 (1989).

[8] D. Chandler and H. C. Andersen, J. Chem. Phys.57,1930 (1972).

[9] K. S. Schweizer and J. G. Curro, Macromolecules21,3070 (1988).

[10] T. Franosch, M. Fuchs, W. Götze, M. R. Mayr, and A. P. Singh, Phys. Rev. E55,7153 (1997).

[11] C. Bennemann, W. Paul, K. Binder, and B. Dünweg, Phys.

Rev. E 57,843 (1998).

[12] C. Bennemann, J. Baschnagel, W. Paul, and K. Binder, Comput. Theor. Polym. Sci.9,217 (1999).

[13] C. Bennemann, J. Baschnagel, and W. Paul, Eur. Phys.

J. B10,323 (1999).

[14] M. Aichele and J. Baschnagel, Eur. Phys. J E5,229 (2001);

5,245 (2001).

[15] The disagreement forq,qmaxmay be due to the fact that center-of-mass density fluctuations are not our explicit dy- namical variables [14], while the one forq.qmaxmay be due to a softer Lennard-Jones potential which is employed in the simulation.

[16] G. B. McKenna, inComprehensive Polymer Science,edited by C. Booth and C. Price (Pergamon, New York, 1989), Vol. 2,p. 311.

[17] W. Götze, A. P. Singh, and Th. Voightmann, Phys. Rev. E 61, 6934 (2000).

[18] M. Fuchs, W. Götze, and M. R. Mayr, Phys. Rev. E 58, 3384 (1998).

[19] J. Baschnagel (private communication).

[20] W. Götze, J. Phys. Condens. Matter11,A1 (1999).

[21] A. van Zon and S. W. de Leeuw, Phys. Rev. E58,R4100 (1998).

[22] A. van Zon and S. W. de Leeuw, Phys. Rev. E 60, 6942 (1999).

185702-4 185702-4

Referenzen

ÄHNLICHE DOKUMENTE

An example of a derivational path within the polyse- mous structure of the verb drop is the following: the primary sense 'let some- thing fall' is derived directly from

The SCN − counterions adopt a monodentate coordination mode: N atoms are coordinated, and S atoms are free of coordination.. Cu(II) ions

Moving on to estimating the number of internodes for regular, lateral long shoots, the independent variables that were identified as influential are the length of the mother GU

The stationary distribution (2.15) or (2.21) will turn out to be a useful tool in the analysis of the migration process. Furthermore, any time dependent solution of this

where (iJ restrains the introduction of the new technique and states that transition from -r to y does not increase the labor force in the establishments, (ii) states that

(Color online) Normalized critical nonergodicity parame- ters of the simulated collective density correlators ss k (t) of the small particles, extracted from Kohlrausch fits for

(2008), which do suggest macroscopic flow inhomogeneity under steady shearing in dense colloidal suspensions. The rest of this paper is organized as follows. Section II

We analyze the slow glassy structural relaxation as measured through collective and tagged-particle density correlation functions obtained from Brownian dynamics