• Aucun résultat trouvé

Theory of the dynamics of tagged chains in interacting polymer liquids: general theory

N/A
N/A
Protected

Academic year: 2021

Partager "Theory of the dynamics of tagged chains in interacting polymer liquids: general theory"

Copied!
16
0
0

Texte intégral

(1)

HAL Id: jpa-00247001

https://hal.archives-ouvertes.fr/jpa-00247001

Submitted on 1 Jan 1994

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Theory of the dynamics of tagged chains in interacting polymer liquids: general theory

T. Vilgis, U. Genz

To cite this version:

T. Vilgis, U. Genz. Theory of the dynamics of tagged chains in interacting polymer liquids: general theory. Journal de Physique I, EDP Sciences, 1994, 4 (10), pp.1411-1425. �10.1051/jp1:1994196�.

�jpa-00247001�

(2)

Classification Physics Abstracts

05.20D 05.40 61.25H

Theory of the dynamics of tagged chains in interacting polymer liquids: general theory

T.A- Vilgis and U. Genz

Max-Planck-Institut für Polymerforschung, Postfach 3148, D-55021 Mainz, Germany (Received 24 February 1994, revised May 1994, accepted 30 June 1994)

Abstract. The effective Rouse dynamics of a surgie polymer chain embedded in

an interact- ing polymer liquid, such as excluded volume melts, blends, or copolymer melts is studied. This

paper suggests an alternative formulation for the problem of the dynamics ofinteracting polymer liquids. This theory uses trie Martin-Siggia-Rose functional formahsm, instead of Mori-Zwanzig

projection operators. The theory is developed in different steps. The first approximation is an effective single chain approximation, that leads to a renormalized Rouse model. Taking into account ail interactions from other chains and performmg a preaveraging over the static and

dynamic structure factors leads to an effective friction function acting on the monomers of the tagged chain. The friction function is determined by ail interactions present in the system.

1. Introduction and motivation.

Static and dynamic properties of individual polymer chains in polymer melts, blends and block copolymers is in general an unsolved problem. The static problem seems to be quite simple

and has been studied previously for vanous situations: melts [1], polymer blends [2] and block

copolymers [3, 7-9]. The one component melt is relatively simple since the conformation of individual chains is ruled by excluded volume interactions and the chain connectivity only. In the case of polymer blends, the situation is more complicated. The underlying phase transi- tion of trie density fluctuations has a dramatic effect on the chain conformation of individual chains. Individual chains in a blend of two homopolyrners show a tendency to shrink when

approaching the phase transition and their radius of gyration reduces up to 15Sl. This has been demonstrated by numerical simulations [10]. In block copolymers, the situation is even

more comphcated. At the rnicrophase separation, the mean field collective structure factor derived first by Leibler [4] diverges at a finite wave vector k*, and not in the small k limit as in the case of a polymer blend. When fluctuations are taken into account it has been shown first

on a general basis by Brazovskii [5] that the divergence is removed and the two point density density correlation function remains finite. This theory has been apphed to block copolymers

[fil where it was shown exphcitly that the structure factor for block copolymer melts does not

(3)

diverge. The physical reason for this fundamental change between mean field theory and renormalized theory is that fluctuations drive the second order phase transition into a weakly

first order transition. Thus correlation lengths, structure factors, etc. remain finite.

The net effect on the static properties of a single chain is that individual blocks in the block

copolymer melt shrink, whereas, because of the strong repulsion between the two blocks, the different blocks are widely separated. This has until present times been found only in numerical

simulations [8] for high density (incompressible system) and for systems including vacancies [8, 9], 1e. compressible systems. In both cases the chain stretching effect has been seen. The

experimental situation is not clear at present. Whether this effect can be seen clearly in real

experiments remains to be demonstrated.

The calculation technique for the static single chain problem is very simple: on the basis of the Edwards Hamiltonian, the partition function for ail chains can be formulated. When

integrating over ail chains except chain 1, an effective single chain Hamiltonian is obtained.

It consists of two parts: the first contribution is due to the chain connectivity which is rep- resented by a Wiener term in the continuous description of the chain. The second term is an effective potent1al consisting of the excluded volume potential of chain 1 itself and the effect due to fluctuations and interactions with the surrounding medium. This effective Hamiltoman

depends on the positions R(s) of the segment s on chain 1 and can be expressed as

N ~ N N

H((R(s))) = ~~)/Î / ds

(~(~~~) + kBT /

ds / ds' U~R(R(s) Ris')) il-1)

2 s

N is the chain length and l~ is the mean squared distance between segments. The influence of the environment on the tagged chain is included in the effective potential UeR(R). Most remarkable is that in the case of polymer mixtures the spatial Fourier transform U~R(k) of this effective potential contains a singular part that changes sign before the phase separation occurs

[2]. This means that the effective potential acting on the chain becomes attractive before phase

separation, hence the radius of gyration of the chains shrinks. Such effects have been found numerically by Sanban and Binder [loi and very recently by light scattenng expenments il Ii.

It has also been shown that the effective potential calculated by reference [2] is responsible for

a weak locaJization of the chains when the size of the cntical droplets is of the order of the radius of gyration [12].

Motivated by the simple static case similar considerations should be possible for the dynamic behavior of tagged chains m interacting polymeric liquids. Unfortunately the dynamic problem

is more comphcated. Let us first summarize a few expenmental results. It is well known from neutron scattenng expenments [13] that in a dense polymer melt, interacting via the excluded volume, (below the cntical molecular weight from where reptation starts) the dynamics of an

individual chain can be described by the Rouse equation,

Ifôt~ ~~(~Î ôS~~) Ris, t) = fis, t) , (1.2)

where ( is the segmental friction coefficient and fis, t) is a random force with the correlation function

(f(s, t) fis', t'))

= 6kBT à(s s') à(t t') (1.3)

Equation (1.3) is the fluctuation dissipation theorem.

On the other hand, equation (1.2) is vahd exactly only for the dynamics of single, non- interacting random walk chains which are self-similar on ail scales. For this ideal system, the

(4)

Hamiltonian [14] is given by

N ~

~ÎÎÎ/~~ ~ÎÎ~~

'

~~'~~

o

where N is the total counter length of the random walk. In the absence of interactions, the friction coefficient ( is the bare friction constant (o which is determined by the heat bath

govemed by the random force fis, t).

In the presence of interactions, it has been shown very often that the friction coefficient (o

is renormalized by the effect of interactions. The simplest model shows an additional term

[15~17]

t " to + j Idi (F(il F1°11 11.51

o

where Fit) is the force on a segment at time t due to excluded volume interactions. The above equation is not at ail as trivial as it looks. The first term represents the bare friction coefficient (o which stems from the heat bath. In Langevin dynamics this contribution can be thought to

have its physical reasons in local random forces. The second contribution, 1-e- the explicitly

written out force force correlation function is due to forces which cannot be assigned to the local fluctuations. Equation (1.5) can be intuitively understood in a coarse grained model.

The first term is due to stochastic processes within a lower cut off scale whereas the second term descnbes the stochastic forces at scales larger than the cut off. In polymer dynamics

such a widely used coarse grained model is the Rouse model [14] and the lower cut off length corresponds roughly to the (static) Kuhn segment length. This point will be discussed later in

more detail.

The aim of this paper is to find a field theoretical way to study the dynamics of interacting polymeric hquids. The dynarnics of a tagged chain in such a multicomponent interacting hquid

can be studied by various methods. The classical one is the Mori Zwanzig projection operator technique. This method was developed by Mon and Zwanzig [18]. Two types of variables are

distinguished: slow and fast variables. Trie fast variables are then projected out and bave an effect on transport quantities such as trie friction coefficient. This very successful standard technique provides a good basis for many problems in trie physics of liquids. Trie result is a

generalized Langevin equation (GLE). This method bas been used to study one component polymeric liquids, where trie chains are interacting by excluded volume forces by Schweizer [19- 22]. In these basic papers it was shown that excluded volume forces lead to several additional effects, including a possible explanation for the Rouse to reptation transition, that is well known

to occur for long chains [14]. In section 2, this method will be outlined and generalized to a

system containing several species, where in contrast to Schweizer we aim for the investigation of trie dynamics of tagged chains near (critical) fluctuations of a phase transition. Trie latter

generahzation is necessary if trie dynamical behavior near a theromdynamic phase transition

is studied.

One intuitively diflicult problem when applying Mon Zwanzig projection operator techniques

to treat a polymer melt is the following: a clear separation between time scales is not possible.

Ail time cales conceming the polymer molecules are essent1al since they contribute to the transport coefficients such as diffusion constants or viscosity. This does not at ail mean that the results derived by the Mon Zwanzig method are unphysical. In fact, the opposite is the

case as the Mon Zwanzig formalism denves a number of formally exact results, which are, however, sometimes diflicult to banale exphcitly without further assumptions. In section 3

(5)

therefore, another alternative method will be introduced which does not use this somewhat artificial separation of time scales by using a functional method. An exact dynamical generating

function is constructed such that correlation and response functions can be calculated simply by differentiation. This functional method based on the dynamical theory of Martin, Siggia

and Rose (MSR) [23] provides a framework which is analogous to trie statistical mechanics

developed for the static case [1-3, 7, 8]. Ànalogously to the Mori Zwanzig method the exact functionals cannot be used explicitly and several approximations have to be employed when

specific systems are discussed. One purpose of this paper is to present a dynamic functional for the behavior of the tagged chain analogously to the static case, where the effective potential

contains the information of the environment. However, for the simplest case, we expect a dynamic functional with a friction function that has most of the information of the environment,

i-e- the other chains in the system and their thermodynamic behavior. Another purpose of the present study is to provide an alternative formalism that uses functionals, for the study of the dynamics of tagged chains in interacting polymer fluids, which will be applied to more specific problems in subsequent investigations. Detailed applications of trie general result obtained in this paper to polymer blends and block copolymer melts are presented separateiy [25]. It

shouid be mentioned that the MSR technique has been appiied to the probiem of collective

dynamics of poiymers by Heifand et ai. [24]. These authors are interested in the dynamics of collective properties of concentrated poiymer solutions, whereas the present study is concemed

by a functional formulation of the tagged chain dynamics in interacting (criticai) polymer systems.

2. Discrete polymer mortels and projection operator techniques.

In a recent serres of publications [19-22], Schweizer addressed the problem of describing polymer dynamics in terms of microscopic chain properties. This avoids an a priori assumption on a

chain reptating in a tube. We present a generaiisation to a system containing severai types of poiymer species. Starting from the Liouvilie equation, Schweizer [19] empioyed Mon.Zwanzig

projection operator techniques [18] to derive an exact equation of motion for trie phase-space

distribution function fi((Ris), (Pis)) of a particuiar 'probe' chain under consideration.

This distribution function depends on the positions (Ris and the momenta (Pis) of ail Ni segments on this chain, which are indicated by trie indices s

= 1,.

,

Ni Trie mathematicai

procedure is niceiy presented in trie appendix of Ref. [19]. As a resuit of projection operator formalism, the time evoiution of fi is expressed in terms of a 'frequency matrix', which is local in time, and a memory contribution due to trie simuitaneous time deveiopment of the

surrounding 'matrix' polymers. The latter ieads to non-Markovian behaviour. In addition, a 'random force' term is present. From the equation for the time evolution of the distribution

fi ((Ris Ii) ), (Pis Ii) )), coupled equations for trie individual chain segments bave been denved.

~vithin a coarse-grained description for trie polymer chains, rapidly decaying contributions to the memory function are taken into account by a friction coefficient ii of a segment on the probe chain. In addition to the probe chain, the system contains several species a having va

polymers. Correlations of intermolecular forces F~, which act between segment s on the probe

chain and the matrix, decay on the time-scale of density fluctuations and have to be taken into account explicitly. For a Gaussian chain, the result can be expressed as

j~/

~~'~~i~'(~'~~) j'~l~~~'~ ~ ~l (2iÉ18 iÉl,S+1 iÉl,S-1) Ù~(~) (~'~)

~ 0

(6)

where ci

= 3kBT/1) is entropic spring constant in this bead spring model. The generalized

friction function (~,~> (t,t') is obtained as

i~,~,ji, i') = iiôjs si) à(i i')

+ £ (Fis expjoi£jt t')j Fis>

=

Îjs s') ôjt t') + Aj~,~>il, t') j2.2)

Fis gives the force on segment s resulting from intermolecular interactions with matrix poly-

mers. £ is the Liouville operator, and the projector Qi is orthogonal to the projection on the

probe polymer distribution function fi fi represents the random force and (...) refers to the

equilibrium average. The essential difference between equation (2,1) and the Rouse equation il.2) is the time dependence of the generalized friction function. If the second term in equation (2.2) vanishes, equation (2.1) reduces to the Rouse equation (1.2). The essent1al problem is

the evaluation of the generalized friction function, and a major difliculty is the treatment of the projected time evolution. Schweizer addressed this problem in references [19, 20]. Here,

we take a more conventional point of view and factorize trie four-point correlation function in equation (2.2) into a product of twc-point correlation functions while replacing trie pro-

jected time evolution operator by trie fuit Liouvilhan. This time of approximation is frequently employed when using a mode-coupling approximation [18]. It bas also been considered by

Schweizer [20, 22].

In a multicomponent melt, the force acting on polymer 1 due to the presence of the sur-

rounding matrix may be expressed in a Fourier presentation as v~

Fis Ii)

=

~ ~j ~j ~jV/~ exp [ik (Ris(t) Ris<Ii))] +

c-c-

,

(2.3)

2

, ~

where c-c- denotes the complex conjugate. In the case of excluded volume interaction, the

constants V?a are excluded volume parameters specifying the interaction between a segment

on chain and a segment of species a. The use of such simple (pseudo) potent1als is a simpli- fication. For microscopic descriptions, realistic potent1als, such as Lennard Jones potentials or

other models, should be employed. For the purpose of the present paper the pseudo potential approximation is suflicient. When inserting equation (2.3) into the memory function, the ad- ditional contribution to the friction function due to interactions with the polymer matrix can be written as

A(ss, ii,t')

=

~j ~j V/~V(~ (exp[ik RiPi exp[Qi £t] exp [iq Ris, P(

,

(2.4)

aT kq

~i>here pi is the Fourier transform of the partial density related to species a,

§ fiÎ~

" ~ ~ ~~P Î~~ iÉÎs(1)1 (~.5)

1 S

In equation (2.5), the summation is performed over ail segments belonging to species a. R$ ii)

gives trie position of segment s on chain at time t. Trie sum of four-point correlation functions in equation (2A) is factonzed in a product of two point correlation functions containing either

the matrix coordinates or the coordinates of the tagged chain. Simultaneously, the projected

(7)

time evolution operator is replaced by the fui] Liouville operator. This leads to

Î~Îss'Il>Î~) " ~ ~ Î~Î~Î~~~ (~XP Î~~ iÉls(~)Î ~XP Î~~' iÉls'(~')Î) (PÎ(Î)p~(Î~)) (2.6)

aT kq

Because the polymer system is isotropic, only terms q = -k contribute to the summation in

equation (2.6), which then simplifies to

Aiss>Ii,t')

=

~ fl V/~V/~ (exPlik lRis(t) Ris> (t'llll (P[(t)P[(t')1 (2.71

It can be seen that the dynamics of the tagged partiale is influenced by the dynamical properties of ail species. The above equation is closely related to the result given by Schweizer [19, 20], 1-e- the renormalized Rouse mortel. Trie difference to Schweizer's result is that here trie projected

propagators are replaced by trie unprojected ones. This possibility has already pointed out in reference [19].

À mode couphng approximation forms an alternative way to evaluate the additional, time-

dependent friction. It leads to more specific predictions on the dependence of the friction function on the segment indices s and s'. Whereas this additional step is very useful for the

explicit evaluation of the memory terms, it does not help to clarify the relation between trie

Martin-Siggia-Rose approach presented in trie following section.

3. Martin-Siggia-Rose formalism.

The aim of this section is the denvation of a generalized Langevin equation describing one tagged chain interacting with other chains. This linear Langevin equation will contain the influence of polymers interacting with polymer 1 via equilibrium properties of the surrounding polymers and will have the form of equation (2.1). So the procedure forms an alternative to trie Mon-Zwanzig projection operator formahsm. We start from u coupled, nonhnear, time local Langevin equations for u polymers including polymer 1. In a continuum description of trie chain, these equations describe trie time evolution of trie vectors R~(s,t) specifying trie

position of segment s on chain at time t,

l~ ~2

~~ ai ~~ ôs~ ~~~~'~~ ~~~~'~~ ~ ~~~'~~ ~ ~'' '~ ~~'~~

(~ is trie microscopic friction constant of a segment on chain1. In the overdamped case consid- ered here, the frictional force, which is trie first term on trie left nana side of equation (3.1), is

matched by ail other forces acting on segment s of chain 1. Three types of forces are present:

Very fast variables, which are not taken into account explicitly in this description, grue use

to random forces ((s,t), which are related to the microscopic friction coefficients (~ by the fluctuation dissipation theorem,

(i~(s, ii i~(s', il))

= 6k~T i~ô~~ à(s s'j à(i il)

,

(3.2)

where ô~j is the Kronecker symbol and à(z) is the Dirac à-function. The second term on the left hand side of equation (3.1) represents the entropic spring forces between subsequent

segments on chain 1. Àdditional excluded volume interactions between segments on the same chain and on different chains j #1 are included in F~(s, t). If this term vanishes, equation (3.1) reduces to the Rouse equation, equation (1.2). Because F~(s,t) depends on ail distances

(R~(s, t) Rj (s',t)) between segment s on chain to other segments s' on polymers j = 1,. u,

Références

Documents relatifs

It has been found that, when conditions are favourable, farmers modify their practices to respond to opportunities generated by good market trends (Cour, 2000). The support of

After finding that the static properties of dense polymers very well compare to ideal chains, it is especially interesting to check the chain dynamics. Since the chain

We also find in agreement with de Gennes that the critical exponents are well described by their mean field values and that the size of the critical region

By Monte Carlo simulation of two lattice models of polymer melts, the enrichment of the chain ends in the vicinity of a wall is observed.. For the bond fluctuation model

by applying a stress to the iuelt. This explains why the initial nuniber of stress points must be considered as constant for a given polyiuer even in the limit of infinitesimal

The intermediate coherent scattering function of entangled polymer melts: a Monte Carlo test of des Cloizeaux’ theory... Classification

In order to obtain a more quantitative insight in trie temperature dependence of trie accep- tance rate Figure 2 presents an Arrhenius plot of trie bulk value Abuik Ii. e., trie

One of the key parameters in viscous fluids is the diffusion coefficient, denoted by D. It affects a lot of properties in glass-forming liquids like relaxation times and