• Aucun résultat trouvé

Random hydrophilic-hydrophobic copolymers

N/A
N/A
Protected

Academic year: 2021

Partager "Random hydrophilic-hydrophobic copolymers"

Copied!
11
0
0

Texte intégral

(1)

HAL Id: jpa-00248120

https://hal.archives-ouvertes.fr/jpa-00248120

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.

T. Garel, L. Leibler, Henri Orland

To cite this version:

T. Garel, L. Leibler, Henri Orland. Random hydrophilic-hydrophobic copolymers. Journal de

Physique II, EDP Sciences, 1994, 4 (12), pp.2139-2148. �10.1051/jp2:1994252�. �jpa-00248120�

(2)

J. Phys. II Fiance 4

(1994)

2139-2148 DECEMBER1994, PAGE 2139

Classification Physics Abstracts

81.205 64.75 75.10N

Random hydrophilic-hydrophobic copolymers

T. Garel

(~),

L. Leibler

(2)

and H. 0rland (~,*)

(~) CEA, Service de Physique Th40rique

(**),

CE-Saclay, 91191 Gif-sur-Yvette Cedex, France (2 Groupe de Physico-Chimie Th40rique (***), E.S.P.C.1., 10 rue Vauquelin, 75231 Paris Cedex

05, France

(Received

15 June 1994, accepted in final form 14 September 1994)

R4sumk Nous 4tudions

un copolymkre al6atoire amphiphile AB dans

un solvant s61ectif

(par

exemple, de

l'eau).

Nous consid4rons deux cas. Dans le cas du d4sordre mobile, les monom4res

hydrophiles

(A)

et hydrophobes

(B)

sont h l'4quilibre chirnique local, et la fraction de monom+res A ainsi que leur position dans l'espace peuvent varier, alors quo dans le cas du d6sordre geld

(qui

est rel16 au probmme des

prot6ines),

la s4quence chimique est fix4e par synt1Jbse. Dans les deux cas, le comportement de la chine depend de son hydrophobicit4 moyenne. Pour une chine

fortement hydrophobe

(grande

fraction de

B),

on trouve un point d'elfondrement 6 continu

ordinaire, avec une grande entropie conformationnelle. Pour une chaine faiblement hydrophobe

ou hydrophile, on trouve une transition inhabituelle du premier ordre. En particulier, dans le

cas du d4sordre gaussien, cette transition discontinue est pilot6e par un changement de signe

du troisibme coefficient du viriel. L'entropie de cette phase collaps4e est fortement r4duite par rapport I cello d'un point 6 ordinaire.

Abstract. We study a single statistical amphiphilic copolymer chain AB in a selective

solvent

(e.g. water).

Two situations

are considered. In the annealed case, hydrophilic

(A)

and hydrophobic

(B)

monomers are at local chemical equilibrium and both the fraction of A

monomers and their location along the chain can vary, whereas in the quenched case

(which

is relevant to

proteins),

the chemical sequence along the chain is fixed by synthesis. In both

cases, the physical behaviour depends on the average hydrophobicity of the polymer chain. For

a strongly hydrophobic chain

(large

fraction of B), we find an ordinary continuous 6 collapse,

with a large conformational entropy in the collapsed phase. For a weakly hydrophobic, or a hydrophilic chain, there is an unusual first-order collapse transition. In particular, for the case of Gaussian disorder, this discontinuous transition is driven by a change of sign of the third virial coefficient. The entropy of this collapsed phase is strongly reduced with respect to the 9

collapsed phase.

(*) Also at Groupe de Physique Statistique, Universit4 de Cergy-Pontoise, 95806 Cergy-Pontoise Cedex, France.

(**) Laboratoire de la Direction des Sciences de la Matibre du Cornmissariat I

l'Energie Atomique.

(***) URA CNRS 1382.

Q Les Editions de Physique 1994

(3)

1 Introduction.

The

study

of random

heteropolymers

in solutions has been

recently developed,

the main reason

being

a

plausible

connection of this

problem

with the protein

folding puzzle

[1, 2]. Various kinds of

quenched

randomness have been considered.

Here,

we wish to

emphasize

the role of the solvent

(water)

in the

folding

process.

Indeed,

it is

widely

believed that the

hydrophobic

effect [3, 4] is the main

driving

force for the

folding

transition: in a native

protein,

the

hydrophobic

residues are burried in an inner core, surrounded

by hydrophilic (or

less

hydrophobic)

residues.

In this paper, we consider a

simple model,

where the monomers of a

single

chain are

(quenched) randomly hydrophilic

or

hydrophobic. They

interact with the solvent molecules

through

an

effective short range

interaction,

which is attractive or

repulsive, depending

on their nature. A related model has been considered

by

Obukhov [5], but he did not

study

its

phase diagram.

We first consider the

corresponding

annealed case, which is

simpler

and

gives

useful

insights

into

the nature of the low temperature

phase.

It may well be of interest for

hydrophilic polymers

in

the presence of

complexing amphiphilic

molecules. In section 2, we describe the model. Section

3 deals with the annealed case, and section 4 with the

quenched

case, both for Gaussian and

binary

distributions of

hydrophilicities.

2 The model.

We consider a chain of N monomers which can be either

hydrophilic (L)

or

hydrophobic (P).

The

hydrophilicity degree

of monomer I is measured

by I,,

and the monomer-solvent interaction is assumed to be

short-ranged (b-interaction).

More

precisely,

we consider a monomer-solvent

interaction of the form:

N+i fit

7ims "

~ ~ l,b(r, Ra) (I)

i=i a=i

where the r, are the

positions

of the N monomers and the

Ra

are the

positions

of the fit solvent molecules.

Here,

the

I,

are

independent

annealed or

quenched

random variables. A

positive (resp. negative)

I

corresponds

to a

hydrophilic (resp. hydrophobic)

monomer.

Using

the

incompressibility

condition of the chain-solvent system [6], I.e. the sum of the monomer and solvent concentration is constant:

pm(r)

+

ps(r)

= po, the solvent

degrees

of freedom can be

eliminated,

and the full Hamiltonian reads:

fl7i(1,)

=

£

[uo +

fl(I,

+

I,

)]

b(r,

2 r, 1#>

+

£

w3

b(r, rj) b(rj rk) (2)

6 1#j#k

+

£

w4 b(r~

rj) b(rj rk) b(rk ri)

~~

'#J#k#1

In

equation (2),

the ~o term accounts for

entropic

excluded volume interactions and represents the

simplest

form of direct monomer-monomer interactions.

(Note

that Obukhov's model [5]

also includes a random

two-body

term

proportional

to

I,lj). Usually,

when

considering

a

polymer collapse transition,

it is sufficient to include terms up to w3.

Here,

we have included

a

repulsive four-body

term w4 which will

play

an essential role.

(4)

N°12 RANDOM HYDROPHILIC-HYDROPHOBIC COPOLYMERS 2141

In the continuous limit

iii,

the

partition

function reads:

with

~ ~

fl~i(>(8))

"

/

d8

/ d8'(~O

+

fl(>(8)

+

>(8~))) b(~(8) ~(8'))

o o

~ N N N

~ ~~

~~'~

~~'~ ~~~~~~ ~~~~~~ ~~~~~~~ ~~~~'~~

(3b)

~ N N N N

+fl /

d8

/ d8'/ d8"/ d8'"b(r(8)-r(8'))

o o o o

6(r(s') r(s")) b(r(s") r(s"'))

where we have

implicitly

assumed that the

origin

of the chain is pinned at 0 and its extremity is free. The parameter a denotes the statistical

segment (Kuhn) length,

and d denotes the

dimension of space. In the

following,

we shall

study

the case where the

hydrophilicities I(s)

are random

independent

Gaussian

variables,

with average lo and variance /h:

P(li)

=

~p

exP

(- ~"~,(°~~ (4)

We shall also discuss

briefly

the

technically

more

complicated

case where the I, are random

independent binary variables, taking

the value lA with

probability

p for a

hydrophilic

monomer

and lB with

probability

q

= 1 p for a

hydrophobic

monomer. This case is more relevant to

synthetic

random

copolymers.

The annealed case amounts to take the average of the

partition

function Z on the random-

ness, whereas the

quenched

case

corresponds

to

taking

the average of

log

Z. Since the

quenched

case is achieved

by introducing replicas

[8] and

using

the

identity log

Z

=

lima-o ~j~,

we

compute

directly

the average W for

integer

n, and then treat

separately

the annealed

(n

= 1) and

quenched (n

=

0)

cases. For Gaussian disorder and

integer

n, the average

yields:

~ /

j~ ~~~~~~

~~~

2~2 ~ ~j~ ~~i~~~

~

~~

~~

~ ~ ~ ~

/~

~

(5a)

x exp

~

'

/

ds

/

ds'

/

ds"

£ b(ra(s) ra(s')) b(rb(s) rb(s"))

2

0 0

~ ~=j

with

N N n

AG "

-(~O

+

2fllO)

ds ds'

£b(ra(s) ra(s'))

2

/ /

~=i

~ N N N n

+@ /

d~

/

d~'

/

d8"

Lb(~a(8) ~a(8')) 6(~a(8') ~a(8"))

° ° °

a=1 ~~~~

~ N N N N

+I£ d8£ d8'£ d8"£

d8"'

L

n

b(~a(8) ~a(8')) b(~a(8') ra(8")) b(~a(8") ~a(8"'))

a=1

(5)

whereas for

binary

disorder:

~ It

~~~~~~

~~

lt ~

~~

t ~~~i~~1~ ~l

~~~~

with

~ ~ ~

AB "

/

ds

/

ds'

£b<ra<s> ra<s'»

2 0 0 ~=i

+~°~

/~

ds

/~

ds'

/~

ds"

fb<ra<s> ra<s'» b<ra<s'> ra(s"»

6 0 0 0 ~=i

~ N N N N

+i£ dS£ dS'£ dS"£

dS"'

~~~,

L

~ b<~a<S> ~a<S'» b(~a<S'> ~a<S"» b<~a<S"> ~a<S"'»

a=1

/~

ds

log (p e~~~~ Ll=i I"

~~'~~~°~~~~~°~~'~~

o

~q ~~~~

Ll=I IT

~~'~(~"(~)~~~(~'))

We now

study separately

the annealed ~nd

quenched

cases.

3. The annealed case.

We set n

= I in

equ~tions

<5> ~nd (6>.

Denoting

the monomer concentr~tion

by p<r>,

we

assume the

validity

of

ground

st~te domin~nce [6, 9]. The

ground

state wave-function

q<r)

is

normalized,

and we

perform

a

saddle-point approximation

of the

partition

function. We quote here

only

the

results, leaving

the details of the calculations for the

quenched

case

<Sect.

4>.

These

approximations yield

the

following

annealed free energy per monomer.

3. I GAUSSIAN DISORDER.

flfG ~[ijjjj~>jj [~j[)~jj[

~

~

j~Tj~ ~~

,

,~

<~~>

<

r

p

TW r o TWr

~~~~~

W~ " W3 3fl~~~ ~~~~

The energy

Eo ~ppe~ring

in (7a> is the

Lagrange multiplier enforcing

the normalization of the wave-function q. The equation of state is obtained

by minimizing

the free energy (7a> with respect to the wave-function

q(r>.

The monomer concentration is

given by:

P(r>

=

Nw~(r>

(8>

The

phase diagram

of the system can be

easily

discussed: for any non-zero concentration of

hydrophobic

monomers, there is a transition between a swollen

phase (p

= 0> at

high

(6)

N°12 RANDOM HYDROPHILIC-HYDROPHOBIC COPOLYMERS 2143

temperature, and a

collapsed phase (finite

p> at low temperature. The nature of the transition

depends

on the average

hydrophobicity

of the chain

lo

and variance I.

For strong average

hydrophobicity <lo

< 0>, there is a second order

collapse transition,

similar to an

ordinary

point for

polymers

in a bad solvent [10]. This occurs when the coefficient of the

q~

term in <7a> becomes

negative

while the coefficient of the

q6

term is still

positive.

This occurs if the average

hydrophobicity lo

satisfies the condition:

->o

>

(~ £<9a>

and the transition temperature is

given by:

To " 210 (9b>

When condition (9a> is not

satisfied,

the

collapse

transition becomes first

order,

and is driven

by

the

change

of

sign

of

w(,

I-e-

by

a

change

of

sign

of the effective third virial coefficient. This is

quite

an unusual mechanism.

Such is the case if the system is

weakly hydrophobic,

or

hydrophilic.

The minimization of the free energy

<7a)

leads to a non-linear

partial

differential

equation

for q which cannot be solved

analytically.

It is

possible

to restrict the minimization to a subset of Gaussian wave-

functions,

and this will be discussed in the section on

quenched

systems <see

Eq.

<20)

below>.

Within this Gaussian approximation, the first order transition temperature T* is

given by

the

equation:

~

~~~~ ~4

~

~~

~~~~~

~3 $) '~~'

3.2 BINARY DISORDER. In this case, the free energy per monomer reads:

~~~~~

G = w~

log (p

e

~~~~~~

+ q

~~~'~~~)

~

~°~~

~~~~~

By taking

derivatives with respect to the chemical

potentials l~,

it. is

easily

seen that the concentrations of

hydrophilic

and

hydrophobic

monomers are

given by:

~+'~' ~~'~'~'p ~Nj~~~~i)~NpAB~~

~ q eNpAB~~ ~~~~

~~'~' ~~' '~'p

e~NpAA~~

+ q eNpAB~~

Note the Boltzman

weights

in (11>, which govern the chemical composition of the

polymer.

It

clearly

appears that the

regions

of

high

monomer concentration are

hydrophobic,

whereas those of low concentration are

hydrophilic.

The main conclusions of the Gaussian case above remain

valid,

except that the transition temperatures cannot be calculated

explicitly.

We find

again

an

ordinary

second order point for

strongly hydrophobic chains,

and a first-order

collapse

for

weakly hydrophobic

or

hydrophilic

chains.

(7)

As can be seen

by

a numerical

study

of this case, the transition is

strongly first-order,

and this

implies

the existence of a

large

latent heat at the

transition,

and thus a

sharp drop

of the entropy. This

drop

in entropy can be

interpreted

as the formation of a

hydrophobic

core, surrounded

by

a

hydrophilic coating,

which is

energetically favourable,

but

entropically

unfavourable. This is confirmed

by equations

II>, which show that

high

concentration

regions

are

hydrophobic,

whereas low concentration

regions

are

hydrophilic. Finally,

we see from (11>

that at low temperature, the total number of

hydrophilic

monomers goes to zero, whereas the number of

hydrophobic

monomers goes to N.

4. The

quenched

case.

We have to take the n

= 0 limit in (5> and (6>.

Introducing

the parameters

qab(r,r'>,

with

a < b, and

pa(r> by:

N

qab(r, r'>

=

/

ds

b(ra(s>

r>

b(rb(s> r'>

°~

(12>

pa(r>

=

/

ds

b(ra(s>

r>

o

we may write

equations

(5> as.

4. I GAUSSIAN DISORDER.

@

=

Dqab(r, r'>D#ab(r, r'>Dpa(r>D#a(r>

exp

(G(qab, lab,

Pa,

la>

+

log ((#ab, la» (13a>

where

#ab(r,r'>

and

#a(r>

are the

Lagrange multipliers

associated with (12>, and:

G(~ab, lab,

Pa,

la>

"

/

dT

L lPa(~>#a(~>

(U0 + 2fl>0>

~j~' ~Pl(~> )(Pl(~>)

~

+

/

dr

/

dr'

£ (iqab(r, r'>#ab(r, r'>

+

fl~l~ qab(r, r'>pa(r>pb(r'>)

a<b

(13b>

and

((4ab, ~a'

"

fl ~~a(S'

(13C~

~

where

w(

is defined in

equation (7b>.

Using

the

identity

of equation (13c> with a

Feynman integral

[9,

iii,

we can write:

((@ab,

la>

"

/ fl d~ra

< ri

rn(e~~~"~4ab'~a~(0.

0 >

(14a>

~

(8)

N°12 RANDOM HYDROPHILIC-HYDROPHOBIC COPOLYMERS 2145

where Hn is a

"quantum-like"

n - 0

Hamiltonian, given by:

~n £ V~

+

£ l~a(ra'

+

£ I#ab(ra,

rb

(14b~

a a a<b

Anticipating

some kind of

(hydrophobically-driven> collapse,

we assume that we can use

ground-state

dominance to evaluate (14>, and write,

omitting

some non-extensive

prefactors:

j(#~~~

j~,

Qf

e~NE0(4ab,S)

= exp

(-N

l~(r)Imin

(< lP(Hn(il

>

-Eo (<

lP(lP >

-1>))

~~~~

where Eo is the

ground

state energy of Hn. At this

point,

the

problem

is still

untractable,

and we make the extra

approximation

of

saddle-point

method

(SPM>.

The extremization with

respect to qab reads:

I#ab(r,r'>

=

-fl~ l~pa(r> pb(r'>

(16>

This

equation

shows that

replica

symmetry is not

broken,

since

jab

is a

product

of two

single- replica quantities

pa. To get more

analytic information,

we use the

Rayleigh-Ritz

variational

principle

to evaluate Eo. Due to the absence of

replica

symmetry

breaking (RSB>,

we further restrict the variational wave-function space to Hartree-like

replica-symmetric wave-functions,

and write:

n

~P(~l;

,~n>

"

fl w(ra>

(17>

a=I

Because of

replica

symmetry, we can omit

replica

indices and take

easily

the n - 0 limit. The variational free energy now reads:

~fll~G(q,

4,

P,~,

§~' "

/dT lp(r~#(r~

(U0 +

2fl~0' ~j~~

~P~(~' j'(P~(~')

/

dT

/

dT'

liq(r, r'>4(r, r'>

+ fl~>~

q(r, r'>p(r>p(r'>)

~2

~~~'

Nl/d~T

W(r>

liv~

+ i

<(r>

W(r>

/

~~~

/ ~~~'~~~'~"~~~~'~~~~"

~

~~°~ / ~~~~~~~' ~'

The SPM

equations

re~d:

P(r>

=

Nw~(r>

q(r, r'>

=

Nw~(r>w~(r'>

wj

~ w~ ~ ~ ~

j

~, , , (19>

~~~~~ ~~° ~ ~~~°~~~~~ ~

2 ~ ~~~ ~

6 ~ ~~~ ~

~

~ ~ ~

~~~'~

~~~~

14(~,

~'> "

~fl~>~P(~>P(~'>

We still have to minimize with respect to the normalized wave-function

q(r>.

This le~ds to a very

complicated

non-linear

Schr6dinger

equation, and we shall restrict ourselves to a one-

parameter

family

of Gaussian wave-functions of the form:

i d/4 ~2

~~~~

27rR2

~~~ 4R2 ~~~~

(9)

where R is the

only

variational parameter.

Using equations

(18> and (19>, the variational free energy per monomer

()

"

GIN>

reads:

Note that the

only

difference between this free energy and the annealed one of equation

(7a),

is the presence, in the

quenched

case, of an additional disorder induced term, which we have

grouped together

with the

q6

term for reasons that will become clear below.

Using

the Gaussian

wave function

(20>,

and the value of

w( given

in

(7b>,

the free energy reads:

~~~ i~~

~

2j~>~

~~~ ~

~

~~~

~~~~

~

(27r~$>~ ~ 2~>~ ~~~

~~ ~~~ ~

~~

~

~

3~7r~

~~~

~~

~

At low temperatures, one has to

study

the

sign

of the third term of

equation (22>.

This

sign

becomes

negative

at low temperature, due to disorder fluctuations.

It is easy to see that this result is not restricted to the choice (20> of the variational function

q(r>. Indeed, using

the

inequality Jd~r(q3(r>

+

xq(r»2

>

0,

for all x>, one finds that the third virial coefficient of

equation

(21> becomes

negative

at low

enough

temperatures, for all

space-dependent

normalized variational functions

q(r>.

We therefore

get

the

following

results:

I> lo > 0: the

hydrophilic

case. In our

approximation,

we find that there is a first

order transition towards a

collapsed phase (R

r-

N~/~>

induced

by

a

negative three-body

term

(and

stabilized

by

a

positive four-body term>.

This transition is neither an

ordinary point (since

the

two-body

term is

positive>,

nor a

freezing

point

(the replica-symmetry

is not

broken>.

Note that since our

approach

is

variational,

the true free energy of the system is lower than the variational one, and we thus expect the real transition to occur at an even

higher

temperature. Since the transition is first

order,

we expect

metastability

and retardation effects to be important near the

transition;

we also expect that

(due

to the latent heat >, there will be

a reduction of entropy in the low temperature

phase, compared

to an

ordinary

second order

point.

ii> lo < 0: the

hydrophobic

case. In this case,

defining

the two temperatures:

To "

2(lo(/uo

~ ~

/

(23>

/w3(1 3d/2/2d>

leads to two

possible

scenarii, in a very similar way as in the annealed case:

a>

Weakly hydrophobic

case:

Ti

> To- The

collapse

transition is

again

driven

by

the disorder fluctuations of the

three-body

interactions. The

resulting

first-order transition is very similar to the

hydrophilic

case I>.

b>

Strongly hydrophobic

case: To > Ti The

collapse

transition is now driven

by

the strong

twc-body

lo term. The

resulting phase

transition is very similar to an

ordinary point,

and is therefore second-order.

(10)

N°12 RANDOM HYDROPHILIC-HYDROPHOBIC COPOLYMERS 2147

The

phase diagram

of the

quenched

case seems

quite

similar to the annealed case,

but,

as we shall see on the case of

binary disorder,

the

physics

is

quite

different. Note also that the transition temperature in the

quenched

case is lower than in the

corresponding

annealed

case, as can be checked from (7a> and (21>. This may be viewed as an effect of

geometrical

frustration induced

by

the chain constraint.

4.2 BINARY DISORDER.

Using

the same

approximations

as in the

previous

case,

namely

ground

state

dominance,

no

replica

symmetry

breaking

and Hartree variational

wave-function,

the free energy per monomer reads:

~~~ '~ ~~~~~~~~~~~

~ ~

~~~~~~~~

+N~'°~ / d~rq~(r>

+ '°~

N~ / d~rq~(r>

Eo

/ d~rq~(r> (~4,

6 24

+~~ffl~A

/

d~T§~~

(~' i~ ~~~

~

1°g /

d~T§~~

(~' ~~~'~ ~~~~)

~_~ P

where I

= AA + lB.

In this case, we see that even the Gaussian variational form is not

tractable,

but a numerical

study

shows that the system has

exactly

the same phase

diagram

as for Gaussian disorder.

However,

the interest of this model is that it allows to calculate the concentration of

hydrophilic

and

hydrophobic

monomers. We obtain:

P+(~'

"

NW~(r>

I j§(-i,i+i (q) eP'lN~~

I=1 i'

i

ddT§a2eflAlN~2

m

P-(r>

=

Np2(r~ £(_i,i+i

q

eP'lN~2

(25>

1=1

i'~

fddT§a2eflAlN~2

It is

easily

seen on (25> that the two concentrations

satisfy

the sum rules:

/d~TP+(4

=

NP

/ d~rp-(r>

~~~,

=

Nq

and thus

(as expected

for a

quenched chain>,

the overall chemical

composition

of the chain remains

unchanged.

5 Conclusion.

We have studied a

simple

model of a

randomly hydrophilic-hydrophobic

chain in water. The randomness may be annealed or

quenched,

and we have considered

binary

and Gaussian dis- tributions. Let us first summarize some technical aspects of our

study.

We have used the

ground

state dominance

approximation

[6,

9],

which is

certainly

apprc- priate in a

collapsed phase.

We have further used a

saddle-point method, together

with a

restricted variational procedure. These extra

approximations

are more difficult to assess, but should be correct at a mean-field level

(see below>.

(11)

We now consider the

physical

output of this calculation. Results are very similar for annealed and

quenched

randomness. The low temperature

phase

is

always collapsed; depending

on the

degree

of

hydrophobicity

of the

chain,

this

collapsed phase

is reached either

through

an usual second order

point,

or

through

a first order transition. In both cases, critical fluctuations are

expected

to be weak. As

expected,

the interior of the compact

globule

is

mainly hydrophobic,

whereas its exterior is

hydrophilic (see Eqs.

(11> and

(25».

Note however that the detailed connection between the monomers and their

spatial positions

is lost when one averages over the disorder. We would like to stress that the first order transition,

resulting

from equation

(22>,

does not

imply

that the

hydrophilic

(A> and

hydrophobic

(B> monomers

undergo

a

macroscopic phase

separation. Since

q2(r>

is

proportional

to the total

polymer density,

one may

identify

the

coexisting phases

as the solvent and a

copolymer-rich phase:

the first order transition can be

thought

of as a

demixing

transition, due to the disorder fluctuations. The detailed

study

of the

spatial

structure of the

chain,

as well as the

properties

of the chain-solvent

interface,

are

beyond

the scope of this work.

Finally,

the

quenched

case may well be relevant to proteins, since

hydrophobic

residues

play

a major role in the

folding

transition.

Indeed, hydrophobicity-based

codes have been

previously proposed by

Goldstein et al. [12], to

predict

some folded structures. Their

approach

is however

linked to the spin

glass approach

to protein

folding,

which differs from the

approach

of the present paper. References

ill

and [2]

give

a more

complete

view of the disordered

polymer approach

to the

protein folding puzzle.

Note added in

proof:

A first order

collapse

transition for random annealed

copolymers

has been obtained

by Grosberg

in Biofizika 29

(1984>

569. We thank A.

Grosberg

for

bringing

this work to our attention and for

stimulating correspondence.

We also benefited from

interesting

discussions with A. Gutin

and E. Shakhnovich.

References

iii

Karplus M. and Shakhnovich E-I-, Protein Folding, T-E- Creighton Ed.

(W.H.

Freeman, New

York, 1992> and references therein.

[2] Garel T., Orland H. and Thirumalai D., New Developments in Theoretical Studies of Proteins,

R. Elber Ed.

(World

Scientific, Singapour, 1994> and references therein.

[3] Tanford C., The hydrophobic effect

(Wiley,

New-York,

1980).

[4] Dill K-A-, Biochemistry 29 (1990> 7133.

[5] Obukhov S-P-, J. Phys. A 19 (1986> 3655.

[6] de Gennes P-G-, Scaling concepts in polymer physics

(Cornell

University Press, Ithaca, 1979).

[7] des Cloizeaux J. and Jannink G., Les Polymbres en Solution

(Editions

de Physique, Les Ulis, 1987>.

[8] Mdzard M., Parisi G. and Virasoro M-A-, Spin glass theory and beyond

(World

Scientific, Sin- gapore, 1987).

[9] Wiegel F-W-, Introduction to path integral methods in physics and polymer science

(World

scientific, Singapore, 1986>.

[10] Flory P-J-, Principles of polymer chemistry

(Cornell

University Press, Ithaca, 1971>.

ill]

Edwards S-F-, Proc. Phys. Soc. London 85

(1965)

613.

[12] Goldstein R-A-, Luthey-Schulten Z-A- and Wolynes P-G., Proc. Nat]. Acad. Sci. USA 89

(1992

4918 and references therein.

Références

Documents relatifs

class of normal Helly circular-arc graphs. In summary, we obtain the following results: we prove that the bend number of normal circular-arc graphs with respect to EPR

The bars represent the daily intake per individual (mg) for each food pairing. Food collection was measured over 12 days in group of termites composed of 70 major workers and 30

After a variable delay (1500 ± 2200 ms) a second auditory signal constituted the GO for the motor task. The motor task was performed using the stimulated right hand in both

preventing  cross  reactions  (negative  selection)  while  more  flexibility  is  allowed... Mornon,

D’après ces observations, nous constatons que le protocole WASRM soit fiable avec un nombre important des membres de groupe multicast (récepteurs) en donnant un

Elle est incapable dans son état actuel d’assumer ce rôle .En effet ,l’absence d’activités et des grands équipements structurants (aéroport, etc...) de

Prior to desalination process of NaCl solutions, membranes were tested in contact with pure water as a feed solution to evaluate the nominal value of water permeate flux (Figure