• No results found

Coil–helix transition

In document This page intentionally left blank (pagina 41-51)

3f (α)

α4 = 0, (1.87)

for the free energy, whereα1is the swollen state andα2is the collapsed state (Figure 1.8(a)).

The nature of the CG transition has been investigated by many researchers. A typical example is polystyrene (PS) in the solvent cyclohexane (θ =34.5C). Neutron scattering, light scattering, osmotic pressure measurements, and viscosity measurements have all confirmed the points (1)–(3) above, but no consensus has yet been reached about the nature of the transition (4). Light scattering experiments on PS of ultrahigh molecular weight (Mw= 2.6 × 107) indicate that the transition is very close to the discontinuous one withy  yc(Figure1.8(b)) [21].

1.8 Coil–helix transition

Some polymers, such as isotactic polypropyrene (iPP), polyisobutadiene (PIB), and poly(ethylene oxide) (PEO), form helices in the crystalline state. A helical structure is represented bypmwithp number of monomers and m the number of turns in one period of the helix. The lengthd of one period is called pitch of the helix. The length along the helical axis per monomer is then given by

b ≡ d/p. (1.88)

For example, iPP forms a 31helix withd = 0.65 nm, and hence b = 0.22 nm; PIB forms an 85helix withd =1.863 nm, b=0.233 nm; and PEO forms a 72helix withd =1.93 nm, b = 0.28 nm(see Figure1.9). Short-range interactions (local interactions), in particular hydrogen bonds, are important for stabilization of a helical conformation.

Several synthetic polypeptides, such as the poly(L-amino acid)s, the poly(γ -L-glutamate) (PBLG), poly(β-benzyl-L-asparate) (PβBA), and poly(L-glutamic acid)

d = 1.93 nm

Fig. 1.9 Helix structure 72of poly(ethylene oxide). The period isd = 1.93 nm. One period includes seven ethylene oxide monomers and forms two turns.

Hydrogen Bonds

R

R

R

R

R R

R R

R R

Fig. 1.10 Alpha helix of polypeptide. Type 185has a period of 2.7 nm. Hence, 3.6 residues (0.54 nm) form a turn. The pitch per monomer isb = 0.15 nm.

(PLGA), also formα-helices in the solid state (see Figure1.10). When they are dis-persed in strongly interacting solvents, polymer chains are not merely separated from each other but also change their conformation from helical to random coil. For instance, PBLG forms anα-helix with b =0.15 nmin chloroform(CF), but melts into a statistical randomcoil in dichloroacetic acid (DCA). Hence, the chain changes its conformation in mixed solvents of CF and DCA depending on the solvent composition. Such a con-formation change is an example of coil–helix transition (referred to as CH transition).

In the transition state, a chain generally takes a conformation with rod-like rigid helices of polydisperse length that are sequentially connected by randomcoil segments. CH transition may also be induced by changing other environmental parameters, such as temperature or pH.

Theoretical studies of CH transitions focus attention on the behavior of the poly(amino acid) in the transition region. The most basic information is the change in the fraction of the residues in helical states as functions of the molecular weight of the chain and of the environmental parameters (solvent compositionx, temperature T , and pH).

Fromthe late 1950s, many papers in the literature studied this problem[23]. Most of them employed either the matrix method or the generating function method to calculate the chain partition function. However, in order to apply the theoretical method directly to many chain problems in solutions and gels, we here reformulate the single chain problem using the combinatorial counting method.

Consider a polymer chain carrying a total numbern of statistical units (Figure1.11).

Let the symbol 0 indicate a monomer (an amino acid) in the random coil part, and let 1 indicate the same in the helical part. Letu be the statistical weight of the adjacent pair (0,0) of monomers,v be that of a pair (1,1), and w be that of the pairs (0,1) and (1,0). These can be derived by integrating over the rotational angle of a monomer under the potential of internal rotation (1.9). Because a hydrogen bond is formed between

1 n

ζ

1 1 v v v

Fig. 1.11 Statistical weight of the conformation of a chain with mixed helical parts (1) and random coil parts (0). After integration over the internal rotation angle, we find the statistical weightsu,v, andw for the adjacent pair (0,0), (1,1) and (0,1), (1,0). The number of repeat units (an amino acid) in the helix is designatedζ .

neighboring pairs in the helical part, the statistical weightv is different from u for the pair in a coil part.

The partition function of the chain is then given by the form Zm(T ) =

(u)w(vv)w(uuuu)w(vvvvvv)··· , (1.89)

where the summation should be taken over all possible distributions of the helical parts along the chain under the given numberm of the helical monomers.

Let us study this partition function froma different viewpoint. In order for the helices to be generated on this chain, helical sequences must be selected from the finite length n. Let jζ be the number of helices with lengthζ = 1,2,3,...,n (counted in the number of statistical repeat units).

We first consider that helices are temporarily contracted into single units. The total length is therefore reduced ton= n −

ζjζ. (In order to distinguish the neigboring helices, we assume that there should be at least one nonhelical monomer between them.)

The number of ways to choose

jζunits fromnis given byn!/(

jζ!)(n− jζ)!, but since we cannot distinguish the states that are obtained by exchanging helices of the same length, we instead have to multiply by the factor(

jζ)!/( jζ!).

We thus find that the number of different ways to select j≡ {j1,j2,j3,...} sequences is given by

ω(j) = (n−

ζjζ)!

(

jζ!)(n−

ζjζ−

jζ)!. (1.90)

We next divide the partition function by its valueunin the reference state of the perfect randomcoil, and introduce the relative statistical unitss(T )≡v/u. The ratio w/u is then

expressed as σ s(T ), where σ ≡ w/v is associated with each boundary between the neighboring coil part and the helical part.

In general, for a run ofζ helical monomers, the statistical weight

ηζ= σs(T )ζ (1.91)

is assigned [24]. The parameterσ associated with the helix boundary is called the helix initiation (nucleation) parameter, or the cooperativity parameter. If it is small, the probability to create the first hydrogen bond to generate the helix (nucleation of the helix) is low due to the large penalty for adjusting the local conformation to form the hydrogen bond. But once one bond is formed, adjacent bonds are formed more easily, so that there is a strong tendency to formcontinuous chains of bonds.

The partition function of a chain measured relative to the random-coil conformation is then given by

Zm(T ) =

j

ω(j)

ζ

ζ)jζ. (1.92)

Because the total numberm ≡

ζjζ of helical monomers is not a fixed number but thermally controlled, we introduce the activityλ of the helical monomers, and move to the grand canonical partition function

F(λ,T ) ≡

m≥0

Zm(T )λm. (1.93)

This function is the helical counterpart of the binding polynomial in the literature [25]

on biomacromolecules, which is frequently used to study the adsorption of ions, legands, protons, etc. onto proteins.

In order to find the most probable distribution (m.p.d.) of helices, we maximize the partition function (1.93) withλ = 1, or minimize the free energy G(T ) of a chain, by changing j. The condition is

∂jζ

lnω(j)+

ζ

jζlnηζ

 =0. (1.94)

By using the Stirling formula for lnω(j), we find that the m.p.d. is given by

jζ/n = (1−θ −ν)ηζzζ, (1.95)

where

θ ≡n

ζ =1

ζjζ/n (1.96)

is the average helical content (number of statistical units in the helical parts divided by the total number of units), and

ν ≡

n ζ =1

jζ/n (1.97)

is the average number of helices on the chain. The parameterz in (1.95) is defined by

z ≡ (1−θ −ν)/(1−θ). (1.98)

Substituting the distribution (1.95) into these definitions, we findθ and ν as

θ = zV1(z)/[1+zV1(z)], (1.99)

and

ν = zV0(z)/[1+zV1(z)], (1.100) where the functionsV (x) are defined by

V0(z) ≡n

ζ =1

ηζzζ, V1(z) ≡n

ζ =1

ζηζzζ. (1.101)

Similarly, substituting (1.95) back into the original partition function (1.92), we find

F(T ) = z−n, (1.102)

whereF(T ) ≡ F(1,T ), or equivalently,

G(T ) = kBT ln z. (1.103)

By definition (1.98), the parameterz must satisfy the equation z

1−zV0(z) = 1. (1.104)

Ifn is allowed to go to infinity in V0(z), this is the same equation as that found by Zimm and Bragg [24] (referred to as ZB):

z 1−z

σ sz

1−sz= 1. (1.105)

The solutionz corresponds to the reciprocal λ−10 of the larger eigenvalueλ0of the original ZB secular equation. However, because the upper limit of the sum inVk is limited to the total numbern of repeat units, the effect of the molecular weight on the transition is easy to study in the present theoretical framework.

1.0

0.8

0.6

0.4

0.2

0.0

z, θ, ν, ζ

–2 –1 0 1 2

TEMPERATURE ln s θ

ν z

n = 100

σ = 0.01 ζ/n

Fig. 1.12 Helix contentθ (solid line), number of helices ν (broken dotted line), mean helix length ¯ζ (broken line), and probabilityz (thin broken line) for a randomly chosen monomer to belong to the random coil part shown as functions of the temperature. The temperature is measured in terms of lns = const +| H|/kBT by using the probability s of hydrogen-bond formation.

Substituting the form(1.91), we find

V0(z) = σszw0(sz), V1(z) = σszw1(sz), (1.106) where the functionsw0andw1are defined by

w0(x) ≡

n ζ =1

xζ −1, w1(x) ≡

n ζ =1

ζxζ −1. (1.107)

Lifson and Roig [26] used a slightly different weight forηζ:

ηζ= v (for ζ = 1), v2wζ −2(for ζ ≥ 2). (1.108) The result does not differ significantly, so that, in the following study, we employ the simpler ZB weight.

Figure 1.12 plots z, θ, ν, and the mean helix length ¯ζ ≡ θ/ν as functions of the temperature. Temperature is measured in terms of lns(T ). The CH transition takes place at around lns = 0. The transition becomes sharper for a smaller nucleation parameter σ (stronger cooperativity). The transition also becomes sharper with molecular weight, and becomes a real phase transition with discontinuousθ in the limit of infinite chain length.

Consider next the CH transition of a polymer chain under tension applied at the chain end (Figure1.13) [27]. For simplicity, let us also assume that the helices are rigid rods and have a pitchd with p monomers in one period. The length along the rod axis per monomer is then given byb (1.88). The length of a helix with monomer sequenceζ is given bybζ .

R f f

Fig. 1.13 Polymer chain forming helices under an applied tension f . Coil parts and helical parts appear alternately along the chain.

Let ei be the unit vector specifying the direction of thei-th helix along the chain, and let rkbe the end-to-end vector of thek-th randomcoil part along the chain. We then have the relation

R=

k

rk+b

i

ζiei (1.109)

for the end-to-end vector of a chain.

Letm be the total number of repeat units in the helical parts. The canonical partition function of a chain with specifiedm and R is written as

Zm(T ,R) =

{j}

ω(j)

ζ

ζ)jζ



···

k

ρ(lk)dlk

i

ρζ(li)dli, (1.110)

under the condition (1.109), whereρζ(l) is the connectivity function (1.3) for the helix of lengthζ.

We next move to the ensemble where the external tension f is the independent variable, and integrate over end-to-end vectors R and orientation ei of helical rods. We then find

Qm(T ,f) ≡



dRZm(T ,R)eβf·R= ˜g(t)n−m

j

ω(j)

ζ

ζ)jζ n

i=1

˜g(κtζ)jζ, (1.111)

where the function ˜g(t) is the Laplace transform(1.16) of the connectivity functionρ(l), t ≡ f a/kBT is the dimensionless tension, and

κ ≡ b/a (1.112)

is the helical pitch per monomer in the unit of the fundamental step length of a repeat unit.

We next introduce the activityλ for a helical monomer, and move to the grand partition function

F(T ,λ,f) ≡

n m=0

λmQm(T ,f)

= ˜g(t)n

j

ω(j)

ζ

ζφζ(t)λζ)jζ, (1.113)

where the new functionφζ is defined by

φζ(t) ≡ ˜g(κtζ)/ ˜g(t)ζ. (1.114)

At this stage, we can see clearly the effect of tension on the CH transition. The statistical weightηζ of a helix with lengthζ is changed to

ηζ−→ ηζφζ(t), (1.115)

where the factorφζ includes the effect of the orientation ˜g(κτζ) of a rod-like helix, and the entropic force ˜g(τ)−ζfromthe corresponding randomcoil segments. In fact, by taking the logarithmof the total statistical weight of a helix, we find that the free energy of a helical sequence of lengthζ is given by

fζ(τ)/kBT = −ln ηζ−ln[sinh(κτζ )/κτζ ]+ζ ln[sinh τ/τ]−ζ ln λ. (1.116) By minimizing this free energy with respect toζ for a given statistical weight ηζ, we can see in a simple way that the average helix length is increased by stretching to the limit where they are finally destroyed. The physical reason why helices are enhanced by tension is that the linear growth of rod-like helices gains a larger end-to-end distance than that of the randomcoils, and hence it is advantageous for a chain under tension.

The m.p.d. of helices is found by maximizing the grand partition function (1.113) by changingjζ. As before, we find

jζ/n = (1−θ −ν)ηζφζ(t)(λz)ζ (1.117)

by variational calculation. The parameterz is defined by

z ≡ (1−θ −ν)/(1−θ). (1.118)

To see the physical meaning of this parameter, we substitute the equilibrium distribu-tion (1.117) into the grand partidistribu-tion funcdistribu-tion, and fix atλ = 1. We find that it is given by

F(T ,t) = [ ˜g(t)/z]n. (1.119)

Since the probability p(m = 0) for finding a completely random coil is given by

˜g(t)n/F(T ,t), we find

p(m = 0) = zn, (1.120)

and hence the physical interpretation of the parameterz is the probability such that an arbitrarily chosen monomer belongs to the random coil part.

Repeating the same procedure given above under no tension, we find

θ = zV1(t,z)/[1+zV1(t,z)], (1.121)

and

ν = zV0(t,z)/[1+zV1(t,z)], (1.122) where the functionsV (t,z) are defined by

V0(t,x) ≡n

ζ =1

ηζ ˜φζ(t)zζ, V1(t,x) ≡n

ζ =1

ζηζ˜φζ(t)zζ. (1.123)

The condition to findz is

z

1−zV0(t,z) = 1. (1.124)

This is basically the ZB equation, but here it is properly extended to include the effect of tension. The solution of this equation gives the probabilityz(t) as a function of the temperature and the external force.

Let us next find the tension–elongation curve. The average end-to-end distanceR can be found by the fundamental relation

R =

∂f[kBT ln F(T ;λ,t)], (1.125) so that we have

R/na = L(t)−(∂z/∂t)/z

= (1−θ(t))[L(t)+κzW1(t,λz)], (1.126) whereL(t) is the Langevin function (1.27), andWkis defined by

Wk(t,x) ≡n

ζ =1

ζkηζφζ(t)L(κtζ )xζ, (1.127)

fork = 0,1,2,....

To find the average square end-to-end distance R2 0in the absence of the external force we expand the partition functionF(T ,t) in powers of the dimensionless force t.

Formal expansion gives

F(T ,t)/F(T ,0) = 1+( R2 0/6a2)t2+··· , (1.128) so that we can find R2 0fromthe expansion

z = z0+z1t2+··· (1.129)

of the parameterz and the expansion of the relation (1.119) in the form

R2 0/na2= 1−6z1/z0. (1.130) Here,z0is the solution of (1.124) under no forcet = 0.

Explicit calculation ofz1leads to

R2 0/na2= [1−θ(0)]+κ2¯ζwθ(0), (1.131) whereθ(0) is the helix contents at t =0, which was studied in the preceding section, and

¯ζw≡ V2(0,z0)/V1(0,z0) (1.132) is the weight-average helix length of the chain under no tension (see Figure1.14).

The end-to-end distance as a function of the tension is calculated in a similar way as before. We find

R(t) = R(c)(t)[1−θ(t)]+κR(h)(t)θ(t), (1.133)

0 2

–1 0 1 2 3 4

0.001

0.01

 = 0.1 5

Temperature ln s 1

Helix Content θ

<R2>/na2

1 0.5 0

Fig. 1.14 Mean square end-to-end distance (solid lines) and helix content (dotted lines) plotted against the temperature. A minimum appears in R2 at the coil–helix transition temperature.

whereR(c)andR(h) are the end-to-end functions defined for the coil part and helical part in a similar way as above [28,29].

In document This page intentionally left blank (pagina 41-51)