PHYSICAL REVIEW E

VOLUME 62, NUMBER 6

DECEMBER 2000

Spin-1 aggregation model in one dimension M. Girardi* and W. Figueiredo† Departamento de Fı´sica, Universidade Federal de Santa Catarina, 88040-900 Floriano´polis, Santa Catarina, Brazil 共Received 5 July 2000兲 We studied a simple model of aggregation in one dimension that resembles the self-assembly of amphiphiles in an aqueous solution. We mapped the water and amphiphilic molecules by Ising spin variables for S⫽1. The zero component of spin represents the water molecules, while the remaining components (⫾1) account for the amphiphilic molecules. We defined an aggregate in one dimension by a set of spin components (⫾1) placed between two zero spin components. There is no difference between up and down components of the spins inside the aggregates. In this way what really matters is the square of the spin component. The grand-canonical partition function and the probability of formation of different aggregate sizes were calculated by the transfer matrix method. We have shown that for any value of the chemical potential and temperature, the system does not exhibit the typical aggregate size distribution which is observed in micellar solutions at low concentrations. The distribution curve for the aggregate size does not show the minimum and the maximum as a function of the concentration which is the signature of the appearance of micelles. We can say that this one-dimensional model does not present any phase transition nor a transition from the micellar to nonmicellar state. PACS number共s兲: 82.60.Lf, 64.75.⫹g, 64.60.Ht

I. INTRODUCTION

It is well known that one-dimensional systems with shortrange interactions do not exhibit a phase transition from an ordered to disordered state 关1兴. In spite of this fact, shortrange forces can give rise to the formation of islands containing a different number of particles at finite temperature. In this work we are concerned with the formation of a special type of aggregates, which we think can mimic the selfassembly of amphiphilic molecules in aqueous solutions 关2–5兴. Some recent models have been presented in the literature to describe the linear self-assembly of molecules. Henderson 关6兴 considered as an example of chain forming amphiphilic solution, a mixture of solvent 共A兲 and solute 共B兲 with repulsive interactions between them, opposed to the interactions (AA) and (BB) considered attractive and of equal magnitude as that of the (AB) pair. He obtained an exact expression for the cluster distribution and showed that the model cannot display a phase transition. Duque and Tarazona 关7兴 presented a simple one-dimensional model with molecular interactions favoring the formation of clusters with a defined optimal size. Their model is a system of hard rods of fixed length, moving along a line and with internal degrees of freedom. They found an exact solution that incorporates excluded volume effects and molecular attraction. They have shown that, at low molecular concentrations, the curve of the chemical potential against molecular density changes its slope at a given concentration. This could be interpreted as the equivalent to the CMC 共critical micellar concentration兲, which is in qualitative agreement with real systems of amphiphilic molecules. Kolomeisky and Widom 关8兴 also considered a one-dimensional lattice model for the hydrophobic attraction. In their model, the accomodation of the solute molecules in the lattice depends on the state of nearest-

*Email address: [email protected]

Email address: [email protected]

1063-651X/2000/62共6兲/8344共5兲/$15.00

PRE 62

neighbor solvent molecules. The model was exactly solved in one dimension, and they found the condition of hydrophobicity at temperature T as a function of the interaction parameters of the model, and the number of possible states of the solvent molecules. The model we consider here consists of a linear array of sites, where each site is occupied by a variable with three possible values. One of these values 共0兲 represents the solvent molecules, while the other two (⫾1) are associated with the amphiphilic molecules. We use the spin 1 Ising model to describe the interactions between nearest neighbor sites, and we restrict the number of components equal to zero to a fixed value, that is, the number of solvent molecules determines the value of the concentration of the solution. We use the transfer matrix method to compute the grand-canonical partition function of the model, and to calculate the probability of finding an aggregate of a given size. We also performed Monte Carlo simulations on this model, and we have seen that the evolution of the system towards its equilibrium states is driven by two simultaneous stochastic processes. The dynamics processes that fit our model are the following: one that conserves the order parameter, of the exchange Kawasaki type 关9兴, and the other, where the order parameter is not conserved, related to the single spin-flip Glauber kinetics 关10兴. We show that this model, albeit very simple, is capable of predicting a change in the slope of the curve of the free amphiphilic molecules against total concentration. We also show that the distribution curve for the aggregate sizes does not exhibit the local minimum and maximum characteristic of micellized systems 关5兴. For a typical system in its micellar state, the local minimum in the aggregate-size distribution curve lies always below the concentration of free amphiphiles, and it is different from zero. The local maximum in this curve, which appears for larger values of the aggregate size than that observed at the minimum, indicates the typical size of the micelles. This work is organized as follows. In Sec. II we present the model and the calculations. In Sec. III we describe the Monte Carlo simulations. In Sec. IV we present our results and conclusions. 8344

©2000 The American Physical Society

PRE 62

SPIN-1 AGGREGATION MODEL IN ONE DIMENSION II. MODEL AND CALCULATIONS

We consider a lattice model to describe the aggregation of amphiphilic molecules in a solvent. The simplest model we can imagine is a linear array of sites, where each site is occupied by an amphiphilic or by a solute molecule. We choose to represent these molecules in the language of spins, and we make use of the spin 1 Ising model, with three possible components. In this way, to each solute molecule we associate the zero component of the spin, and the two other components (⫾1) are attached to the amphiphilic molecules. For instance, the values (⫾1) are related to the hydrophobic and hydrophilic properties of the amphiphilic molecules. If two nearest neighbor amphiphiles have the same value of the spin components, it can be thought that two hydrocarbon chains are touching each other in a real solution. On the other hand, two antiparallel spin components mimic the repulsion between the hydrophilic head and the hydrophobic chain of the amphiphile. We define an aggregate in one dimension by a set of spin components (⫾1) placed between two zero spin components. There is no difference between an up and down component of the spin inside the aggregate, but the different orientations of spins can take into account the possible values of the aggregate energy. In order to simulate a fixed concentration in the real solution, in this model, the number of zero spin components is made constant. We define the following Hamiltonian model for the system:

␭⫽

8345



1 a 1⫹ ⫹ab 2 b ⫹

冑冉

a a2 1⫺2 ⫺2ab⫹ 2 ⫹2a 2 ⫹a 2 b 2 ⫹8a b b

冊册

, 共5兲

and we defined a⫽exp(␤␮) and b⫽exp(␤J). In the next step we define the probability to find an aggregate of size n in the system. It is chosen as the sequence (0 ⫾ 1 ⫾ 1 ⫾ 1•••⫾ 1 ⫾ 1 0), where we have n nonzero spin components between two zero components. Therefore, the probabilility is given by P 共 S k ,S k⫹1 , . . . ,S k⫹n⫹1 ,S k⫹n⫹2 /0,⫾1, . . . ,⫾1,0兲 ⫽ P共 S 兲 ⫽

1 具 ␦ 共 1⫺ ␦ S k⫹1 ,0兲 ••• 共 1⫺ ␦ S k⫹n⫹1 ,0兲 ␦ S k⫹n⫹2 ,0典 ⌶ S k ,0 共6兲

where the brackets mean

具␧典⫽ 兺

L



兵 S i 其 i⫽1

␧e ␤ JS i S i⫹1 ⫹( ␤␮ /2)(S i )

2 ⫹( ␤␮ /2)(S

i⫹1 )

2

,

共7兲

L

H⫽⫺J

兺 S i S i⫹1 ,

i⫽1

共1兲

where J represents the coupling constant between nearest neighbor spins, and L is the linear size of the lattice. By imposing periodic boundary conditions, we have S L⫹1 ⫽S 1 . The constraint of constant number of zero spin components can easily be handled by employing the formalism of the grand-canonical ensemble. The grand-partition function is expressed by the equation L

⌶⫽



e ⫺ ␤␮ N Z,

N⫽0

共2兲

where ␤ ⫽1/k B T, and k B is the Boltzmann constant. ␮ is the L (S i ) 2 is the number of amphichemical potential, N⫽ 兺 i⫽1 philes, and Z is the canonical partition function for a fixed N. The grand-partition function can also be written in the form

and the ␦ S k ,0 are the delta’s Kronecker. Defining ␦ S k ,0⫽1 ⫺S 2k , and using the transfer matrix, we can write the equation P共 S 兲⫽

兺兿

兵 S i 其 i⫽1

R 共 S i ,S i⫹1 兲 ,

共3兲

␤ JS i S i⫹1 ⫹( ␤␮ /2)(S i ) 2 ⫹( ␤␮ /2)(S i⫹1 ) 2

where R(S i ,S i⫹1 )⫽e , are the elements of the 3⫻3 transfer matrix. Finally, in the thermodynamic limit, the grand-canonical potential density is

共8兲

⫻ 共 1⫺S 2 兲 U 兴 ,

where Tr means the trace of the matrix, U is the matrix that diagonalizes the transfer matrix R, 1 is the unity matrix and S is the matrix representation of the spin. This last equation can be used, after we take the thermodynamic limit L→⬁, to find the distribution of the aggregate size as a function of temperature and the chemical potential. Also using the formalism of the transfer matrix we obtain, in the thermodynamic limit, the density of amphiphiles, that is given by

␳⫽

L

⌶⫽

1 Tr关共 U ⫺1 RU 兲 L⫺n⫺1 U ⫺1 共 1⫺S 2 兲 R•••R ⌶

1 共 bc⫹ab 2 ⫺b⫺3a 兲共 2b 2 ⫹bc⫺b⫹a⫹ab 2 兲 , 共9兲 4 bc 共 ab 2 ⫺a⫺b⫹b 2 兲

where c⫽

冑冉



a a2 1⫺2 ⫺2ab⫹ 2 ⫹2a 2 ⫹a 2 b 2 ⫹8a . b b

III. MONTE CARLO SIMULATIONS

⌽⫽⫺ 共 ␤ 兲

⫺1

ln共 ␭ 兲 ,

共4兲

where ␭ is the largest eingenvalue of the transfer matrix that can be written as

We have performed Monte Carlo simulations in order to understand the underlying mechanism that leads the system towards its equilibrium state. We used linear sizes up to L ⫽104 and we have taken into account periodic boundary

8346

M. GIRARDI AND W. FIGUEIREDO

FIG. 1. Reduced chemical potential ␤␮ versus total concentration ␳ . This isotherm was obtained for ␤ J⫽2.5.

conditions. We also used different initial random configurations to guarantee that the final equilibrium state is the same. First of all, we define the concentration by fixing the number of zero spin component in the system. The dynamics employed here is a mixture of two different stochastic processes 关11兴: one of them is the exchange Kawasaki process, where two nearest neighbor spins are exchanged accordingly with the Metropolis prescription 关12兴. For this dynamical process the order parameter is conserved, which in our case is equivalent to keep constant the concentration of the solution 共 the number of zero spin component is fixed兲. The other dynamical process used here is the single spin-flip Glauber kinetics where a given ⫾1 spin is also flipped if it satisfies the Metropolis rule. This special Glauber mechanism changes the order parameter without changing the concentration. We follow the steps described below to find the equilibrium states of the system for given values of temperature, concentration and q, that gives the probability to perform the Kawasaki process in each step 关the probability of choose the Glauber process is (1⫺q)兴: we choose at random a site in the lattice, and a random number to select the process to apply. If the Kawasaki process is chosen, we select at random a nearest neighbor of the given site and we exchange them if they satisfies the requirements of Metropolis prescription. On the other hand, if the Glauber process is chosen, we only flip the given spin if it is nonzero and if it satisfies the Metropolis rule. To attain the equilibrium we need nearly 2⫻104 Monte Carlo steps 共MCs兲, where each MCs represents L random trials to select a spin in the lattice. After thermalization we used 105 MCs to obtain the mean value of the aggregate size distribution. IV. RESULTS AND CONCLUSIONS

We exhibit in Fig. 1 the plot of the isotherm of the reduced chemical potencial ( ␤␮ ) as a function of the number density of amphiphiles ( ␳ ). It is easy to see, that in the limit

PRE 62

FIG. 2. Free amphiphile concentration as a function of total concentration for ␤ J⫽0.5. The dashed line indicates the ideal gas behavior.

␮ →⬁ the lattice is saturated by amphiphiles, and this corresponds, in our case, to the usual S⫽1/2 Ising model. In the regime of very small densities, we have almost isolated amphiphiles, and the behavior is that of the ideal gas, where ( ␤␮ ) is a logarithmic function of the density. In Fig. 2, we show the density of isolated amphiphiles as a function of the total density for a fixed temperature. As to be expected, at very small densities, the slope of the curve is equal to 1. However, slightly increasing the total density, the slope of the curve in Fig. 2 decreases, and we can associate this fact with the appearance of small aggregates inside the solution. The plot of Fig. 2 resembles with the typical CMC curves observed in micellar systems. However, this is not sufficient to characterize these aggregates as being true micelles. It is well accepted in the literature of micellar systems 关5兴 that, with the exception of CMC curves, to characterize a true micellar aggregate, it is necessary the presence of a local minimum and a local maximum in the distribution curve of aggregates. We already used this property to study the behavior of two- and three-dimensional diluted systems as a function of temperature 关13,14兴. We have shown in these works that the parameter that controls the micellization process is the difference in height between the maximum and the minimum of the distribution curve. This difference in height vanishes linearly with temperature in two dimensions, while it vanishes quadratically in three dimensions. In both cases we defined a temperature at which the difference in height becomes zero, and it represents a transition from the micellar to a nonmicellar state. We present in Fig. 3 the plot we obtain for the distribution curve of aggregate size n for the same temperature and concentration in the range of Fig. 2. As we can see, the distribution curve is monotonically decreasing with the aggregate size. Then, there is no typical aggregate, and the system does not exhibit a true micellar behavior. The behavior observed in Fig. 3 is qualitatively the same up to densities of 20%. For higher values of density, the curve exhibits a single local maximum, but never shows

PRE 62

SPIN-1 AGGREGATION MODEL IN ONE DIMENSION

FIG. 3. Distribution function n ␳ (n) as a function of the size of the aggregates n for total concentration ␳ ⫽5.5% and ␤ J⫽2.5. The lines serve to guide the eyes.

up the corresponding local minimum, as we can see in Fig. 4, for the particular value of density ␳ ⫽50%. The general behavior observed in Figs. 3 and 4, can be obtained from Fig. 5, where we exhibit, for each aggregate size, the density of all molecules aggregated as a function of the total density. From this very simple one-dimensional model, we can conclude that it cannot support the existence of micelles. We also exhibit in Fig. 4 the distribution curve of aggregate sizes obtained through Monte Carlo simulations for q⫽1/2. For instance, in Fig. 6, choosing q⫽0, that is, only the Glauber dynamics is present, the distribution of aggregate sizes is exactly that random one that we generate at the beggining of the simulation. This is expected, because the Glauber process

FIG. 4. Distribution function n ␳ (n) as a function of the size of the aggregates n for total concentration ␳ ⫽50% and ␤ J⫽2.5. The line represents the exact solution and the small circles are the results of Monte Carlo simulation for q⫽1/2.

8347

FIG. 5. Distribution function n ␳ (n) as a function of the total concentration ␳ for different aggregate sizes, as indicated in the figure. We used ␤ J⫽2.5.

cannot change the size of the aggregates at any time. The only effect here is to change the value of order parameter, which is reflected only in the energy of the aggregates. On the other hand, for q⫽1, that is, using only the exchange Kawasaki dynamics, the simulation results are still different from the exact values. This happens because with this dynamics, which conserves the order parameter, only a finite region of the phase space is explored. Finally, for any other value of q, the simulation curve and the exact one coincide for all aggregate sizes. This clearly indicates that both dy-

FIG. 6. Distribution function n ␳ (n) as a function of the size of the aggregates n for ␤ J⫽2.5 and total concentration ␳ ⫽50%. The full line is the exact result and the connected circles represent Monte Carlo simulation results for q⫽1 共pure Kawasaki dynamics兲. The connected crosses give Monte Carlo results for q⫽0 共pure Glauber dynamics兲.

8348

M. GIRARDI AND W. FIGUEIREDO

FIG. 7. Free amphiphile concentration as a function of total concentration for ␤ J⫽2.5. The full line is the exact result and crosses represent the Monte Carlo results for q⫽1/2.

namical processes are essential to describe the evolution of the system towards the equilibrium states. Figure 7 displays the agreement between simulation and the exact results for the density of free amphiphiles as a function of total concentration. In conclusion, we have presented a simple one-

关1兴 C.J. Thompson, Mathematical Statistical Mechanics 共Princeton, New Jersey, 1972兲. 关2兴 C. Tanford, The Hydrophobic Effect: Formation of Micelles and Biological Membranes 共Wiley, New York, 1980兲. 关3兴 Y. Chevalier and T. Zemb, Rep. Prog. Phys. 53, 279 共1990兲. 关4兴 Micelles, Membranes, Microemulsions, and Monolayers, edited by W. Gelbart, A. Ben-Shaul and D. Roux 共Springer, New York, 1994兲. 关5兴 H. Wennerstrom and B. Lindman, Phys. Rep. 52, 1 共1979兲. 关6兴 J.R. Henderson, Phys. Rev. Lett. 77, 2316 共1996兲. 关7兴 D. Duque and P. Tarazona, J. Chem. Phys. 107, 10 207 共1997兲. 关8兴 A.B. Kolomeisky and B. Widom, Faraday Discuss. 112, 81 共1999兲. 关9兴 K. Kawasaki, in Phase Transitions and Critical Phenomena,

PRE 62

dimensional model of aggregation of amphiphiles to understand the process of micellization. The model was mapped onto a spin S⫽1 Ising model where the density of the solution was associated to the number of zero spin components. We were able to solve it exactly to obtain the densities of aggregate sizes as a function of total density, temperature, and the chemical potential. We also performed Monte Carlo simulations on this model and we demonstrated that the equilibrium states can only be attained through a combination of Glauber and Kawasaki dynamical processes. The model exhibits some characteristics similar to that of micellar systems, but the aggregate-size distribution curve does not show up the minimum and maximum that must be present in a micellized system. Although in Fig. 4 we observe a local maximum, we cannot say that this behavior caracterizes a true micellar system. The presence of the minimum in the distribution curve, below the free amphiphile concentration, is necessary. The disappearance of this minimum would lead to the coalescence of the two relaxation processes involved in the kinetics of micellization: the fast process, which accounts for the exchange of monomeric surfactants between micelles and solution, and the slow process, attributed to the micelle formation breakdown 关15兴. The model does not present any phase transition, and the transition from the micellar to nonmicellar state is absent. ACKNOWLEDGMENTS

This work was supported by the Brazilian agencies CNPq and FINEP.

关10兴 关11兴 关12兴 关13兴 关14兴 关15兴

edited by C. Domb and M.S. Green 共Academic, London, 1974兲, Vol. 4. R.J. Glauber, J. Math. Phys. 4, 294 共1963兲. ˜o, B.C.S. Grandi, and W. Figueiredo, Phys. Rev. E J.R.S. Lea 60, 5367 共1999兲. N. Metropolis, A.W. Rosenbluth, M.N. Rosenbluth, A.H. Teller, and E. Teller, J. Chem. Phys. 21, 1087 共1953兲. J.N.B. de Moraes and W. Figueiredo, J. Chem. Phys. 110, 2264 共1999兲. M. Girardi and W. Figueiredo, J. Chem. Phys. 112, 4833 共2000兲. E.A.G. Aniansson and S.N. Wall, J. Phys. Chem. 78, 1024 共1974兲.

using standard syste

zero component of spin represents the water molecules, while the remaining components (1) account for the amphiphilic molecules. We defined an ... centration of free amphiphiles, and it is different from zero. The local maximum in this curve, which ..... and the slow process, attributed to the micelle formation breakdown 15.

67KB Sizes 1 Downloads 142 Views

Recommend Documents

USING STANDARD SYSTE
directed sandpiles with local dynamical rules, independently on the specific ..... is needed to define a meaningful correlation length. In the latter case, on the ...

using standard syste
Mar 29, 2001 - *Electronic address: [email protected]. †Present address: Department of ... which implies that the dendrites act as a low-pass filter with cutoff frequency . ...... The most robust signatures of cortical modes are ...

using standard syste
May 19, 2000 - high-spin states, such as the deformed configuration mixing. DCM 4–7 calculations based on the angular momentum projection of the deformed ..... ration; Th.3 is MONSTER 28; Th.4 is the (f7/2)6 shell mode 27;. Th.5 is the rotational m

using standard syste
May 19, 2000 - 41. 372. aReference 25. bReference 26. cReference 27. TABLE IV. .... Sharpey-Schafer, and H. M. Sheppard, J. Phys. G 8, 101. 1982. 28 K. W. ...

using standard syste
In order to test this possibility, we have performed .... tency check of our results, we have checked that our expo- nents fulfill ... uncertainty in the last digit. Manna ...

using standard syste
One-particle inclusive CP asymmetries. Xavier Calmet. Ludwig-Maximilians-Universität, Sektion Physik, Theresienstraße 37, D-80333 München, Germany. Thomas Mannel and Ingo Schwarze. Institut für Theoretische Teilchenphysik, Universität Karlsruhe,

using standard syste
Dec 22, 2000 - ... one being the simplicity in definition and computation, another the fact that, for the ca- ...... search School, FOA Project No. E6022, Nonlinear ... the computer simulations were carried out on the Cray T3E at NSC, Linköping ...

using standard syste
May 1, 2000 - distance physics and Ta are the generators of color-SU3. The operators ... meson. Due to the different final states cu¯d and cc¯s, there are no.

using standard syste
rules: i each burning tree becomes an empty site; ii every ... the simple form 1 is meaningful. .... better to use analysis techniques that use the whole set of.

using standard syste
Jun 7, 2000 - VcbVcs*„b¯c V A c¯s V A b¯Tac V A c¯Tas V A…H.c.,. 5. PHYSICAL REVIEW D, VOLUME 62, 014027. 0556-2821/2000/621/0140275/$15.00.

using standard syste
4564. ©2000 The American Physical Society ..... (x,t) (x,t) 0, we can express all the functionals as ..... shifts i.e., in a log-log plot of a versus ) required for a.

using standard syste
Feb 20, 2001 - and the leaky integrate-and-fire neuron model 12. SR in a periodically ... where dW(t) is a standard Wiener process and I(t) is the deterministic ...

using standard syste
May 22, 2001 - 13 D. J. Watts, Small Worlds: The Dynamics of Networks Be- tween Order and Randomness Princeton University Press,. New Jersey, 1999. 14 A.-L. Barabási and R. Albert, Science 286, 509 1999; A.-L. Barabási, R. Albert, and H. Jeong, Phy

using standard pra s
Mar 20, 2001 - convex cloud to the desired state, by means of an external action such as a ..... 5 M. R. Matthews, B. P. Anderson, P. C. Haljan, D. S. Hall, C.

using standard prb s
at these low doping levels, and the effects due to electronic mistmach between Mn .... MZFC curves of Cr samples below TC is a signature of a well established ...

using standard pra s
Feb 15, 2001 - Electron collisions with the diatomic fluorine anion .... curves are Morse potential-energy curves obtained from experimental data as derived by ...

using standard prb s
Significant changes in the 3d electron population (with respect to the pure metal) are observed ... experimental arrangement as well as data analysis have been.

using standard prb s
Applied Physics Department, University of Santiago de Compostela, E-15706 Santiago de Compostela, ..... Values for the Curie constant, Curie-Weiss, and Cu-.

using standard pra s
Dec 10, 1999 - spectra, the data indicate that the detachment cross section deviates from the ... the detached electron is 1 for the Ir and Pt ions studied.

using standard prb s
Department of Physics, Santa Clara University, Santa Clara, California 95053. (Received 25 August ..... invaluable technical support of S. Tharaud. This work was funded by a Research Corporation Cottrell College Science. Grant, Santa Clara ...

using standard pra s
So the degree of mis- registration artifact associated with each pixel in a mix- ture of misregistered basis images can be measured as the smaller of the artifact's ...

using standard prb s
(10 15 s), and in an optical regime using lower penetration depth. 50 nm and ... time (10 17 s). ..... 9 Michael Tinkham, Introduction to Superconductivity, 2nd ed.

using standard prb s
Electron-spin-resonance line broadening around the magnetic phase ... scanning electron microscopy SEM. ... The magnetization values to fit the ESR data.

using standard prb s
Mar 6, 2001 - material spontaneously decomposes into an electronically spatially ... signed dilution refrigerator and in a pumped 4He cryostat. The films were ...