9037 PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Kinetic and thermodynamic control of protein adsorption

J. Satulovsky, M. A. Carignano, and I. Szleifer


PNAS 2000;97;9037-9041; originally published online Jul 25, 2000;
doi:10.1073/pnas.150236197
This information is current as of February 2007.

Online Information High-resolution figures, a citation map, links to PubMed and Google Scholar,
& Services etc., can be found at:
www.pnas.org/cgi/content/full/97/16/9037
This article has been cited by other articles:
www.pnas.org/cgi/content/full/97/16/9037#otherarticles
E-mail Alerts Receive free email alerts when new articles cite this article - sign up in the box
at the top right corner of the article or click here.
Rights & Permissions To reproduce this article in part (figures, tables) or in entirety, see:
www.pnas.org/misc/rightperm.shtml
Reprints To order reprints, see:
www.pnas.org/misc/reprints.shtml

Notes:
Kinetic and thermodynamic control of
protein adsorption
J. Satulovsky, M. A. Carignano, and I. Szleifer*
Department of Chemistry, Purdue University, West Lafayette, IN 47907

Communicated by Stuart A. Rice, University of Chicago, Chicago, IL, May 22, 2000 (received for review October 22, 1999)

Control of nonspecific protein adsorption is very important for the surface induces a gradient in the chemical potential of the
design of biocompatible and biomimetic materials as well as drug protein. This gradient arises from the sudden inhomogeneous
carriers. Grafted polymer layers can be used to prevent protein environment that appears in the direction perpendicular to the
adsorption. We have studied the molecular factors that determine surface. The proteins in the close vicinity of the surface feel the
the equilibrium and kinetic control of protein adsorption by bare attractive interaction of the surface as well as the repulsions
grafted polymer layers. We find that polymers that are not at- induced by the polymer molecules. The balance between these
tracted to the surface are very effective for kinetic control but not interactions will result in a final equilibrium density profile and
very good for equilibrium reduction of protein adsorption. Poly- amount of protein adsorbed.
mers with attractions to the surface show exactly the opposite The time evolution from the homogeneous system to the new
behavior. The implications for molecular design of biocompatible equilibrium induced by the presence of the modified surface can
materials also are discussed in this paper. be described (in the free-draining limit) with the help of a
diffusion equation (15–18) of the form

P 冋 册
rotein adsorption plays a major role in a variety of important
⭸ ␳ pro 共z;t兲 ⭸ ⭸
biological-related processes. Biocompatible materials are ⫽ D pro ␳ 共z;t兲 ␮ pro 共z;t兲 , [1]
required to eliminate, or largely reduce, the adsorption of blood ⭸t ⭸z pro ⭸z
proteins, for example, to avoid surface-induced thrombosis (1,
2). In recent years special attention has been given to the where we have assumed that the diffusion in the xy plane is faster
modification of protein–surface interactions using grafted poly- than the adsorption. This assumption is justified for the relatively
mers (3, 4, 5, 6). The idea, borrowed from many years of research low surface densities that we treat here (19). The time-
dependent local chemical potential is defined by ␮pro(z;t) ⫽
in colloidal stabilization (7) and mimicking one of the roles of
polysaccharides in cell membranes (8), is to build a steric barrier ␦W
, where W is the free-energy density of the system. This
to the proteins by the presence of the polymer layer (9, 10). In ␦␳pro(z;t)
practice, however, there is still no systematic way of modifying free energy includes the contribution of the grafted polymer
surfaces that are successful in preventing protein adsorption. molecules, the solvent, and the surface, under the condition of
This is because of the complexity involved in the process caused the fixed protein concentration profile given by ␳pro(z;t) at each
by the interplay between changing average structure of the time t.
polymer–protein layer, strong protein–surface attractions, and The adsorption kinetics can be described with the help of Eq.
polymer–surface interactions (11). For example, liposomes† 1, assuming that the time scale of the diffusion of the proteins is
formed by mixtures of lipid and lipid-polymer have shown much slower than that of the polymer and solvent rearrange-
increased longevity in the bloodstream and are now being used ment. This is a physically motivated approximation, because the
as drug delivery systems (12, 13). It is believed that the role of solvent molecules (water) are much faster than the protein
the polymers tethered to the liposome interface is to present a motion, and the rearrangements of the polymer molecules
steric barrier to approaching proteins from the blood. However, involve only local moves, which, for the flexible chains of interest
the amount of polymer on the surface is rather small, and here, are much faster than the typical motion of the larger
experimental observations of the same polymer grafted on protein molecules.
hydrophobic surfaces, at the same surface coverage, show that The free energy of the system is determined by using a
although the amount of protein adsorbed is lower than in the molecular mean-field theory. The predictions of the theory for
absence of polymer, there is still a significant amount of protein equilibrium systems have been shown to be in excellent agree-
adsorption (14). ment with experimental observations for conformational and
In this paper we demonstrate that the molecular factors that thermodynamic properties of tethered polymer layers (20, 21) as
determine the ability of the polymer layer to prevent protein well as for the adsorption of proteins on hydrophobic surfaces
adsorption are different for the kinetic control as compared with (11, 14). The basic idea of the theory is to treat each molecule
the equilibrium case. We will demonstrate that our theoretical with all of its intramolecular and surface interactions exactly
predictions can quantitatively reproduce experimental observa- taken into account (within the chosen model system). The
tions of protein adsorption isotherms. We will show that sur- intermolecular interactions are considered within a mean-field
face–polymer interactions play a central role in the ability of the approximation. The mean field is determined by the average
polymer layer to prevent protein adsorption. However, the role
of those interactions is different for kinetic as compared with Abbreviations: pdf, probability distribution function; EO, ethylene oxide; PEO, polyethyl-
equilibrium control. Our findings enable us to rationalize why ene oxide.
the surface-modified liposomes have an increased longevity but *To whom reprint requests should be addressed. E-mail: igal@purdue.edu.
polymers on hydrophobic surfaces are not as effective to prevent
†Liposomes
protein adsorption. Finally, our theoretical predictions can be are closed spherical bilayers formed by lipid molecules with varying radii from
BIOPHYSICS

10 nm to micrometers.
used to decide the type of surface modification necessary for the
The publication costs of this article were defrayed in part by page charge payment. This
rational design of biomaterials. article must therefore be hereby marked “advertisement” in accordance with 18 U.S.C.
Consider a protein solution of density ␳b that is put into §1734 solely to indicate this fact.
contact with a surface. The surface has polymers that are Article published online before print: Proc. Natl. Acad. Sci. USA, 10.1073兾pnas.150236197.
chemically attached to it at one of their ends. The presence of the Article and publication date are at www.pnas.org兾cgi兾doi兾10.1073兾pnas.150236197

PNAS 兩 August 1, 2000 兩 vol. 97 兩 no. 16 兩 9037–9041


properties of all of the molecules in the mixture, and thus the
theory is in essence a self-consistent approach.
We write the free energy of the system in complete analogy
with the equilibrium approach (22) but with all of the quantities
now assumed to be time-dependent. The explicit expression for
the time-dependent free energy density (per unit area) reads

␤W ⫽ ␴ 冉冘␣
P共 ␣ ;t兲lnP共 ␣ ;t兲 ⫹ ln␴ ⫹ 冊 冕 ␾ s 共z;t兲ln␾ s 共z;t兲dz ⫹

冕 ␳ pro 共z;t兲关ln␳ pro 共z;t兲 ⫹ ␤ U ps 共z兲兴dz, [2]

where the first two terms are the contribution from the tethered
polymers, with the first being the conformational entropy of the
chains, with P(␣;t) being the time-dependent probability distri-
bution function (pdf) of chain conformations and the second
being the tethered polymers two-dimensional translational en-
tropy. The third term is the solvent entropy, with ␾s(z;t) being the
volume fraction of solvent at distance z from the surface at time Fig. 1. Theoretical predictions for the amount of protein adsorbed as a
t. The last two terms correspond to the translational entropy of function of the surface coverage of EO oligomers (full lines). The experimental
the proteins and the protein–surface interaction, respectively, observations (symbols) are from Prime and Whitesides (24, 25), where they
measured the thickness of the adsorbed film (in nm), which is proportional to
Ups(z) is the bare surface–protein interaction. This interaction is
the surface density of protein, as a function of composition of a mixed
taken from ref. 23 where the potential between lysozyme and a self-assembled film. The mixed films are composed by SC10OH and SC11(EO)n,
hydrophobic surface was calculated by using atomistic potentials. with n ⫽ 6 and 17, on gold surfaces. The relationship between the experi-
Note that the solvent and protein contributions are integrated mentally observed thickness and the protein density was fitted so that at zero
over z to account for all of the molecules. composition the predictions and the experimental observations agree. The
Eq. 2 assumes that the intermolecular attractions between all parameters used in the calculations are from refs. 11 and 22.
of the components in the system are the same. This is actually the
model that we successfully used to compare experimental ob-
where q(t) is the time-dependent normalization constant. For
servations and the predictions of the theory in early work (14)
the solvent-density profile, the expression is
and in the comparisons shown below. The excellent agreement
between the predictions of the theory and the experimental ␾ s 共z;t兲 ⫽ exp[⫺␤␲(z;t)v 0 ]. [5]
observations supports the assumption.
Eq. 2 does not include the repulsive interactions between the The numerical values of the Lagrange multipliers are obtained
molecules. The hard-core repulsions are accounted for by pack- by replacing the expression for the pdf, Eq. 4, and the solvent-
ing constraints in which the available volume to the molecules, density profile, Eq. 5, into the constraint equation, Eq. 3, where
at each distance z from the surface, is filled by polymer, protein, the protein contribution is given as an input obtained from the
or solvent molecules. This constraint reads diffusion equation, Eq. 1. Once the Lagrange multipliers are
obtained, one can calculate the time-dependent protein chemical

␴具ng共z;t兲典v0 ⫹ 冕 ␳pro共z⬘;t兲v共z;z⬘兲dz⬘ ⫹ ␾s共z;t兲 ⫽ 1, for all z, [3]


potential that is needed for the time evolution of the adsorption.
This calculation is done by differentiating the free-energy den-
sity to obtain
where the first term is the volume fraction of polymer at z, with
具ng(z;t)典dz ⫽ ⌺␣P(␣;t)ng(z,␣;t)dz being the average number of
polymer segments of volume v0 at time t at distance z from the
␤␮ pro 共z;t兲 ⫽ ln␳ pro 共z;t兲 ⫹ ␤ U ps 共z兲 ⫺ 冕 ␤␲ 共z⬘;t兲v共z⬘;z兲dz⬘,

[6]
surface. The second term is the volume fraction of proteins with
v(z;z⬘)dz being the volume that the proteins at distance z⬘ from which is readily obtained once the Lagrange multipliers are
the surface contribute to z. The last term is the volume fraction known. Thus, the time evolution of the system is obtained in the
of solvent. following way. At time t ⫽ 0, a tethered polymer layer in
The assumption of solvent and polymer equilibration in the equilibrium with pure solvent is brought into contact with a
time scale of protein motion implies that the time-dependent pdf protein solution. The initial density profile of proteins enables
of tethered polymer configurations and the solvent-density the calculation of ␲(z;0) from which ␮pro(z;0) is obtained by use
profile are those that minimize the free energy (Eq. 2) subject of Eq. 6. From it, the diffusion equation (Eq. 1) is iterated a time
to the packing constraint (Eq. 3) for fixed protein density profile. step. The new density profile of proteins now is given as input for
Thus, at each time t the density of proteins calculated from the the calculation of the new Lagrange multipliers from which the
diffusion equation (Eq. 1) is given as input to the free energy and chemical potential profile is obtained and the time evolution is
this free energy is minimized subject to the packing constraints. further iterated. This procedure is continued until the new
This minimization is done by introducing time-dependent La- equilibrium condition is achieved in which ␮pro(z;t) ⫽ ␮pro.eq for
grange multipliers, ␤␲(z;t). The pdf of chain conformations is all z.
Eq. 6 shows that the motion of the proteins toward the surface

P共 ␣ ;t兲 ⫽
1
q共t兲
exp ⫺ 冋冕 册
␤␲(z;t)v 0 n g 共z, ␣ ;t兲dz , [4]
contains three contributions. The first one is the purely diffusive
motion that is found in any ideal system because of the gradient
in composition. The second arises from the force exerted to the

9038 兩 www.pnas.org Satulovsky et al.


Fig. 2. (A) Amount of proteins adsorbed, in molecules per Å2, as a function of time. The dashed lines correspond to the equilibrium amount of protein adsorbed.
The lines correspond to the following number of EO units: black, no polymer; red, n ⫽6; green, n ⫽15; blue, n ⫽ 25; and orange and magenta, n ⫽ 50. The magenta
curve corresponds to a hydrophobic surface. (B) The potential of mean force, in kJ兾mol, as a function of the distance from the surface for: black, bare
protein–surface interaction; and red, n ⫽ 50 before the beginning of the adsorption. The green and blue curves are for n ⫽ 50 at the times marked by the arrows
in A. (C) Maximal steric repulsion for n ⫽ 50 as a function of time. Note the changes in the shape and strength of the potential as the adsorption process takes
place. All of the results correspond to a grafted density of ␴ ⫽ 0.0012Å⫺2. Higher surface coverage of polymer will result in slower kinetics but qualitatively similar
behavior.

proteins by the surface. These two terms do not depend on the phobic glass surfaces was found to be closely linked to the fact
intermolecular interactions and will lead to a fast approach of that the EO segments are strongly attracted to the surface and
the proteins to the surface. We refer to them as the purely compete with the proteins for adsorption sites (14).
diffusive motion (see e.g., Fig. 2 and discussion thereafter). The How good is the equilibrium assumption? To this end, we solve
last term is the one arising from the intermolecular (protein– the diffusion Eq. 1 with the free energy obtained from the same
polymer and protein–protein) interactions. This term will be theoretical approach that was used in the equilibrium predictions
responsible for steric barriers and it is expected to lead to of Fig. 1. Fig. 2A shows the time evolution of the amount of
regimes in which the motion is dominated by kinetic barriers. protein adsorbed on the surface for a variety of PEO chain
The interplay between the diffusion and barrier crossing to lengths. All of the profiles but one correspond to PEOs that are
determine the total adsorption process is discussed in detail not attracted to the surface; we discuss the case of the attractive
below. surface later. There are several interesting features that can be
Fig. 1 shows the predicted equilibrium adsorption isotherms observed from the profiles. First, as the chain length of the
for lysozyme and fibrinogen on surfaces with short oligomeric grafted polymers increases up to 25 segments (corresponding
polyethylene oxide (PEO) together with the experimental ob- close to PEO-1,000), the equilibrium amount of protein ad-
servations of Prime and Whitesides (24, 25). The calculations sorbed decreases. However, for PEO-2,000, the amount of
were carried out by full minimization of the free energy (ther- protein adsorbed at equilibrium is almost identical to that of
modynamic equilibrium), i.e., with respect to the protein degrees PEO-1,000. The equilibrium amount of protein adsorbed is
of freedom as well, as explained in detail in refs 14 and 22. The independent of polymer chain length once the thickness of the
agreement between the theory and the experimental observa- layer is larger than the protein size (22).
tions is excellent. The adsorption isotherms show that as the Second, the kinetics of protein adsorption is qualitatively
surface coverage of polymer increases, the amount of protein similar for all chain lengths up to PEO-1,000. There are two
adsorbed decreases. There is a surface coverage of oligomers qualitatively different regimes. For short times, the adsorption is
above which there is zero protein adsorption. The role of the very fast, and it is controlled by the diffusion of the proteins to
short chains is basically to fill the space close to the surface, the surface, i.e., the first two terms of Eq. 6. The second regime
reducing the amount of available surface for the adsorbing is a much slower one, and it is controlled by the repulsive barrier
proteins (11, 22, 24, 25). The most important point to note is that that the proteins already adsorbed on the surface and the grafted
these chains are short and that the surface coverage at which the polymers present to the approaching proteins, i.e., the motion is
adsorption goes to zero is very large. Those surface coverages are dominated by the last term in Eq. 6. The total time to reach
an order of magnitude larger than are typical surface coverages equilibrium in these cases supports the equilibrium assumption
obtained by grafting PEO-2,000 or PEO-5,000‡. Further, the used to obtain Fig. 1.
calculations assume that the ethylene oxide (EO) oligomers are Third, for PEO-2,000, there is no fast regime, i.e., there is no
not attracted to the surface. In contrast, the ability of long- diffusion-controlled adsorption. At this molecular weight and for
grafted PEO to reduce protein adsorption on modified hydro- the surface coverage shown, the polymer chains present a steric
BIOPHYSICS

barrier that causes the adsorption to be controlled by the kinetics


of barrier crossing, Fig. 2B.
‡PEO-M, also called PEG-M for polyethylene glycol, is a chain of molecular weight M. Thus,

the number of segments in the chain is given by M兾45. We refer here to the commonly used
The dependence of the time scales for adsorption on polymer
experimental molecular weights. The exact number of segments used in the calculations molecular weight also can explain the success of pegolated
is given in the figure legends. liposomes having increased longevity in the bloodstream. Typ-

Satulovsky et al. PNAS 兩 August 1, 2000 兩 vol. 97 兩 no. 16 兩 9039


Fig. 3. (Left) Schematic representation of the ability of grafted polymers to prevent protein adsorption. For polymers attracted to the surface, there is no kinetic
barrier but the protein competes with the polymer for adsorption sites. Polymers that are not attracted to the surface present a large steric barrier but not very
good thermodynamic prevention because of the ability of the protein to deform the polymer layer. (Right) The curves on the right show the calculated average
shapes of two neighboring polymers as a function of time. The shape is defined here (20) by the lateral average radius of gyration of the chains as a function

ical molecular weights used in those systems are PEO-2,000 to polymer molecular weights below PEO-2,000. For the surface
PEO-5,000. As we can see from Fig. 2 for PEO-2,000, the chains coverage considered here, we see that in all of the cases there is
are long enough that the adsorption is completely controlled by a diffusion-controlled regime of the adsorption kinetics. This
the kinetic barriers presented by the polymers, and the time scale finding implies that the proteins approaching the surface feel a
to achieve equilibrium is much longer. Actually, the system only purely attractive interaction potential with no barriers. The
reaches equilibrium in many hours. Moreover, the calculations dramatic change in slope in the kinetics is attributable to
presented here are for lysozyme, which is a relatively small the presence of a barrier in the potential of mean force that the
protein as compared to fibrinogen. Thus, for fibrinogen the time proteins approaching the surface feel¶. For the hydrophobic
scale for adsorption is going to be much longer, because its large surfaces, the same effect is observed but up to much longer
size results in very large repulsions caused by the interactions molecular weights (see above). For the PEO-2,000 with no
with the polymers before reaching the surface§. attraction with the surface, the polymers present a kinetic barrier
The time-dependent adsorption for the case of PEO-2,000 even in the absence of any adsorbed proteins (Fig. 2 B and C).
with attractive interactions with the surface (hydrophobic sur- The presence of the barrier implies that there is no diffusion-
faces) shows two important differences with the nonattractive controlled regime. Furthermore, any molecular weight above
case. First, the equilibrium amount of adsorbed proteins is much 2,000 will show a stronger barrier and thus the time scale of the
lower for the hydrophobic surface as compared with the same initial adsorption will be even longer.
molecular weight for nonattractive surfaces. Namely, grafted The main conclusion from this work is that the best surface
polymers that are attracted to the surface are more effective in modifiers for the kinetic control of protein adsorption are
reducing protein adsorption. Second, the time scale for adsorp- polymers that are not attracted to the adsorbing surface. How-
tion, both for the initial adsorption and to reach equilibrium, are ever, the best type of polymers for the thermodynamic control
orders of magnitude faster for surfaces with attractive interac- of protein adsorption are those that are attracted to the surface.
tions with the polymers. The reason for the faster kinetics is that The molecular picture that emerges is summarized in Fig. 3,
the polymers have almost all of their segments close to the where a cartoon is shown of how the different types of polymers
surface and thus present no steric barrier to the approaching kinetically or thermodynamically prevent the protein adsorption,
proteins up to the point of close vicinity with the surface. The along with the calculated variation of the shape of the grafted
reason for being better at preventing the equilibrium adsorption polymers as a function of time. The complexity involved in the
is that the proteins need to compete for adsorption sites with the adsorption process can be seen in the time-dependent average
polymers. The short time scale for adsorption on hydrophobic shape of the polymer molecules. The kinetics of protein adsorp-
surfaces suggests that the recent comparisons (14) with exper- tion on surfaces with grafted polymers involves the competition
imental observations assuming equilibrium are probably very good. between the strong bare attractive interactions between the
We now discuss in some detail the origin and characteristics surface and the proteins and the conformational entropy loss
of the repulsive barriers that the proteins feel in the different that the grafted polymers have to pay to accommodate the
cases. First, let us consider the case of no grafted polymer or proteins. This process involves a constantly varying environment,

§Thebare potential of fibrinogen with surfaces is not known and therefore it is very hard ¶The potential of mean force here refers to the interaction (as a function of the distance
to provide a reliable estimate for the potential of mean force. The calculations presented from the surface) of a protein with the surface averaged over all of the configurations of
here use the same attractive contribution between the surface and fibrinogen that was the polymer molecules, solvent, and other proteins in the system, which is given exactly by
proposed in ref. 14. This interaction is only the attractive strength at contact. the last term in Eq. 6.
and therefore the kinetic behavior cannot be characterized in is necessary for protein rejection in each of these cases is
simple terms. We find that the kinetic barriers for protein different. In systems where it is necessary to control protein
adsorption vary constantly in time, and the whole potential is a adsorption over finite time scales, it is probably best to use
complicated function that varies with distance and time as a relatively dense polymer layers with long polymers that are not
result of the constant rearrangement of the polymer–protein attracted to the surface of the biomaterial. This scenario should
layer (see Fig. 2 B and C). provide the best kinetic control. On the other hand, for materials
We have presented a detailed study of the equilibrium and that need to be in contact with the bloodstream for years, the
kinetic adsorption of proteins on surfaces with grafted polymers. optimal type of polymers may be one that is attracted to the
The extension of the molecular theory for tethered polymer surfaces, then providing thermodynamic control. It is important
layers to dynamical systems enables the study of kinetic processes
to mention that very high surface coverage of grafted polymers
over several orders of magnitude in time without compromising
is extremely hard to obtain experimentally (14, 26). Thus,
molecular detail. Note that although this approach does not
provide the detail of molecular dynamics simulations, it enables alternative approaches may be needed, for example, mixtures of
the study of the dynamic processes over 15 orders of magnitude polymers that present optimal kinetic and thermodynamic con-
in time, thus bridging the gap between microseconds and days in trol under experimental realizable conditions. The theory used
complex systems with the ability to follow changes on average here provides quantitative guidelines on how to build those
molecular properties. optimal layers.
Biocompatible materials can be in contact with the blood-
stream for relatively short times (e.g., drug carriers) or very long This work is supported by the National Science Foundation through
times (e.g., artificial organs). The type of surface modifier that Grant CTS-9624268. I.S. is a Camille Dreyfus Teacher兾Scholar.

1. Horbett, T. A. (1993) Cardiovasc. Pathol. 2, 137S–148S. 11. Szleifer, I. (1997) Curr. Opin. Solid State Mater. Sci. 2, 337–344.
2. Salzman, E. W., Merrill, E. W. & Hent, K. C. (1994) in Hemostasis and 12. Lasic, D. & Martin, F., eds. (1995) Stealth Liposomes (CRC, Boca Raton, FL).
Thrombosis, eds. Colman, R. W., Hirsh, J., Marder V. J. & Salzman, E. W. 13. Fisher, L. & Lasic, D. (1998) Curr. Opin. Colloid Interface Sci. 3, 509–510.
(Lippincott, Philadelphia), pp. 1469–1485. 14. McPherson, T., Kidane, A., Szleifer, I. & Park, K. (1998) Langmuir 14, 176–186.
3. Milton Harris, J. & Zalipsky S., eds. (1997) Poly(ethylene glycol): Chemistry and 15. Fraaije, J. G. E. M. (1993) J. Chem. Phys. 99, 9202–9212.
Biological Applications (Am. Chem. Soc., Washington, DC). 16. Hasegawa, R. & Doi, M. (1997) Macromolecules 30, 3086–3089.
4. Andrade, J. D., ed. (1985) Surface and Interfacial Aspects Of Biomedical 17. Kondepudi, D. & Prigogine, I. (1998) Modern Thermodynamics: From Heat
Polymers (Plenum, New York). Engines to Dissipative Structures (Wiley, New York).
5. Torchilin, V. P. & Trubetskoy, V. S. (1995) Adv. Drug Delivery Rev. 16, 141–155.
18. Marconi, U. M. B. & Tarazona, P. (1999) J. Chem. Phys. 110, 8032–8044.
6. Needham, D., McIntosh, T. J., Simon, S. A. & Zhelev, D. (1998) Curr. Opin.
19. Ravichandran, S. & Talbot, J. (2000) Biophys. J. 78, 110–120.
Colloid Interface Sci. 3, 511–517.
7. Napper, D. H. (1983) Polymeric Stabilization of Colloidal Dispersions (Aca- 20. Szleifer, I. & Carignano, M. A. (1996) Adv. Chem. Phys. 94, 165–260.
demic, New York). 21. Szleifer, I. (1996) Curr. Opin. Colloid Interface Sci. 1, 416–423.
8. Holland, N. B., Yongxing, Q., Ruegsegger, M. & Marchant, R. E. (1998) Nature 22. Szleifer, I. (1997) Biophys. J. 72, 595–612.
(London) 392, 799–801. 23. Lee, S. J. & Park, K. (1994) J. Vac. Sci. Technol. 12, 1–7.
9. Ikada, Y. (1984) Adv. Polym. Sci. 57, 104–140. 24. Prime, K. L. & Whitesides, G. M. (1991) Science 252, 1164–1167.
10. Jeon, S. I., Lee, J. H., Andrade, J. D. & de Gennes, P. G. (1991) J. Colloid 25. Prime, K. L. & Whitesides, G. M. (1993) J. Am. Chem. Soc. 115, 10714–10721.
Interface Sci. 142, 149–158. 26. Halperin, A., Tirrell, M. & Lodge, T. P. (1992) Adv. Pol. Sci. 100, 31–71.

BIOPHYSICS

Satulovsky et al. PNAS 兩 August 1, 2000 兩 vol. 97 兩 no. 16 兩 9041

You might also like