You are on page 1of 33

Annu. Rev. Phys. Chem. 2001. 52:10737 Copyright c 2001 by Annual Reviews.

All rights reserved

MECHANISMS AND KINETICS OF SELF-ASSEMBLED MONOLAYER FORMATION


Daniel K Schwartz
Annu. Rev. Phys. Chem. 2001.52:107-137. Downloaded from www.annualreviews.org by Universite Toulouse 3 -Paul Sabatier on 09/16/10. For personal use only.

Department of Chemical Engineering, University of Colorado, Boulder, Colorado 80309; e-mail: daniel.schwartz@colorado.edu

Key Words thin lm, coatings, SAM, monolayer growth s Abstract Recent applications of various in situ techniques have dramatically improved our understanding of the self-organization process of adsorbed molecular monolayers on solid surfaces. The process involves several steps, starting with bulk solution transport and surface adsorption and continuing with the two-dimensional organization on the substrate of interest. This later process can involve passage through one or more intermediate surface phases and can often be described using two-dimensional nucleation and growth models. A rich picture has emerged that combines elements of surfactant adsorption at interfaces and epitaxial growth with the additional complication of long-chain molecules with many degrees of freedom.

INTRODUCTION
The adsorption of amphiphilic surfactant molecules at interfaces is a well-known phenomenon that is at the heart of all detergency applications. A single molecular layer (monolayer) of surfactant stabilizes oil droplets and gas bubbles in an aqueous environment, enhancing the stability of emulsions and foams. In addition to adsorption at liquid-liquid and liquid-vapor interfaces, amphiphilic molecules also adsorb at the solid-liquid interface. Self-assembled monolayers (SAMs) are distinguished from ordinary surfactant monolayers by the fact that one end of the molecule (generally the hydrophilic one) is designed to have a favorable and specic interaction with the solid surface of interest (the substrate). This results in the formation of a stable monolayer lm that remains intact even after the substrate is removed from solution. Due to the specic interaction between molecule and substrate, the adsorption can often be carried out in a variety of solvents, polar and nonpolar, allowing greater exibility in molecular design and, therefore, in the types of surface properties that can be modied and controlled. Since the monolayer lms are thin and homogeneous, they have found frequent use as model surfaces in research applications. Much of the interest in these lms lies in their potential as inexpensive and versatile surface coatings for applications including control of wetting and
0066-426X/01/0601-0107$14.00

107

108

SCHWARTZ

Annu. Rev. Phys. Chem. 2001.52:107-137. Downloaded from www.annualreviews.org by Universite Toulouse 3 -Paul Sabatier on 09/16/10. For personal use only.

adhesion, chemical resistance, biocompatibility, sensitization for photon harvesting, molecular recognition for sensor applications, and many others. Zisman is often credited with originating the SAM concept in his 1946 paper (1). Work in the early 1980s by Nuzzo & Allara (thiols on gold) (2) and Maoz & Sagiv (trichlorosilanes on silicon oxide) (3) introduced what were to become the two most popular SAM systems and brought SAMs into the popular scientic consciousness. Interest in these monolayer lms has continuously increased since that time, and the development and application of surface-sensitive experimental techniques (e.g. scanning probe microscopy, vibrational spectroscopy, and synchrotron X-ray sources) has resulted in an improved understanding of the lm structure and growth process. Poirier recently reviewed scanning tunneling microscopy (STM) measurements of thiol-based SAMs (4), and a more general review of SAM structure (with some information about lm growth) was previously published by Ulman (5). Ulmans book (6) serves as a useful introduction to SAMs and thin organic lms in general. The current review is more narrowly focused on the growth process of a variety of SAM systems.

THE BIG PICTURE: General Growth Mechanisms Bulk Transport and Adsorption
Many processes are involved in SAM growth. A rst step is clearly the solutionphase transport of adsorbate molecules to the solid-liquid interface, which can involve some combination of diffusive and convective transport. This is followed by adsorption on the substrate with some adsorption rate (related to a sticking probability). The overall adsorption dynamics may be diffusion-controlled, adsorptionrate controlled, or in an intermediate mixed-kinetic regime. This part of the selfassembly process is closely related to the adsorption of surface-active molecules at the liquid-vapor interface, an area that has been thoroughly studied. Although the typical quantity of interest at the liquid-vapor interface is surface tension rather than surface concentration (or coverage), the two quantities are related by the surface equation of state. In fact, most dynamic adsorption models are actually written in terms of surface concentration and translated into dynamic surface tension predictions, using an equation of state determined by applying the Gibbs equation (7) to equilibrium surface tension data. The dynamics of surfactant adsorption were thoroughly reviewed by Chang & Franses (8), and most of the mathematical development presented by them is directly relevant to the initial adsorption stage of SAM formation. Quantitative aspects of this process are discussed later in this review.

Self-Organization on the Surface


Two-dimensional (2D) molecular organization is a key ingredient for SAM stability and function. This differs from surfactant monolayers at the air-water interface,

SELF-ASSEMBLED MONOLAYER FORMATION

109

Annu. Rev. Phys. Chem. 2001.52:107-137. Downloaded from www.annualreviews.org by Universite Toulouse 3 -Paul Sabatier on 09/16/10. For personal use only.

for example, because their primary function is simply to reduce surface tension. In SAM formation, therefore, there must be an evolution of the molecular order as adsorption progresses and the surface coverage increases. For example, the very early stages of adsorption can be pictured as isolated adsorbed molecules, conformationally disordered and randomly distributed on the substrate. The nal lm involves close-packed adsorbate molecules with relatively uniform molecular orientation and conformation. Although one might imagine a continuous path from the former structure to the latter, experimental evidence points to a stepwise process that can be thought of as an isothermal path through a quasiequilibrium 2D-phase diagram like the one schematically illustrated in Figure 1. Possible states alluded to in this phase diagram include (a) a low-density vapor phase in which isolated, mobile adsorbate molecules are randomly deposited on the surface, (b) an intermediate-density phase that could involve conformationally disordered molecules or ones lying at on the surface, and (c) a nal, high-density solid phase in which the molecules are conformationally ordered, close packed, and

Figure 1 Schematic quasi-equilibrium 2D-phase diagram for a generic SAM system. The dotted lines represent hypothetical isothermal paths of SAM growth at temperatures below (T1) and above (T2), the triple point (Ttriple).

110

SCHWARTZ

standing approximately normal to the surface plane with a possible polar tilt angle of about 30 . As discussed below, other states are, of course, possible. In a hypothetical situation in which the adsorption rate is much slower than any other process, the monolayer system would follow the equilibrium phase diagram. There are two qualitatively different growth processes suggested by the lines at temperatures T1 and T2 in Figure 1. If the temperature is lower than the triple point (e.g. temperature T1), the growth sequence will be similar to the one shown in Figure 2a. Initially, adsorbed molecules will form a dilute 2D-vapor phase. At a relatively low surface concentration, the monolayer will enter a coexistence region between the vapor and the high-density condensed (solid) phase. Domains (islands) of solid phase will nucleate and grow, surrounded by isolated adsorbate molecules in the vapor phase. Eventually, these domains will grow to cover the entire substrate. This mechanism is analogous to the three dimensional (3D) process of crystal nucleation and growth from a vapor phase precursor, and the 2D scenario is typical for epitaxial lm growth from the vapor phase (e.g. molecular-beam epitaxy) (9). At a temperature above the triple point (e.g. T2 in Figure 1), a more complicated progression will occur as illustrated in Figure 2b. When the vapor phase reaches a certain surface concentration, islands of an intermediate, lowdensity condensed phase will nucleate and grow. This phase may be a disordered 2D-liquid phase or an ordered phase with lower density than the solid phase (e.g. a lying-down phase where the molecular axis is parallel to the surface plane). Eventually the vapor phase is completely converted to the low-density condensed phase. As adsorption continues, a second transition occurs involving nucleation, growth, coalescence, etc, of solid-phase islands surrounded by the low-density condensed phase. Note that, at any temperature, a snapshot of an incomplete lm during growth will often involve islands of one phase surrounded by another, in particular, islands of solid phase surrounded by either liquid or vapor phase. It is important to recognize that the picture painted in the previous paragraph is somewhat oversimplied. For example, the adsorption rate will not always be much slower than other surface processes, and, therefore, partial monolayers may be quite far from equilibrium. If the nucleation and growth of condensed-phase domains do not keep up with the deposition rate, the less condensed phase will become super concentrated (i.e. it will have a density greater than the equilibrium coexistence concentration), and, thus, its density may vary considerably during the growth of the more condensed phase. This behavior is well known in vapor phase thin-lm deposition, where the surface concentration of free adsorbate atoms is understood to vary during island nucleation and growth, and is likely to occur during SAM growth as well. However, the surface concentration in the vapor phase will always be small and amount to a negligible fraction of the molecules on the surface. In the case of a 2D-liquid phase, however, the surface density is not negligible, and, in fact, the lm thickness is directly related to the surface concentration. Therefore, variation of the surface density in a 2D liquid in coexistence with solid-phase islands will have a signicant effect on the appearance of the partial-monolayer lm.

Annu. Rev. Phys. Chem. 2001.52:107-137. Downloaded from www.annualreviews.org by Universite Toulouse 3 -Paul Sabatier on 09/16/10. For personal use only.

Annu. Rev. Phys. Chem. 2001.52:107-137. Downloaded from www.annualreviews.org by Universite Toulouse 3 -Paul Sabatier on 09/16/10. For personal use only.

SELF-ASSEMBLED MONOLAYER FORMATION

Figure 2 Cartoons depicting typical sequences of a self-assembled monolayer structure during growth below (A) and above (B) a triple point like that shown in Figure 1. (A) Below the triple point, growth proceeds from a 2D-vapor phase, through a solid-vapor coexistence region, to the solid phase. (B) Above the triple point, the SAM must pass through three phases and two coexistence regions. The intermediate low-density phase may be a disordered (liquid) phase, a lying-down phase, etc.

111

112

SCHWARTZ

Annu. Rev. Phys. Chem. 2001.52:107-137. Downloaded from www.annualreviews.org by Universite Toulouse 3 -Paul Sabatier on 09/16/10. For personal use only.

The molecules used to create SAMs have numerous degrees of freedom, and, therefore, it is quite possible that the equilibrium phase diagram could be more complicated and involve a greater number of condensed phases than implied in Figure 1. It could include a lying-down phase and a disordered-liquid phase, for example. However, there are numerous other possibilities. Langmuir monolayers of long-chain fatty acids, for example, are known to display a variety of liquid crystalline and crystalline phases (10) that differ in the polar tilt angle, the azimuthal direction of molecular tilt (i.e. nearest-neighbor vs next-nearest-neighbor direction), and rotational freedom (herringbone vs rotator). To date there is no rm experimental evidence for liquid-crystalline phases, transient or equilibrium, in SAMs. Given their ubiquity in Langmuir monolayers and Langmuir-Blodgett lms, however, one would not be surprised if they were observed in SAMs with the appropriate experimental studies. Therefore, although it is certainly overly simplistic, the phase diagram of Figure 1 will be used as a conceptual framework to describe the experimentally observed growth mechanisms of various SAM systems discussed in the following sections. Two general experimental strategies have been used to study monolayer growth: (a) in situ studies under actual deposition conditions in real time and (b) studies on quenched partial monolayers removed from solution and possibly rinsed to remove loosely attached adsorbate molecules. Although in situ experiments have become increasingly important in recent years, many publications in the literature report experiments that used quenched lms. The clear advantage of in situ experiments is that one avoids the issue of whether the lm structure is altered by the quenching process. This is not a trivial matter, since there is clear evidence that quenching can alter the lm coverage and morphology in some molecular systems. On the other hand, working with quenched lms permits the use of certain techniques not applicable in situ, such as contact angle and X-ray photoelectron spectroscopy. Furthermore, one can work over a longer range of time scales (i.e. concentrations). Although experiments on quenched lms often report reliable and useful information (particularly on qualitative issues), one must view subtle quantitative conclusions based on quenched lms with appropriate skepticism until they are conrmed by more direct experiments.

Vapor Phase-Deposited Thiol Films


Although not strictly considered SAMs, lms created by vapor phase (molecular beam) deposition of alkylthiols on gold share many structural characteristics with solution-deposited lms. Furthermore, studies on these lms have the advantages of ultra-high-vacuum substrate cleanliness and the availability of traditional in situ surface characterization techniques. Although it is clear that solvent interactions are potentially important for SAMs (perhaps even more so during the growth process than for complete lms), many of the results obtained on vapor-deposited lms are relevant to SAM structure and growth, and they are briey reviewed here.

SELF-ASSEMBLED MONOLAYER FORMATION

113

Poirier & Pylants (11) STM observations of vapor-deposited thiols on singlecrystal Au(111) were the rst report of the general mechanism of thiol monolayer growth. Studying the formation process of C6C10 alkylthiols, both methyl and hydroxy terminated, Poirier & Pylant reported a two-step process starting with the nucleation and growth of islands of striped phases from a lowerdensity lattice-gas phase. Based on the observed periodicity of these striped phases, it was proposed that they consisted of molecules lying at on the gold surface, in either a head-to-head or a head-to-tail arrangement. The growth of the stripe phase islands was accompanied by the appearance of gold atom vacancies (pits). After the surface was covered by the stripe phase, continued deposition resulted in islands of molecules arranged in a way consistent with an epitaxial overlayer ( 3 3)R30 on the Au(111) surface. The lateral density necessary to form this structure implied thiol molecules with an orientation nearly perpendicular to the substrate. This sequence is similar to the one suggested in Figure 2b. Schreiber and coworkers (12) presented a multitechnique study (X-ray diffraction, atom diffraction, and X-ray photoelectron spectroscopy) of the vapor phase growth process of C10 thiol on Au(111). Their results were qualitatively consistent with Poirier & Pylants observations (11) at low temperatures with a few added details. The atom diffraction suggested that the striped phase became disordered prior to nucleation of the upright [( 3 3)R30 ] phase. Also, experiments at temperatures above 15 C found an additional intermediate 2D-liquid phase between the striped phase and the standing-up solid phase. To incorporate this into a phase diagram like Figure 1, one would have to add another low-density condensed phase. It is interesting that Schreiber and coworkers found that the size of correlated [( 3 3)R30 ] domains that grew from the liquid phase were signicantly larger than those that nucleated from the striped phase. This suggested that the defect structure of the nal lm may be intimately related to deposition conditions and mechanisms. In two subsequent papers (13, 14), the growth pro cess of the [( 3 3)R30 ] phase was studied in greater detail. It was found that at >15 C, the growth rate of the [( 3 3)R30 ] phase was approximately proportional to the adsorbate pressure in the gas phase. However, at lower temperatures, the growth rate was proportional to the square of the pressure, suggesting that a bimolecular process may be rate limiting.

Annu. Rev. Phys. Chem. 2001.52:107-137. Downloaded from www.annualreviews.org by Universite Toulouse 3 -Paul Sabatier on 09/16/10. For personal use only.

Thiol on Gold Self-Assembled Monolayers


Contact angle and ellipsometry experiments on quenched, incomplete alkanethiol SAMs on gold by Bain et al (15) revealed at least two time scales in the growth process. For a typical solution of 1 mM in ethanol, the contact angle and lm thickness were observed to reach 90% of their nal values within the rst minute of growth. A slower second step was observed in which the nal lm properties were reached only after several hours. This suggested multiple processes at work,

114

SCHWARTZ

with time scales differing by 2 orders of magnitude. They also noted that the monolayer properties of longer-chain n-alkyl thiols (n > 8) were consistent but that shorter-chain thiol monolayers were qualitatively different in a way that suggested greater disorder. Sun & Crooks (16) decorated defects in quenched partial SAMs by underpotential electrochemical deposition of Cu. They monitored the decrease in SAM defect density as a function of exposure time using STM and found that the number of defects disappeared on a time scale of several hours, consistent with Bains (15) results. This study also provided evidence, albeit indirect, that an islanding mechanism was involved in alkylthiol SAM growth. They also observed pits 0.5 nm deep that were characteristic of the SAM-covered gold surface. These pits did not appear to be holes in the SAM layer, however, because Cu islands did not nucleate in these locations. These intriguing results inspired a multitude of measurements designed to conrm and later explain the existence of multiple time scales. Numerous studies involving measurements of lm mass and average thickness were conducted to explore the overall coverage kinetics of thiol SAMs on gold. The results of many of these were qualitatively consistent with the observations of Bain et al (15), nding fast and slow time scales. Shimazu and coworkers (17) performed in situ quartz crystal microbalance (QCM) experiments on ferrocenesubstituted thiols in hexane solution. They observed a fast adsorption step (a few seconds in 0.5 mM solution) followed by a process with a time scale 2 orders of magnitude slower. Their results were consistent with a single molecular layer after 800 s. QCM and STM studies on quenched monolayers by Kim et al (18) detected a slow build-up of multilayers (over a period of days) during C18 thiol SAM growth from ethanol solution. Schneider & Buttrys in situ QCM experiments (19) also considered multilayer formation. However, their results suggested gradual conversion of physisorbed multilayers to a chemisorbed monolayer. Schneider & Buttry also observed a signicant solvent effect. In dimethylformamide solution, adsorption was rapid; however, a complete monolayer was never formed. In acetonitrile solution, on the other hand, adsorption was slower, but the physisorbed lm was slowly converted to a densely packed monolayer. Schneider & Buttry suggested that the nal monolayer quality had an inverse relationship with the solubility of the thiol in the solvent. In situ QCM experiments by Frub se & Doblhofer (20) o revealed two distinct time scales in adsorption from 0.1 mM thiol solutioninitial adsorption in 2 min followed by a much slower process taking >1 h. Their measurements of gradually decreasing electrochemical impedance during that latter process suggested that the slow time scale corresponded to healing of the SAM. In situ surface plasmon resonance (SPR) experiments by DeBono and coworkers (21) found two adsorption time scales differing by a factor of 100. For C12 or C16 thiols from ethanolic solution, the initial fast step resulted in 80% of monolayer coverage. The C16 rate constant was faster than the C12 for both steps. For the C6 thiol, only 50% coverage was reached in the initial step. Peterlinz & Georgiadis (22) also performed in situ SPR experiments on n-alkyl thiol SAMs grown from ethanol and heptane solutions (see Figure 3). For ethanol solution, they observed

Annu. Rev. Phys. Chem. 2001.52:107-137. Downloaded from www.annualreviews.org by Universite Toulouse 3 -Paul Sabatier on 09/16/10. For personal use only.

SELF-ASSEMBLED MONOLAYER FORMATION

115

Annu. Rev. Phys. Chem. 2001.52:107-137. Downloaded from www.annualreviews.org by Universite Toulouse 3 -Paul Sabatier on 09/16/10. For personal use only.

Figure 3 Chain length dependence of formation kinetics for C8 (circles), C12 (squares), C16 (triangles), and C18 (diamonds) thiols from 1.0 mM ethanolic solutions. The lm thicknesses were calculated from in situ surface plasmon resonance measurements. Up to three distinct kinetic regimes were observed, depending on chain length. (a) Details of kinetics at short times. (b) Overview of kinetics for 4872 h. (Figure adapted with permission from Reference 22.)

116

SCHWARTZ

multiple time scales. For the initial fast adsorption step (25 min for a 1 mM solution), the rate decreased with increasing chain length, whereas the opposite trend was observed in the second step which lasted many hours. Monolayers composed of shorter-chain molecules (C8 and C12) reached 80% of the nal thickness in the rst step while the longer-chain lms (C16 and C18) obtained only 40%50%. In the longer-chain thiol systems (C16 and C18), a third step with an even longer time scale was observed. Growth from heptane solution for the C16 thiol had a rst step with a rate somewhat faster than that of ethanol solution. In addition, the lm reached >80% of its nal thickness in this rst step. This was followed by a gradual evolution to the nal lm thickness over a period of 2 days. Shao and coworkers (23) used in situ electrochemical methods to show that the growth of an azobenzenethiol also involved a two-step process. Buck et al (24) noted only a single, relatively fast time scale (e.g. <1 min for 0.045 mM solution) from their in situ second-harmonic-generation experiments on n-alkyl thiols in ethanolic solution. The in situ QCM measurements by Karpovich & Blanchard (25) [later followed up by Schessler et al (26)] also reported only a single time scale for SAMs grown over a wide range of concentrations, using hexane or cyclohexane as the solvent. They noted, however, that the kinetic data for concentrations above 0.1 mM deviated systematically from the predictions of the Langmuir model, which involves only a single time scale. In both of these studies, the measurements may not have extended to long enough times to observe very slow later steps of growth. The in situ SHG experiments of Dannenberger et al (27) also probed only the initial stages of growth. X-ray photoelectron spectroscopy experiments by Kawasaki and coworkers (28) on quenched partial lms probed the rst 2 h of growth of C8 thiol adsorbed from 0.001 to 0.1 mM ethanolic solutions on atomically at sputter-grown gold surfaces. In this time period Kawasaki and coworkers observed two-step growth kinetics for the 0.001 mM solution but a one-step growth process for the 0.01 and 0.1 mM solutions. Spectroscopic techniques have been used to observe the evolution of molecular order within thiol SAMs in the various stages of growth that were previously identied. Near-edge X-ray adsorption ne structure spectroscopy experiments by H hner et al (29) on quenched lms probed the molecular conformation of C22 a thiol SAMs after 10-min and 43-h immersions in 3 M ethanolic solution. They reported that the angular dependence of the C-H resonance was highly anisotropic for the 43-h sample, suggesting that the molecules were conformationally ordered and uniformly tilted 35 from the surface normal. Although the data from the 10-min sample were more difcult to analyze, the anisotropy present at the long exposure time was not present, implying that the alkyl chains were disordered. They suggested that the slow time scale observed in kinetics experiments corresponded to an ordering process in which alkyl chains straighten as they disentangle. Truong & Rowntree (30) followed the infrared spectra (in the C-H stretch region) of quenched C4 thiol SAMs as a function of immersion time in 5 M methanolic solution. The signicant changes in the spectra occurred over the rst 1015 min, so these experiments probed the structural evolution in the initial period of SAM

Annu. Rev. Phys. Chem. 2001.52:107-137. Downloaded from www.annualreviews.org by Universite Toulouse 3 -Paul Sabatier on 09/16/10. For personal use only.

SELF-ASSEMBLED MONOLAYER FORMATION

117

formation. The relative band intensities suggested that the alkyl chains were, on average, lying close to the surface at rst and closer to the surface normal after 15 min of immersion. The invariance of the band frequencies suggested that the local molecular environments were insensitive to coverage. These data were consistent with a picture in which islands of vertically oriented molecules form and grow to cover the surface. Terrill and coworkers (31) obtained IR spectra from quenched C16 thiol SAMs immersed in ethanolic solutions (106102 M) for periods of 11 days. Using the position of the antisymmetric methylene stretch as a signature of chain disorder, they found that long times (from several hours to several days depending on concentration) were necessary to reach the most conformationally ordered state. They also observed that the ordering was faster on smoother substrates. On the other hand, Bensebaa et al (32) reported that this same peak position reached its ultimate value, representative of well-ordered alkyl chains, after only a 45-s immersion in 5 M ethanol solution (for quenched lms of a C22 thiol). Himmelhaus et al (33) performed sum frequency generation spectroscopy studies on quenched C22 thiol SAMs adsorbed from 3 M ethanolic solution, monitoring the various C-H stretch bands over 2 days of immersion time. They found three distinct regimes of growth. The rst stage (initial 5 min of immersion) involved formation of Au-S bonds. The coverage reached 80%90% after this stage. The second stage (515 min of immersion) was characterized by a transition of the hydrocarbon chains from a highly kinked to an all-trans conformation. The nal stage (20 min to 2 h of immersion) involved reorientation of the terminal methyl groups from a state in which methyl groups were disordered relative to one another to one in which they were aligned. The authors pointed out that this sequence implies that the ordering process can be viewed as consecutive steps originating at the gold interface and moving toward the lm surface. Humbert and coworkers (34) performed SFG studies on quenched para-nitroanilino C12 thiol SAMs deposited from 2 M ethanolic solution. They observed a marked change in molecular orientation over the rst 30 min, followed by a slower change over the next 90 min, after which their observations ended. In recent years, several scanning-probe-microscopy experiments have shed light on the thiol growth process. In a sequence of two papers (35, 36), Yamada & Uosaki performed in situ STM experiments monitoring alkylthiol growth on Au(111) from micromolar heptane solutions. They observed three basic steps. Initially, patches of adsorbed molecules were observed, but no periodic structures were detected on molecular-length scales. The authors suggested that these patches might correspond to a disordered phase. In this stage of growth, pits (or vacancy islands) were formed in the gold. The second step involved the appearance of patches in which striped patterns were observed. Although several periodic length scales were found, all were greater than the molecular length (and increased systematically with alkyl chain length). These observations and the structural similarities to the features observed on vapor-phasedeposited thiol lms suggested that these stripe phase domains were composed

Annu. Rev. Phys. Chem. 2001.52:107-137. Downloaded from www.annualreviews.org by Universite Toulouse 3 -Paul Sabatier on 09/16/10. For personal use only.

118

SCHWARTZ

Annu. Rev. Phys. Chem. 2001.52:107-137. Downloaded from www.annualreviews.org by Universite Toulouse 3 -Paul Sabatier on 09/16/10. For personal use only.

of thiol molecules lying down on the surface in various ordered epitaxial arrangements. In the third and nal stage of growth, islands of apparently greater lm thickness formed and grew to cover the surface. A hexagonal pattern was observed on these islands, consistent with the ( 3 3)R30 epitaxial arrangement of thiol molecules well-known from STM and scattering experiments (as discussed above). This growth sequence is reminiscent of the path discussed in an earlier section of this review through a quasi-equilibrium phase diagram at temperatures above the triple point (Figure 2b). Xu and coworkers (37) performed detailed and quantitative in situ atomic-force microscopy (AFM) experiments to follow the growth of C18 and C22 thiol molecules from 2-butanol solution (see Figure 4). Their observations were consistent with those of Yamada & Uosaki but provided direct height information. They rst observed the formation of patches that were 0.5 nm high, consistent with molecules lying on the surface. At longer exposure times, islands that were 1.8 nm higher than the lying-down phase were observed to nucleate and grow, consistent with a structure in which molecules were approximately vertically oriented. The transition from lying down to standing up was faster for the C22 than the C18 thiol lm. For 0.2 mM C18 thiol solution, the time elapsed between the initial appearance of lying-down

Figure 4 In situ topographic atomic-force microscopy images of Au(111) obtained at various times after injection of a solution of C18 thiol (0.2 mM in 2-butanol). The area of each frame is 150 nm 150 nm2, and the bright spot in the lower left corner and the Au(111) steps at the bottom of each frame provide landmarks for comparison of the images. (Figure adapted with permission from Reference 37.)

SELF-ASSEMBLED MONOLAYER FORMATION

119

patches, and essentially complete coverage of the standing-up phase was 1015 min. Tamada et al (38) also observed island growth with AFM on quenched partial monolayers.

Silane Self-Assembled Monolayers


The growth of trichlorosilane (similarly trimethoxy- or triethoxysilane)-based SAMs is unique among SAM systems in that it involves an irreversible covalent cross-linking step. This is critical to the desirable properties of this class of SAMs, including their chemical and mechanical robustness on a variety of substrates. There is also the potential for hydrolytic bond formation to OH surface groups that would immobilize adsorbate molecules. Again, this is critical to the stability of the nal monolayer. The kinetics of this step relative to the self-assembly process can clearly have dramatic implications on the growth mechanisms and nal lm structure. This adds a number of complications, because the rate of hydrolysis is sensitive to water content, pH, and temperature. It is interesting that, since the molecular packing is ultimately determined by the covalent siloxane network, one does not nd long-range molecular order in these SAMs as one does in thiol SAMs, where the epitaxial arrangement on the Au lattice dictates the molecular arrangement. In two X-ray reectivity studies, Wasserman et al (39) and Tidswell et al (40) determined the electron density proles of quenched partial and complete monolayers formed from C10C18 trichlorosilanes on silicon oxide substrates. They found that the structure of partial monolayers was inconsistent with molecular islands. An AFM study by Schwartz and coworkers (41) of quenched C18 silane SAMs on mica, on the other hand, explicitly observed 2-nm-high islands that grew to cover the surface with increasing immersion time. They found that the islands were fractal in shape and that the scaling exponent (fractal dimension) of 1.7 in the early stages of growth was consistent with 2D-diffusionlimited aggregation. This suggested an island growth mechanism involving collisions between adsorbate molecules moving randomly on the surface and immobile islands. The assumption of irreversible attachment to islands led to the fractal shape. This was essentially a view of the sequence shown in Figure 2a from a kinetic (rather than a thermodynamic) perspective and was clearly inconsistent with the conclusions of the prior X-ray studies. However, the mica surfaces were known to have only isolated-exposed OH sites appropriate for anchoring the monolayer, whereas such sites were ubiquitous on silica substrates. A later AFM study by Bierbaum & Grunze (42) on quenched C18 (and longer) silane SAMs on silicon oxide observed similarly shaped islands. They did not observe islands on partial C3 silane monolayers. Interestingly, second harmonic generation experiments by Zhao & Kopelman (43) showed that only a small minority of the surface OH groups on the silica surface are actually bound to the SAM. Therefore, the picture of large, interconnected monolayer domains anchored to the substrate only in isolated locations may be appropriate for a variety of substrates. Using AFM,

Annu. Rev. Phys. Chem. 2001.52:107-137. Downloaded from www.annualreviews.org by Universite Toulouse 3 -Paul Sabatier on 09/16/10. For personal use only.

120

SCHWARTZ

Annu. Rev. Phys. Chem. 2001.52:107-137. Downloaded from www.annualreviews.org by Universite Toulouse 3 -Paul Sabatier on 09/16/10. For personal use only.

Kropman and coworkers (44) observed dendritic C18 silane islands in quenched partial monolayers prepared even on SrTiO3 substrates. As mentioned above, the competition between various time scales in silane SAM growth makes the process quite sensitive to variations in preparation conditions. For example, contact angle (45) and IR spectroscopy (46) studies showed that quenched C18 silane SAMs prepared on silica substrates at a temperature below 30 C contained well-ordered alkyl chains, while those prepared at higher temperatures contained increasing chain disorder. This led to a quasi-equilibrium picture of silane SAM growth similar to those steps illustrated in Figure 1 and Figure 2 above. However, the nomenclature used was borrowed from the Langmuir monolayer literature; therefore the liquid phase was labeled LE and the solid phase LC. Carraro et al (47) obtained AFM images of quenched partial C18 silane SAMs on silicon oxide over a range of temperatures (see Figure 5). At a low temperature (10 C), dendritic islands were observed to grow and coalesce to cover the surface, while at a high temperature (40 C) only a homogeneous uniform lm was

Figure 5 Atomic-force microscopic images of partial octadecyltrichlorosilane self-assembled monolayers on silicon oxide removed from 2 mM solution (using hexadecane-carbon tetrachloride as a solvent) after 30 s of immersion at 10 C, 25 C, and 40 C (left to right). The height distribution histograms are also shown. The differences in structure as a function of temperature are consistent with a quasi-equilibrium-phase diagram as discussed in the text. (Figure adapted with permission from Reference 47.)

SELF-ASSEMBLED MONOLAYER FORMATION

121

observed. At an intermediate temperature (25 C), some dendritic islands were observed to nucleate and grow. However, before they could coalesce and cover the entire surface, the continuous phase between the islands gradually increased its thickness to that of the islands, ending the lm formation. A similar AFM study by Goldmann and coworkers (48) observed quenched partial C18 silane SAMs on silicon oxide prepared at 12, 21.5, 26.5, 35, and 43 C. They observed regions of three heights that they considered to be vapor, liquid, and solid, respectively (although their nomenclature was G, LE, and LC). At 26.5 C and below, they observed sequential transitions from vapor to liquid to solid involving domain nucleation and growth. The vapor-liquid coexistence region was characterized by a foamlike morphology. At 35 C, only the evolution from vapor to liquid was observed. The results of both papers can be interpreted via a phase diagram like Figure 1, where the triple point is between 10 and 12 C if one assumes that the growth is not under quasi-equilibrium conditions. Below the triple point, the growth process is typical of 2D vapor-solid coexistence. Above the triple point (but below 30 C), the vapor-liquid transition is observed followed by the liquid-solid transition. Although solid-phase islands nucleate from the liquid phase, they do not grow quickly enough to maintain quasi-equilibrium conditions, and the surrounding liquid phase becomes more and more super concentrated, therefore thicker. Above 30 C, it is possible that concentrations necessary to nucleate the solid phase are not reached before the SAM growth is quenched by cross-linking or that the surface density necessary for nucleation of the solid phase is not accessible via spontaneous adsorption from solution. A recent experiment by Sung et al (49) on a quenched C18 silane SAM explicitly demonstrated a phase transition consistent with the results of these temperature-dependent adsorption studies. A partial lm was prepared in the vapor-solid coexistence region at 10 C, removed from solution, and heated to 30 or 60 C. The sample heated to 30 C had a lower area fraction of solid-phase islands than the unheated lm and the islands had disappeared completely on the lm heated to 60 C. If the lm was heated to 60 C and then cooled to 30 C, islands were observed to form; if cooled to 10 C, the islands were larger and covered more of the surface. This study explicitly veried the phase diagram paradigm as well as the mobility of the molecules even after quenching. The mobile state was found to last for several minutes, after which cross-linking and grafting apparently froze the lm morphology. In a series of several papers (5052), the Hoffmann group explored the effect of deposition conditions (water content and solution age) and substrate on the structure of quenched alkylsiloxane SAMs formed at room temperature. Their AFM images suggested simultaneous growth by island formation and a continuous disordered-liquid phase. The relative contributions of the two mechanisms were sensitive to growth conditions. Island growth was favored in deposition solutions with higher water content (in toluene solution) or solutions that had been aged longer. In addition, they found that adsorption was faster on mica than on silicon oxide. They also prepared mica substrates coated with a specic number of silicon oxide layers and found that the growth rate decreased exponentially with

Annu. Rev. Phys. Chem. 2001.52:107-137. Downloaded from www.annualreviews.org by Universite Toulouse 3 -Paul Sabatier on 09/16/10. For personal use only.

122

SCHWARTZ

Annu. Rev. Phys. Chem. 2001.52:107-137. Downloaded from www.annualreviews.org by Universite Toulouse 3 -Paul Sabatier on 09/16/10. For personal use only.

increasing oxide coating thickness up to about six layers. In situ IR spectroscopy experiments found that the silane molecules adsorbed initially in a disordered conformation and gradually aligned and stood up as the coverage increased. Furthermore, they observed an enhancement in the adsorption kinetics with increasing water content of the deposition solution. In a later in situ AFM study (53) conducted at room temperature, the same group observed only islands 2.5 nm high during C18 silane growth. The discrepancy between these observations and the AFM images of quenched lms (which showed both islands and continuous phase) casts some doubt on the relevance of the quenched-lm studies. Richter and coworkers (54, 55) performed X-ray reectivity studies of C18 silane lm growth on silicon oxide at room temperature from micromolar-concentration heptane solution. In in situ experiments, they found density proles that suggested that the maximum lm thickness did not change during growth but that the average density gradually evolved to that of a complete monolayer. This was consistent with island growth of approximately vertically oriented molecules. Richter and coworkers compared the structure of partially formed monolayers during these in situ experiments with quenched partial monolayers and found systematic differences. The quenching process apparently introduced free area into the lm that was not restored by reintroducing the quenched lm into solvent. This suggested that some adsorbate molecules were removed during quenching. Although both the average lm thickness and density were affected by the quenching process, the density decreased more dramatically, which suggested that regions of relatively densely packed molecules were not signicantly affected by quenching, while other, less dense regions lost most of their molecules, with the remaining molecules tilting over dramatically. This experiment again sounds a warning regarding overinterpretation of experiments based on quenched partial monolayers.

Other Self-Assembled-Monolayer Systems: Organic Acids and Ions


Although thiol- and silane-based systems represent the bulk of the SAM literature, there are a number of reports of monolayers based on organic acids or ions. For example, alkyl carboxylic, sulfonic, and phosphonic acids have been demonstrated to form organized monolayers on several metal or metal oxide surfaces. Also, organic ions, such as quaternary ammonium salt detergents, form stable-monolayer lms on substrates like mica that have a nonzero net charge at accessible pH values. Aside from the practical signicance of expanding the range of substrates that may be coated with SAMs, these systems offer the opportunity to explore how the adsorbate-substrate interaction affects the assembly process, because the type of interactions (acid-base or ionic) are in stark contrast with those in thiol or silane SAMs. The growth process of octadecylphosphonic acid (OPA) on mica from tetrahydrofuran solution was explored by our group in a series of papers (5660). AFM observations of quenched samples showed evidence of island growth (56) and

SELF-ASSEMBLED MONOLAYER FORMATION

123

suggested that the regions between islands were virtually bare, since the mica atomic lattice could be imaged. Complementary IR spectroscopy and contact angle studies (58) conrmed that the islands were surrounded by a 2D vapor. In situ AFM experiments (57, 59, 60) conrmed these qualitative observations, putting OPA on mica within the class of SAMs that forms via the sequence shown in Figure 2a. Looking closely at the coverage kinetics of quenched OPA-mica lms (58), it was determined that a signicant amount of adsorbate molecules (5%25% coverage over the concentration range of 0.022 mM) was deposited during sample removal from solution via a Langmuir-Blodgettlike transfer process. This process also resulted in a dramatically altered island size distribution, as new islands nucleated and existing islands grew during sample removal. After an attempt was made to account for the effects of quenching, the island coverage kinetics were explored as a function of concentration in the range 0.022 mM (58). The data were collapsed onto a single curve with the introduction of a concentration-dependent time scale. However, the curve was not in good agreement with Langmuir kinetics. Although the growth kinetics could be explained using a diffusion-limited functional form, the diffusive time scales extracted from the ts were much slower than expected for solution-phase molecular diffusion and did not display the expected concentration dependence. It is interesting that Peterlinz & Georgiadis (22) made similar observations regarding thiol SAM growth. For OPA, however, it is not clear that the apparent form of the coverage kinetics may not simply be an artifact of studying quenched lms, since later in situ AFM experiments found that the early stages of growth were consistent with Langmuir kinetics and became complicated in the later stages (which is addressed in more detail in a later section). Hayes and coworkers (61) explored the monolayer growth of the organic cation octadecyltrimethylammonium bromide (OTAB) on mica from aqueous solution. The room temperature experiments were performed below the Krafft point (the temperature above which micelles are observed). Their AFM, contact angle, and IR spectroscopy observations on quenched partial monolayers found extremely slow growth of islands, with complete monolayers forming only after 2 weeks of immersion time. The atomic lattice of mica could not be imaged with AFM on the areas surrounding the islands, and IR and contact angle data suggested that these regions were covered by a thin lm of disordered molecules (a 2D liquid). Thus the OTAB monolayer appeared to grow via a sequence consistent with the one shown in Figure 2b. Although AFM data may suggest which phase is in coexistence with solid islands by means of island heights, relative friction, or stiffness (using phase contrast), these observations are rather indirect. In general, it is preferable to use complementary techniques that are sensitive to molecular conformation to clarify this issue. In this regard, it is instructive to compare the contact angle (Figure 6) and IR (Figure 7) signatures of 2D vapor-solid coexistence [as exemplied by OPA SAMs on mica (58)] with those of 2D liquid-solid coexistence [using OTAB SAMs on mica as an example (61)].

Annu. Rev. Phys. Chem. 2001.52:107-137. Downloaded from www.annualreviews.org by Universite Toulouse 3 -Paul Sabatier on 09/16/10. For personal use only.

124

SCHWARTZ

Annu. Rev. Phys. Chem. 2001.52:107-137. Downloaded from www.annualreviews.org by Universite Toulouse 3 -Paul Sabatier on 09/16/10. For personal use only.

Figure 6 Comparison of the cosine of the contact angle of water to the fractional coverage of partially formed lms of octadecyltrimethylammonium bromide (OTAB) on mica (open circles) and octadecylphosphonic acid (OPA) on mica ( lled circles). The different behavior at low coverage suggests that solid-phase islands on partial OPA self-assembled monolayers (SAMs) are surrounded by a two-dimensional (2D) vapor, while those on OTAB SAMs are surrounded by a 2D-liquid phase. (Figure adapted with permission from Reference 61.)

Applying the Cassie equation (62) to the case of solid-phase islands in coexistence with a dilute phase yields the following predictions for the cosine of the contact angle : cos = cos dilute + island (cos island cos dilute ) where island is the fractional surface coverage of the island phase and dilute and island are the contact angles on a surface composed purely of the respective phase. This equation predicts that the extrapolated value of cos at zero island coverage will be equal to the cosine of the contact angle on a surface composed purely of the dilute phase, which surrounds the islands. As shown in Figure 6, the extrapolated value for OPA is close to unity, implying that water would wet the dilute phase. This is consistent with the dilute phase being bare mica or mica with a very low coverage of adsorbed surfactant moleculesa 2D-vapor phase of OPA molecules. On the other hand, for OTAB, cos extrapolates to 0.4 at zero coverage, implying that the dilute phase is fairly hydrophobic, with a contact angle of 65 . This is consistent with a disordered layer (a 2D liquid), for example. IR studies of partial OPA and OTAB monolayers were consistent with these conclusions. Figure 7 shows the methylene stretch region of partial monolayers

SELF-ASSEMBLED MONOLAYER FORMATION

125

Annu. Rev. Phys. Chem. 2001.52:107-137. Downloaded from www.annualreviews.org by Universite Toulouse 3 -Paul Sabatier on 09/16/10. For personal use only.

Figure 7 Transmission IR spectra of quenched (a) octadecylphosphonic acid (OPA) and (b) octadecyltrimethylammonium bromide (OTAB) self-assembled monolayers on mica taken after increasing immersion times in 0.2 mM and 0.1 mM solution, respectively. (Figure adapted with permission from References 58 and 61.)

of OPA (Figure 7a) and OTAB (Figure 7b) after increasing immersion times (top to bottom in each gure). For OPA, although the peaks were observed to grow as immersion time and island coverage increased, the peak positions remained at the same wavelength. For example, the antisymmetric methylene stretch was found at 2919 cm1, consistent with well-ordered all-trans alkyl chains (63), even at a fractional island coverage of only 0.1. This suggested that a negligible fraction of

126

SCHWARTZ

Annu. Rev. Phys. Chem. 2001.52:107-137. Downloaded from www.annualreviews.org by Universite Toulouse 3 -Paul Sabatier on 09/16/10. For personal use only.

the molecules on the surface were disordered. On the other hand, the peak positions for OTAB gradually shifted to lower wave numbers with increasing immersion time and island coverage. The antisymmetric peak position moved from 2924 cm1, consistent with disordered alkyl chains, to 2919 cm1. For samples partially covered by islands, this peak is presumably the convolution of two peaks, one for the molecules within islands and one for the molecules in the dilute phase between islands. The relative weighting of the two peaks changes with the island coverage, resulting in the apparent peak position shift, which is consistent with a signicant coverage of disordered molecules in the region between islands. For both systems, the contact angles and IR spectra tell the same storythe dilute phase surrounding the islands may be either a 2D vapor (as for OPA) or a 2D liquid (as for OTAB), depending on the system chemistry and thermodynamic conditions.

QUANTITATIVE ASPECTS OF GROWTH PROCESSES Rate Constants/Time Scales


Many reported rate constants or time scales for SAM growth have been collected in Figures 8a and 8b, cast in terms of time constants. For cases in which rates constants were reported, the time constants were calculated simply as the inverse of the rate constants. Although rate and time constants clearly contain the same information, time constants are presented in the hope that the plotted values will have more intuitive value to the reader. Figure 8a displays data for thiol SAMs, and Figure 8b includes data for silanes and acid-based monolayers. The large scatter in the thiol data is particularly noticeable. For any concentration in the range 106103 M, the spread in the measured values of time constants is typically 2 orders of magnitude. Data have been included for a range of chain lengths (C12C22) and a variety of solvents (mostly alcohols and alkanes). Also, several different theoretical models were used to extract rate or time constants. However, these variables typically introduce variations in rates only of order unity. Data included in Figure 8a were determined using a variety of techniques, both in situ and on quenched partial lms; however, there is no real pattern or consistency even when considering only individual techniques or methods. Thus, one is left with the impression that there may be real differences in the growth kinetics of thiol SAMs in different laboratories. It is unclear which parameters are not controlled; one possibility that has been suggested is the substrate roughness or microcrystallinity. In contrast, the data in Figure 8b is surprisingly consistent even though time scales are included for silane SAMs on a variety of substrates (open symbols) in addition to carboxylic and phosphonic acid SAMs ( lled symbols). Again, the data represent a variety of techniques, some in situ and some on quenched partial lms. A casual inspection also reveals that these SAMs grew consistently more slowly than the thiol SAMs. The dotted lines on the graphs in Figure 8 represent the inverse dependence between concentration and time constant (equivalent to a linear dependence of

SELF-ASSEMBLED MONOLAYER FORMATION

127

Annu. Rev. Phys. Chem. 2001.52:107-137. Downloaded from www.annualreviews.org by Universite Toulouse 3 -Paul Sabatier on 09/16/10. For personal use only.

Figure 8 Time scales of self-assembled monolayer (SAM) growth versus solution concentration summarized from a variety of reports. (a) All symbols represent Alkanethiol SAMs. (b) Silane (open symbols) and acid ( lled symbols) SAMs.

128

SCHWARTZ

Annu. Rev. Phys. Chem. 2001.52:107-137. Downloaded from www.annualreviews.org by Universite Toulouse 3 -Paul Sabatier on 09/16/10. For personal use only.

rate on concentration) that one would expect for any model involving adsorptionlimited growth. There appears to be a reasonable agreement for silanes and acids (Figure 8b) above 105 M. Some of the individual thiol data sets (15, 21, 27) in Figure 8a are approximately consistent with the inverse dependence of time constant and concentration. However, others, such as those from the Blanchard group (25, 26) and the Georgiadis group (22), report time constants that decrease much more slowly. It has been suggested (25, 26) that this can be explained by a signicant desorption rate, which is not concentration dependent. The discrepancy regarding the concentration dependence of growth kinetics between different laboratories remains unresolved.

Functional Form of Coverage Kinetics


In most cases, the coverage kinetics are compared to the simple Langmuir or Avrami kinetic model, which assumes that the rate of deposition is proportional to the free space on the surface, that is, d/dt = k(1 ), where k is a rate constant. This leads to the integrated expression = 1 exp(kt), which has frequently been used to t kinetic data for SAM growth, sometimes because it agreed quantitatively with the data and on other occasions simply because it was considered the simplest model to use when the low precision of the data did not justify using a more complicated model. The time constants reported in the previous section are simply the inverse of the rate constant, that is, time constant = 1/k. In situ nonlinear optical experiments on thiol SAM growth (24, 27, 33, 64) have typically found that Langmuir kinetics [or a variant involving multiple adsorption sites (27)] described the early stages of adsorption reasonably well, as did several in situ SPR (21, 65), QCM (25, 26), and AFM (37) studies. In a careful SPR study that divided the growth process into three regimes, Peterlinz & Georgiadis (22), however, found that the early stages were equally well described by second-order Langmuir kinetics or a diffusion-limited form. A second step was found to obey zero-order kinetics. In electrochemical studies of azobenzenecontaining thiol monolayers, Shao et al (23) found that kinetics based on a Frumkin isotherm approach (which includes adsorbate-adsorbate interactions) described the data better than Langmuir kinetics. In situ IR (51, 66) and X-ray reectivity (55) studies of alkylsilane SAM growth also found reasonable agreement with Langmuir kinetics, except for early times (55, 66) or for solutions with high water content (51). IR spectra of quenched partial monolayers of alkyl carboxylic acids on aluminum oxide (67) were consistent with Langmuir kinetics, as were in situ AFM studies (59) of the early stages of phosphonic acid monolayer growth on mica. Given what is known about the details of the growth processes for the various SAM systems discussed above, it is somewhat surprising that the simple Langmuir form effectively describes SAM formation that involves island nucleation and growth, multiple 2D-phase transitions, etc. Certainly in some cases the data may simply be too imprecise to justify using other models with greater numbers of

SELF-ASSEMBLED MONOLAYER FORMATION

129

free parameters. Also, although it often describes the coverage kinetics well over a large range of the growth, there are clear limitations to the Langmuir model at very early times and for late times. At early times, there is often a discrepancy between observed coverage and the prediction from the Langmuir form. Sometimes an induction period is observed before growth starts. Also, the last 10% 20% of monolayer growth is generally found to evolve with a slower time scale than the earlier regimes. However, the simple Langmuir model is remarkably robust, and thus it is worthwhile to consider what this tells us about the growth process. For systems in which SAM growth involves a 2D-vapor-to-solid transition, the signal observed by any of the typical techniques is dominated by the molecules in the solid islands. Thus Langmuir kinetics are consistent with a growth rate that is proportional to the area uncovered by islands. If adsorption occurs primarily on these uncovered regions and the overall SAM growth kinetics is adsorptionlimited, this would explain why Langmuir kinetics work well in such cases. For cases in which a low-density phase (2D-liquid or lying-down phase) forms before the solid phase, however, it is somewhat more difcult to justify a simple Langmuir model. With thiols, for example, one wonders which part of the growth process is being observed in SPR, QCM, or spectroscopic measurements. Is it the formation of the lying-down phase, the conversion to the standing-up phase, or some combination of the two? It is certainly reasonable that the formation of the original lying-down phase would follow Langmuir kinetics. However, this would account for only 20% of the nal lm coverage, whereas many experiments assert that Langmuir kinetics describes the initial 50%80% of lm growth. In the only experiment that could directly separate the two processes, an in situ AFM study (37), growth of the two phases were observed to occur sequentially but with approximately the same time constant. It is not clear, however, that the kinetics of each of the processes have the same concentration dependence. Another possibility is that some of the techniques used might not be particularly sensitive to molecules in the lying-down phase, and/or the signal caused by this phase is part of the experimental baseline. In such a situation, the experiment would report essentially the growth of the solid-phase islands, and Langmuir kinetics would imply that the kinetics are limited by adsorption in the regions covered by the lying-down phase. These assumptions are not particularly satisfying, however, and the connection between the detailed growth mechanism and the macroscopically averaged coverage kinetics remains an open question for thiol SAM growth. Particularly useful in resolving these issues will be in situ techniques that are capable of discriminating between surface-bound molecules that are in different 2D phases.

Annu. Rev. Phys. Chem. 2001.52:107-137. Downloaded from www.annualreviews.org by Universite Toulouse 3 -Paul Sabatier on 09/16/10. For personal use only.

Solvent Effects
Several studies have addressed the effect of different solvents on the growth kinetics of thiol SAMs; solvents typically used were n-alkanes and ethanol. Peterlinz &

130

SCHWARTZ

Annu. Rev. Phys. Chem. 2001.52:107-137. Downloaded from www.annualreviews.org by Universite Toulouse 3 -Paul Sabatier on 09/16/10. For personal use only.

Georgiadis (22) found that the initial stage of thiol SAM growth from heptane solution was 35% faster than that from ethanolic solution. Karpovich & Blanchard (25) did not observe signicant differences using hexane or cyclohexane as solvents. Dannenberger et al (27) found that solvents affected the thiol SAM growth kinetics in the order (fastest to slowest) hexane > ethanol > dodecane > hexadecane. Although this order coincides with the solvent viscosity (which would affect the molecular diffusivity in solution), there is ample evidence to suggest that SAM growth is typically not limited by bulk diffusion at micromolar concentrations or above. It was suggested that a limiting step might involve the displacement of solvent molecules by adsorbate molecules at the surface, so that solvents with stronger surface interactions would result in slower adsorption and SAM growth.

Chain Length Effects


The literature is full of dramatically conicting reports regarding the effects of chain length on thiol SAM growth kinetics. Regarding the initial fast stage of growth, Bain et al (15) found that C18 grew faster than C10 from ethanolic solution, Xu and coworkers (37) reported that C22 formed more quickly than C18 from 2-butanol solution, and Jung & Campbell (65) performed a systematic SPR study and found that the growth rate increased with chain lengths in the range C2C18 from ethanolic solution. Thus, these studies consistently found that adsorption rate increased with chain length. Other studies reported exactly the opposite trend, however. Peterlinz & Georgiadis (22) reported growth rates for the initial step in the order C8>C12>C16>C18 from ethanolic solution, and Dannenberger et al (27) found that growth rates obeyed the trend C4>C12>C22 for both ethanolic and hexane solution. Complicating the matter even further, two additional reports were inconsistent with all of these results. DeBono and coworkers (21) found that the initial stages of growth for C16 thiol occurred at about the same rate as C6 and that both were faster than C12 from ethanolic solution. Karpovich & Blanchard (25) found that the early stages of growth for C8 and C18 thiols (from hexane solution) were approximately equal in overall rate. Analyzing the concentration dependence of the growth kinetics, they reported that the adsorption rate for C18 was greater than that for C8, but that the desorption rates had the opposite behavior. There is also some confusion regarding the chain length dependence of the later slow-growth regime. Peterlinz & Georgiadis (22) reported that the rate of this process increased with chain length from C12 to C16 to C18. DeBono et al (21) also found that C16 was faster than C12, but they observed that the trend was reversed for C6, which was equally as fast as C16. There is, unfortunately, little basis on which to critically analyze these results. The discrepancies do not divide along the lines of experimental technique, solvent, concentration range, or any other obvious parameter. In addition, it is not intuitively clear which trend should be correct for a given regime. In considering

SELF-ASSEMBLED MONOLAYER FORMATION

131

a hypothetical activated process for adsorption, one might think that the enhanced interactions between a longer chain and the surface would lower the energy barrier and increase the adsorption rate (65). On the other hand, if mobility is an issue, longer chains might move more slowly. It is clear that none of the results summarized above are dominated by bulk solution-phase molecular diffusion because of the absolute rates, the details of the time dependence, and the concentration dependence of the rate constants. However, one cannot rule out the importance of molecular mobility in moving through a hypothetical physisorbed layer (22, 27), etc.
Annu. Rev. Phys. Chem. 2001.52:107-137. Downloaded from www.annualreviews.org by Universite Toulouse 3 -Paul Sabatier on 09/16/10. For personal use only.

Adsorption Energetics
Several approaches have been used to get at the energetics of thiol SAMs. Bain et al (15) determined desorption rates of alkanethiol SAMs into hexadecane at 83 C. Assuming an Arrhenius-type expression, they found that the activation energy for desorption increased by 0.2 kcal/mol for each methylene group. Their estimate for the absolute activation energy for C22 thiol was 28 kcal/mol. Jung & Campbell (65) determined the sticking probabilities of various-chain-length thiols by analyzing the observed SAM growth kinetics with a model incorporating molecular diffusion in solution and adsorption from the subsurface layer. Again assuming an activated energy process for adsorption, they reported that the activation free energy for adsorption decreased by 0.16 kcal/mol per methylene group and that the absolute activation extrapolated to 11 kcal/mol for zero chain length. The Blanchard group (25, 26) determined the free energy of adsorption, Gads, of C8 and C18 thiols by analyzing the concentration dependence of the observed growth rate constant. They found that Gads = 5.5 kcal/mol for C18 and 4.4 kcal/mol for C8 thiol SAMs. By measuring the temperature dependence of Gads for C18, they found the molar enthalpy of adsorption, Hads = 48 kcal/mol, and the entropy of adsorption, Sads = 48 cal mol1 K1. It should be noted that these measurements are not completely consistent. For example, one would expect that Gads should be approximately the difference between the activation energies for desorption and adsorption. Using the values from the above references, this would give approximately Gads 827 = 19 kcal/mol for C18 thiol compared with the 5.5 kcal/mol quoted by Karpovich et al (25). However, these absolute free energies involve an approximate value of the pre-exponential frequency factor in the Arrhenius expression and are, therefore, somewhat arbitrary. Considering the change with chain length, one nds that the activation energy measurements predict that longer chains will be stabilized by approximately 0.2 + 0.16 = 0.36 kcal/mol per methylene group. This predicts that Gads for C18 should be 3.5 kcal/mol lower than that for C8, whereas the value quoted is only 1.1 kcal/mol lower. Of course, there were numerous simplications in the analyses, not least of which was the assumption of Arrhenius-like behavior in experiments performed at only one temperature (15, 65). SAM energetics should be a fruitful area for future work.

132

SCHWARTZ

Submonolayer Island Nucleation, Growth, and Size Distributions


Doudevski et al (59) analyzed their in situ AFM images of OPA SAM growth on mica to determine the time dependence of the island number density as well as the growth rates of individual islands. The island density (and more generally the full island size distribution) is frequently used in the literature of vaporphase epitaxial growth to characterize the submonolayer lm morphology (9, 68 72). Figure 9 shows the island number density per site, (a site is calculated as the approximate cross-sectional molecular area) N, as a function of time. Three regimes of growth were observed. For short times (growth regime; <600 s), the number of islands increased, indicating that nucleation of new islands occurred. At intermediate times (aggregation regime; 6003000 s), the island density was approximately constantnucleation of new islands virtually stopped, and existing

Annu. Rev. Phys. Chem. 2001.52:107-137. Downloaded from www.annualreviews.org by Universite Toulouse 3 -Paul Sabatier on 09/16/10. For personal use only.

Figure 9 A log-log plot of the number density of islands per site (estimated at 0.25 nm2) vs time, extracted from in situ atomic-force microscopy images during the growth of octadecylphosphonic acid self-assembled monolayers on mica from 0.15 mM tetrahydrofuran solution. The various regimes of growth are described in the text. The line represents the best t to a power law time dependence at early times. The best-t exponent is 0.31 0.5, consistent with the point island model value of one-third. (Figure adapted with permission from Reference 59.)

SELF-ASSEMBLED MONOLAYER FORMATION

133

islands gradually grew. At later times (coalescence regime; >3000 s), the island density decreased rapidly due to the merging of individual islands. At short times, the number density was found to have a power law dependence on deposition time with an exponent of 0.31 0.05, consistent with the point island model prediction of one-third (68, 69). This time dependence also implied that the critical nucleus consisted of two molecules. The growth kinetics of individual islands also had a power law form, with an exponent of 0.70 0.08, again consistent with the point island model prediction of two-thirds (68, 69). By comparing the rates of island nucleation and growth, Doudevski et al (59) inferred a value of the surface diffusivity for adsorbate molecules of D = 1.1( 0.1) 106 sites/s = 2.9(0.3) 109 cm2/s. In other work, Doudevski & Schwartz (60) measured the island size distribution in the aggregation regime (in which the island number density was approximately constant) of OPA SAM growth. As expected qualitatively, with increasing coverage the peak position of the distribution gradually moved to larger island size, and the distribution broadened considerably. They used these distributions to test one of the fundamental assumptions of cluster growth, the dynamic-scaling assumption. The essential concept of the dynamic-scaling assumption is that, at a given stage of growth, there is only a single characteristic length scale. This length can be taken to be S, the average island size, which is a function of the fractional island coverage . If this assumption is correct, then the island size distribution function can be written as Ns ( ) = S2 f (s/S ) (68), where Ns( ) is the number density of islands containing s molecules at coverage . That is, the island size distribution can be factored into two partsone that contains all dependence on coverage and length scale and another that is a scale-invariant fundamental distribution function, f. Upon applying this scaling form, the island size distributions obtained at various stages of the aggregation regime were found to collapse onto a single function f (s/S) = S2 1 Ns (), consistent with the dynamic-scaling-assumption prediction. The shape of this fundamental size distribution was different than expected from kinetic Monte Carlo simulations of epitaxial growth (68, 72), in that the peak was shifted to smaller island sizes and the distribution did not extrapolate to zero for small island sizes. This suggested the importance of additional processes not included in these simulations, such as desorption from island edges or long-range interactions. The shape of the distribution did rule out the possible inuence of Ostwald ripening, however (73).

Annu. Rev. Phys. Chem. 2001.52:107-137. Downloaded from www.annualreviews.org by Universite Toulouse 3 -Paul Sabatier on 09/16/10. For personal use only.

MANIPULATING GROWTH WITH EXPERIMENTAL PROBES


Owing to the current interest in nanotechnology, there has been recent interest in using localized experimental probes (typically AFM tips) to fashion small features in SAMs. Although not strictly within the topic of this review, such scanning probe lithography represents a promising technological area that relies in many

134

SCHWARTZ

Annu. Rev. Phys. Chem. 2001.52:107-137. Downloaded from www.annualreviews.org by Universite Toulouse 3 -Paul Sabatier on 09/16/10. For personal use only.

ways on the SAM growth process. Two approaches are particularly relevant. The Liu group has developed a technique they call nanoshaving/nanografting (74, 75). The concept involves applying a large enough load while scanning with an AFM tip so that molecules are mechanically removed from a SAM. If this procedure is performed while the SAM is in contact with a solution of adsorbate molecules, those molecules adsorb and assemble to ll in the region previously shaved by the AFM tip. Using this method, the group has demonstrated the ability to create features with typical length scales of 10 nm. The features can be raised or recessed, or they can have a different chemical surface functionality relative to the surrounding surface, depending on the choice of adsorbate molecules used to create the original resist SAM and those present in solution during the shaving. For thiol SAMs, they found that, if the dimension of the shaved region was smaller than the extended length of the adsorbate molecule, the regrowth of SAM in the shaved region was approximately an order of magnitude faster than SAM growth on a bare surface (76). In practice, these conditions were obtained by scanning the AFM probe slowly so that the regrowth was faster than the shaving and the dimension of the shaved patch was related to the size of the AFM probe tip. Under these conditions, there were no observable defects or scars at the boundary of the grafted features. The authors suggested that, in the spatially conned environment, the usual sequence of growth was altered. Instead of progressing through a lying-down phase to a standing-up phase, the molecules adsorbed directly in a standing-up conformation. The Mirkin group (77, 78) reported a technique they labeled dip-pen nanolithography that involves casting a layer of thiol adsorbate molecules onto an AFM tip and subsequently transferring these molecules to a gold substrate by scanning the tip under ambient conditions. The authors presumed that the thiols are transported to the surface via a water capillary bridge formed between the tip and surface. Although originally used to write SAM features on a bare substrate, this technique has recently been combined with the nanoshaving technique to create features within existing SAMs (79).

CONCLUSIONS
Although kinetic data on SAM growth have been collected for the past decade, we are only beginning to develop a reliable qualitative picture of the process, and we are far from a complete quantitative understanding. Given the current awareness regarding the complexity of the growth process (largely due to the application of scanning probe microscopy), that is, multiple 2D phases forming sequentially via nucleation and growth processes, it is now clear why techniques that do not discriminate between different surface species are insufcient to characterize the growth mechanisms and kinetics. Additionally, clear evidence now exists that quenching the SAM growth process during growth often results in modication

SELF-ASSEMBLED MONOLAYER FORMATION

135

Annu. Rev. Phys. Chem. 2001.52:107-137. Downloaded from www.annualreviews.org by Universite Toulouse 3 -Paul Sabatier on 09/16/10. For personal use only.

of the SAM structure, coverage, and morphology. This reduces our condence in results obtained by techniques that cannot be used in situ, under actual growth conditions. The paradigm that is emerging for SAM growth is a fascinating one, combining aspects of surfactant science, epitaxial growth, and nonequilibrium thermodynamics in two dimensions. Qualitative pictures of the growth process are now reasonably well established for several important systems. However, few of the mechanisms are understood at a quantitative level. The temperature dependence of a proposed 2D-phase diagram has been explored for alkylsilane SAMs. The details of island nucleation and growth have been touched on for alkylphosphonic acid SAMs. The basic issues of solvent and chain length effects, etc, have been addressed in numerous studies for thiol SAMs; however, the discrepancies in the literature are dramatic on these issues, both quantitative and qualitative, and essentially unexplainable. A few studies have concerned themselves with the energetics of adsorption and desorption, but, as yet, no consistency has emerged. These are complicated and rich systems with a variety of interactions of the same order in strength, that is, adsorbate-solvent, adsorbate-adsorbate, adsorbate-substrate, etc. Although variable solvent and chain length studies have the ability to address the rst and second types of interaction (at least in a limited way), there has been little effort to probe the ways in which the adsorbate-substrate interaction affects the growth process or the effects that qualitatively different types of intermolecular interactions might have. Because SAMs are being proposed as a route for surface modication in increasing numbers of applications, involving a greater variety of substrates and adsorbate chemical functionality, these issues will become increasingly important. ACKNOWLEDGMENTS Thanks go to Gang-Yu Liu, Rosina Georgiadis, and Roya Maboudian (and their coworkers) for contributing gures to this manuscript and to Chad Taylor for his careful and critical reading of this manuscript. The author is grateful for support from the National Science Foundation (award 9980250).
Visit the Annual Reviews home page at www.AnnualReviews.org

LITERATURE CITED
1. Bigelow WC, Pickett DL, Zisman WA. 1946. J. Colloid Interface Sci. 1:513 2. Nuzzo RG, Allara DL. 1983. J. Am. Chem. Soc. 105:105 3. Maoz R, Sagiv J. 1984. J. Colloid Interface Sci. 100:465 4. Poirier GE. 1997. Chem. Rev. 97:1117 27 5. Ulman A. 1996. Chem. Rev. 96:153354 6. Ulman A. 1991. An Introduction to Ultrathin Organic Films. Boston, MA: Academic 7. Adamson AW, Gast AP. 1997. Physical Chemistry of Surfaces. New York: Wiley Interscience, Wiley & Sons 8. Chang C-H, Franses EI. 1995. Colloids Surf. 100:145

136

SCHWARTZ 30. Truong KD, Rowntree PA. 1996. J. Phys. Chem. 100:1991726 31. Terrill RH, Tanzer TA, Bohn PW. 1998. Langmuir 14:84554 32. Bensebaa F, Voicu R, Huron L, Ellis TH, Kruus E. 1997. Langmuir 13:533540 33. Himmelhaus M, Eisert F, Buck M, Grunze M. 2000. J. Phys. Chem. B 104:576 84 34. Humbert C, Buck M, Calderone A, Vigneron JP, Meunier V, et al. 1999. Phys. Status Solidi A. 175:12936 35. Yamada R, Uosaki K. 1997. Langmuir 13:521821 36. Yamada R, Uosaki K. 1998. Langmuir 14:85561 37. Xu S, Cruchon-Dupeyrat SJN, Garno JC, Liu GY, Jennings GK, et al. 1998. J. Chem. Phys. 108:500212 38. Tamada K, Hara M, Sasabe H, Knoll W. 1997. Langmuir 13:155866 39. Wasserman SR, Whitesides GM, Tidswell IM, Ocko BM, Pershan PS, Axe JD. 1989. J. Am. Chem. Soc. 111:585261 40. Tidswell IM, Ocko BM, Pershan PS, Wasserman SR, Whitesides GM, Axe JD. 1990. Phys. Rev. B 41:111128 41. Schwartz DK, Steinberg S, Israelachvili J, Zasadzinski JAN. 1992. Phys. Rev. Lett. 69:335457 42. Bierbaum K, Grunze M, Baski AA, Chi LF, Schrepp W, Fuchs H. 1995. Langmuir 11:214350 43. Zhao XL, Kopelman R. 1996. J. Phys. Chem. 100:1101418 44. Kropman BL, Blank DHA, Rogalla H. 1998. Thin Solid Films 329:18590 45. Brzoska JB, Shahidzadeh N, Rondelez F. 1992. Nature 360:71921 46. Parikh AN, Allara DL, Azouz IB, Rondelez F. 1994. J. Phys. Chem. 98:75777590 47. Carraro C, Yauw OW, Sung MM, Maboudian R. 1998. J. Phys. Chem. B 102:444145 48. Goldmann M, Davidovits JV, Silberzan P. 1998. Thin Solid Films 329:16671 49. Sung MM, Carraro C, Yauw OW, Kim

9. Zhang ZY, Lagally MG. 1997. Science 276:37783 10. Kaganer V, Mohwald H, Dutta P. 1999. Rev. Mod. Phys. 71(3):779819 11. Poirier GE, Pylant ED. 1996. Science 272:114548 12. Schreiber F, Eberhardt A, Schwartz P, Wetterer SM, Lavrich DJ, et al. 1998. Phys. Rev. B 57:1247681 13. Eberhardt A, Fenter P, Eisenberger P. 1998. Surf. Sci. 397:L28590 14. Schwartz P, Schreiber F, Eisenberger P, Scoles G. 1999. Surf. Sci. 423:20824 15. Bain CD, Troughton EB, Tao Y, Evall J, Whitesides GM, Nuzzo RG. 1989. J. Am. Chem. Soc. 111:32125 16. Sun L, Crooks RM. 1991. J. Electrochem. Soc. 138:L2325 17. Shimazu K, Yagi I, Sato Y, Uosaki K. 1992. Langmuir 8:138587 18. Kim Y-T, McCarley RL, Bard AJ. 1993. Langmuir 8:194144 19. Schneider TW, Buttry DA. 1993. J. Am. Chem. Soc. 115:1239197 20. Frub se C, Doblhofer K. 1995. J. Chem. o Soc. Faraday Trans. 91:194953 21. DeBono RF, Loucks GD, Dellamanna D, Krull UJ. 1996. Can. J. Chem. 74:67788 22. Peterlinz KA, Georgiadis R. 1996. Langmuir 12:473140 23. Shao HB, Yu HZ, Cheng GJ, Zhang HL, Liu ZF. 1998. Ber. Bunsenges. Phys. Chem. 102:11117 24. Buck M, Grunze M, Eisert F, Fischer J, Trager F. 1992. J. Vac. Sci. Technol. A 10:92629 25. Karpovich DS, Blanchard GJ. 1994. Langmuir 10:331522 26. Schessler HM, Karpovich DS, Blanchard GJ. 1996. J. Am. Chem. Soc. 118:9645 51 27. Dannenberger O, Buck M, Grunze M. 1999. J. Phys. Chem. B 103:220213 28. Kawasaki M, Sato T, Tanaka T, Takao K. 2000. Langmuir 16:171928 29. H hner G, Woll C, Buck M, Grunze M. a 1993. Langmuir 9:195558

Annu. Rev. Phys. Chem. 2001.52:107-137. Downloaded from www.annualreviews.org by Universite Toulouse 3 -Paul Sabatier on 09/16/10. For personal use only.

SELF-ASSEMBLED MONOLAYER FORMATION Y, Maboudian R. 2000. J. Phys. Chem. B 104:155659 Vallant T, Brunner H, Mayer U, Hoffmann H, Leitner T, et al. 1998. J. Phys. Chem. B 102:719097 Vallant T, Kattner J, Brunner H, Mayer U, Hoffmann H. 1999. Langmuir 15:533946 Brunner H, Vallant T, Mayer U, Hoffmann H, Basnar B, et al. 1999. Langmuir 15:1899901 Resch R, Grasserbauer M, Friedbacher G, Vallant T, Brunner H, et al. 1999. Appl. Surf. Sci. 140:16875 Richter AG, Durbin MK, Yu C-J, Dutta P. 1998. Langmuir 14:598083 Richter AG, Yu CJ, Datta A, Kmetko J, Dutta P. 2000. Phys. Rev. E 61:60715 Woodward JT, Ulman A, Schwartz DK. 1996. Langmuir 12:362629 Woodward JT, Schwartz DK. 1996. J. Am. Chem. Soc. 118:786162 Woodward JT, Doudevski I, Sikes HD, Schwartz DK. 1997. J. Phys. Chem. B 101:753541 Doudevski I, Hayes WA, Schwartz DK. 1998. Phys. Rev. Lett. 81:492730 Doudevski I, Schwartz DK. 1999. Phys. Rev. B 60:1417 Hayes WA, Schwartz DK. 1998. Langmuir 14:591317 Cassie AB. 1952. Discuss. Faraday Soc. 75:5041 Porter MD, Bright TB, Allara DL, Chidsey

137

50.

64. 65. 66. 67. 68. 69. 70. 71. 72. 73.

51. 52.

Annu. Rev. Phys. Chem. 2001.52:107-137. Downloaded from www.annualreviews.org by Universite Toulouse 3 -Paul Sabatier on 09/16/10. For personal use only.

53.

54. 55. 56. 57. 58.

74. 75. 76. 77. 78. 79.

59. 60. 61. 62. 63.

CED. 1987. J. Am. Chem. Soc. 109:3559 68 Dannenberger O, Wolff JJ, Buck M. 1998. Langmuir 14:467982 Jung LS, Campbell CT. 2000. Phys. Rev. Lett. 84:516467 Cheng SS, Scherson DA, Sukenik CN. 1992. J. Am. Chem. Soc. 1114:54365437 Chen SH, Frank CW. 1989. Langmuir 5:97887 Amar JG, Family F, Lam PM. 1994. Phys. Rev. B 50:878197 Tang LH. 1993. J. Phys. I 3:93550 Evans JW, Bartz JA, Sanders DE. 1986. Phys. Rev. A 34:143448 Amar JG, Family F. 1996. Thin Solid Films 272:20822 Amar JG, Family F. 1995. Phys. Rev. Lett. 74:206669 Zinke-Allmang M, Feldman LC, Grabow MH. 1992. Surf. Sci. Rep. 16:377 463 Xu S, Liu GY. 1997. Langmuir 13:12729 Xu S, Miller S, Laibinis PE, Liu GY. 1999. Langmuir 15:724451 Xu S, Laibinis PE, Liu GY. 1998. J. Am. Chem. Soc. 120:935661 Piner RD, Zhu J, Xu F, Hong S, Mirkin CA. 1999. Science 283:66163 Hong SH, Zhu J, Mirkin CA. 1999. Langmuir 15:7897900 Amro NA, Xu S, Liu GY. 2000. Langmuir 16:30069

Annual Review of Physical Chemistry Volume 52, 2001

CONTENTS
A Free Radical, Alan Carrington State-to-State Chemical Reaction Dynamics in Polyatomic Systems: Case Studies, James J Valentini Recent Progress in Infrared Absorption Techniques for Elementary GasPhase Reaction Kinetics, Craig A Taatjes, John F Hershberger Surface Biology of DNA by Atomic Force Microscopy, Helen G Hansma On the Characteristics of Migration of Oligomeric DNA in Polyacrylamide Gels and in Free Solution, Udayan Mohanty, Larry McLaughlin Mechanisms and Kinetics of Self-Assembled Monolayer Formation, Daniel K Schwartz Crossed-Beam Studies of Neutral Reactions: State-Specific Differential Cross Sections, Kopin Liu Coincidence Spectroscopy, Robert E Continetti Spectroscopy and Hot Electron Relaxation Dynamics in Semiconductor Quantum Wells and Quantum Dots, Arthur J Nozik Ratiometric Single-Molecule Studies of Freely Diffusing Biomolecules, Ashok A Deniz, Ted A Laurence, Maxime Dahan, Daniel S Chemla, Peter G Schultz, Shimon Weiss Time-Resolved Photoelectron Spectroscopy of Molecules and Clusters, Daniel M Neumark Pulsed EPR Spectroscopy: Biological Applications, Thomas Prisner, Martin Rohrer, Fraser MacMillan Fast Protein Dynamics Probed with Infrared Vibrational Echo Experiments, Michael D Fayer Structure and Bonding of Molecules at Aqueous Surfaces, GL Richmond Light Emitting Electrochemical Processes, Neal R Armstrong, R Mark Wightman, Erin M Gross Reactions and Thermochemistry of Small Transition Metal Cluster Ions, PB Armentrout Spin-1/2 and Beyond: A Perspective in Solid State NMR Spectroscopy, Lucio Frydman From Folding Theories to Folding Proteins: A Review and Assessment of Simulation Studies of Protein Folding and Unfolding, Joan-Emma Shea, Charles L Brooks III Polymer Adsorption-Driven Self-Assembly of Nanostructures, Arup K Chakraborty, Aaron J Golumbfskie Biomolecular Solid State NMR: Advances in Structural Methodology and Applications to Peptide and Protein Fibrils, Robert Tycko Photofragment Translational Spectroscopy of Weakly Bound Complexes: Probing the Interfragment Correlated Final State Distributions, L Oudejans, RE Miller Coherent Nonlinear Spectroscopy: From Femtosecond Dynamics to Control, Marcos Dantus Electron Transmission through Molecules and Molecular Interfaces, Abraham Nitzan 1 15 41 71 93 107 139 165 193 233 255 279 315 357 391 423 463 499 537 575

Annu. Rev. Phys. Chem. 2001.52:107-137. Downloaded from www.annualreviews.org by Universite Toulouse 3 -Paul Sabatier on 09/16/10. For personal use only.

607 639 681

Early Events in RNA Folding, D Thirumalai, Namkyung Lee, Sarah A Woodson, DK Klimov Laser-Induced Population Transfer by Adiabatic Passage Techniques, Nikolay V Vitanov, Thomas Halfmann, Bruce W Shore, Klaas Bergmann The Dynamics of "Stretched Molecules": Experimental Studies of Highly Vibrationally Excited Molecules with Stimulated Emission Pumping, Michelle Silva, Rienk Jongma, Robert W Field, Alec M Wodtke

751 763

811

Annu. Rev. Phys. Chem. 2001.52:107-137. Downloaded from www.annualreviews.org by Universite Toulouse 3 -Paul Sabatier on 09/16/10. For personal use only.

You might also like