Molecular Structure of Poly(methyl methacrylate) Surface II: Effect of

Oct 13, 2014 - Chains of atactic (atac-), isotactic (iso-), and syndiotactic (syn-) PMMA were constructed in the Discover module of Materials Studio 6...
0 downloads 16 Views 12MB Size
Subscriber access provided by Northeastern University Libraries

Article

Molecular Structure of Poly(methyl methacrylate) Surface II: Effect of Stereoregularity Examined through All-atom Molecular Dynamics Kshitij C Jha, He Zhu, Ali Dhinojwala, and Mesfin Tsige Langmuir, Just Accepted Manuscript • Publication Date (Web): 13 Oct 2014 Downloaded from http://pubs.acs.org on October 14, 2014

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Langmuir is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Molecular Structure of Poly(methyl methacrylate) Surface II: Effect of Stereoregularity Examined through All-atom Molecular Dynamics Kshitij C. Jha*, He Zhu, Ali Dhinojwala, and Mesfin Tsige∗ Department of Polymer Science, The University of Akron, Akron, Ohio 44325 E-mail: [email protected],[email protected]

Abstract Utilizing all-atom molecular dynamics (MD), we have analysed the effect of tacticity and temperature on the surface structure of poly(methyl methacrylate) (PMMA) at the polymervacuum interface. We quantify these effects primarily through orientation, measured as tilt to the surface normal, and the surface number densities of the α-methyl, ester-methyl, carbonyl, and backbone methylene groups. Molecular structure at the surface is a complex interplay between orientation and number densities, and is challenging to capture through sum frequency generation (SFG) spectroscopy alone. Independent quantification of number density and orientation of chemical groups through all-atom MD presents a comprehensive model of stereoregular PMMA at the surface. SFG analysis presented in part I of this joint publication measures orientation of molecules that are in agreement with MD results. We observe the ester-methyl groups as preferentially oriented, irrespective of tacticity, followed by the αmethyl and carbonyl groups. SFG spectroscopy also points to the ester-methyl being dominant at the surface. The backbone methylene groups show a very broad angular distribution and lie ∗ To

whom correspondence should be addressed

1 ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

along the surface plane. The surface number density ratios of ester-methyl to α-methyl groups show syndiotactic PMMA with the lowest value. Isotactic PMMA has the highest ratios of ester- to α-methyl. These subtle trends in relative angular orientation and number densities that influence the variation of surface structure with tacticity are highlighted in the present paper. A more planar conformation of the syndiotactic PMMA along the surface (x-y plane) can be visualized through the trajectories from all-atom MD. Results from conformation tensor calculations for chains with any of their segments contributing to the surface validate the visual observation.

Introduction Surface energy plays a critical role in determining functionality of polymers whether it be utilization as commodity plastics or for high end composite applications such as biomaterials. Wetting, adhesion, segregation, specific binding and toxicity are all controlled in some measure by the surface energy. The proliferation in different end usage of PMMA from bone materials, 1 and MEMS 2 to pharmaceutical systems 3,4 makes the study of its surface behavior not only a basic scientific problem of interest, but also one with immediate implications in material design. In contrast to the recent mining of surface properties for various applications, analyses of PMMA so far have focused largely on its bulk properties as a pure polymer or barrier properties in composites and blends. 5–15 A plethora of surface characterization methods for polymers are available, each with its own set of challenges. The most popular of these methods remains the traditional contact angle measurement preferred in part due to its simplicity. Loss of specific chemical information, and additional complications in interpreting data that may also reflect molecular rearrangement of the polymer surface upon contact with water, limits surface characterization through contact angle. For analysis in vacuum, x-ray photoelectron spectroscopy (XPS) is an attractive tool that offers chemical information on surface groups. However, small shifts in carbon binding energies serve as indicators of various hydrocarbon groups and their differentiation is non-trivial. Sampling techniques 2 ACS Paragon Plus Environment

Page 2 of 28

Page 3 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

such as attenuated-total-reflectance (ATR) infrared, while providing surface molecular information, analyse thicknesses that are larger than those affecting interfacial properties such as wetting and adhesion. Infrared-visible SFG has been widely used as a method of studying surfaces and interfaces. We have previously employed such analysis to probe the molecular structure of both polystyrene (PS) 16 and PS/comb-polymer 17 surfaces. The behavior of PMMA at polymer, water, and air interfaces has also been investigated by a number of groups using SFG. 18–29 However, in these studies, conflicting accounts exist on the orientation of chemical groups for PMMA, and the effect of annealing on changes in surface structure especially in the work by Wang et al. 30 and Tateishi et al., 27 both of whom utilized SFG for characterization of the PMMA surface. Experimentally, there is ∼5-10 cm−1 resolution 31–34 between the α- and ester-methyl groups, and even systematic deuteration studies have given conflicting accounts of assignment of peaks 20,29,30 and the manner in which the chemical groups orient themselves. 31,33–35 This has necessitated the need for a careful delineation of the molecular structure at the surface of PMMA by combining SFG with models that complement the information available through spectroscopy. There is, thus, a clear case to be made for combining MD and SFG in order to understand the surface number densities and orientation of chemical groups such as α-methyl, ester-methyl, carbonyl, and the backbone methylene groups as surface applications become more common, and the segregation and wetting properties of the polymer gain increased interest from the polymer community. In the first part of this joint publication, hereafter referred to as paper I, 36 the molecular structure of PMMA at the air interface has been systematically analysed using surface sensitive SFG in conjunction with MD and ab initio models to aid in understanding the SFG results. The dominance of ester-methyl at the surface, irrespective of tacticity and temperature is captured by SFG analysis. Spectra computed from orientation distribution obtained from all-atom MD match experimental spectra, providing a validation for the surface structure model hypothesized in the present study. Through this combination of all-atom MD simulations and SFG we propose a model of chem-

3 ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ical group orientation that shows both α- and ester-methyl, and the carbonyl groups coming to the surface with the backbone methylene lying flat along the surface plane with broad distribution in orientation. The relative number density of the two methyl groups varies with tacticity and temperature, and these changes are captured through orientation and number density distributions obtained from all-atom MD simulations. The change in conformation at the surface with tacticity can also be visually seen through inspection of the trajectories. A sample snapshot for each tacticity at 560 K is shown in Figure 1.

Figure 1: Chain orientation for different tacticities of PMMA at 560 K. We observe the effects of tacticity on the orientation of the chain at the air interface. Clearly, syndiotactic PMMA shows the most propensity for the chains to lie parallel to the surface. As observed in Figure 1, the overall chain orientation is a strong function of tacticity and this may be a result of the packing of the chains in the bulk. We have quantified the anisotropy through computation of conformation tensor of the chains at the surface as a function of tacticity. Having analysed the molecular structure of PMMA at the surface through SFG in paper I, 36 we provide further insight by analysing the subtle differences in interfacial structure with tacticity through all-atom MD. In the current study, we find the number density ratios of the ester-methyl to the α-methyl and carbonyl groups to be tacticity dependent. We also find the relative orientation of the different chemical groups, especially the α- and ester-methyls to be tacticity dependent.

4 ACS Paragon Plus Environment

Page 4 of 28

Page 5 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Methods Force Field Parameters The interfacial behavior and dynamics of PMMA, along-with local relaxation and activation energies have been validated through MD simulations using Optimized Potentials for Liquid Simulations (OPLS) force field. 15,37 OPLS has also been used to study the behavior of PMMA at graphene interfaces. 38 The use of class I force fields such as OPLS has the additional advantage of affording equilibration over longer periods of time and has been used in the present study.

Simulation Details Chains of atactic (atac-), isotactic (iso-) and syndiotactic (syn-) PMMA were constructed in the Discover module of Materials Studio 6.0 from Accelrys Inc. and randomly placed in a box. 39 The bulk of the polymer, thus constructed, was equilibrated at 600 K for 5 ns for each tacticity in an NPT ensemble with periodic boundary conditions in all directions. The initial equilibration generates a melt structure that is energetically stabilized with equilibrated bulk density at that temperature. Chain length of 40-mers was chosen because it balances the objectives of being long enough to distinguish between the tacticities while also being computationally efficient. To generate the film from bulk, periodicity was maintained in the x and y directions while periodicity in the z direction was removed and was kept open to vacuum. Details of the method have been discussed in our previous work on polymer films. 40,41 The box dimensions and the film thickness were chosen so that no chain interacts with both the top and the bottom surface and a chain can fully adsorb at the surface. The thickness of the film for the different structures was initially about 90 Å along the z direction, with approximately 74 Å on both sides in the x-y plane. The total number of chains that satisfy these conditions for the system was found to be 64 for a chain length of 40-mers. MD simulations were also done for systems with a wider surface area (100x100x100 Å3 ) and double the thickness (74x74x180 Å3 ) at a few selected temperatures. The results were similar to the results obtained using the smaller 5 ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

system, which is used for all current analyses. The large time needed for equilibration of PMMA has been mentioned as an issue in previous studies, 5 and we found that a minimum of 100 ns at a given temperature was needed so that the ratios of the different chemical groups at the interface equilibrate to an average value that does not change with time. This is checked through computation of the ratios of the number density of chemical groups at different times of the simulation, as also their orientation distribution. Both the ratios and the orientation distributions stabilize close to 100 ns. For each tacticity, the PMMA film was cooled from 600 K to 300 K at the rate of 20 K/2 ns in the NVT ensemble. Selected temperatures of 500 K, 520 K, 540 K, and 560 K were run for longer simulation times than 100 ns (NVT), which would bring the total simulation time for a given tacticity to be greater than 0.4 µs. In running these simulations we employed particle-particle particle-mesh (PPPM) Ewald summation with 10−5 accuracy and a 12 Å cutoff for van der Waals summation using 126 potential. Since no chain leaves the interface, the choice of a longer cutoff for van der Waals is sufficient to capture the interfacial structure. All simulations were carried out using LAMMPS code. 42 Utilizing these simulation time protocols, surface number density ratios of chemical groups were seen to equilibrate to an average value and their variance with tacticity and temperature gave a measure of the surface structure of PMMA. Our discussion will mainly focus on 500 K and 560 K, which represent the lower and higher ends of the temperature range.

Sum Frequency Generation (SFG) Spectroscopy To measure the response of the molecules at the surface, SFG mixes a visible laser beam with a tunable infrared (IR) beam. Approximation of the electric dipole yields only signal through ordered species at the surface. Enhancement of the signal occurs through overlap of the stretching vibrations of the surface molecules with the incident IR. The SFG setup and apparatus are detailed in paper I 36 alongside discussion of its particular utility to resolve the ordering and orientation of chemical groups at the PMMA-air interface. Spectra were obtained for syn- and iso-PMMA for solid and melt state (above Tg for syn-PMMA and above Tm for iso-PMMA). Through angular 6 ACS Paragon Plus Environment

Page 6 of 28

Page 7 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

distribution obtained from all-atom MD, SFG spectra were computed that match experimental observations.

Results and Discussion Characterization of the surface, followed by analyses of the orientation and number densities at the surface of the four chemical groups (ester-methyl, α-methyl, backbone methylene, and carbonyl) and their dependence on stereoregularity is discussed in the following sections. We conclude with a discussion on the effect of tacticity on chain conformation at the surface.

Mass Density Profile Mass density represents the most straightforward way to characterize the surface. We computed the mass density profile using partition bins of 1 Å in the z direction of the simulation cell. The masses of the atoms were summed in the partition volume, and density was calculated and averaged over length of the trajectory at a given temperature.

Figure 2: Density distribution along the PMMA film in the Z-direction at 560 K with z=0 in the middle of the film. Left shows mass density distribution of the films. Right is the zoomed in distribution for the upper 25 Å of the films to show density variation with tacticity. A similar behavior is seen for temperatures from 500 K- 560 K.

7 ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 2 shows time averaged mass density plots for the three tacticities at 560 K. The mass densities show two symmetric and smooth polymer-vacuum interfaces for all tacticities with the nature of the curve at the interfaces exhibiting tan hyperbolic behavior. Same behavior is observed for temperatures from 500 K- 560 K. The average densities in the middle of the film are within 5% of experimental values 43 and decreased with increasing temperature, as expected. The film thickness of iso-PMMA is slightly smaller as can be seen in the left panel of Figure 2. This means iso-PMMA has higher density, which has also been observed experimentally and in simulations. 5 Additionally, both Soldera et al. 5 and Lu et al. 7 based on density and free volume computations respectively argue for a more compact packing of iso-PMMA.

Orientation at the Surface Utilizing the vectors for orientation shown in Figure 3, distribution plots for were computed (Figure 4) to analyze the orientation of different chemical groups. All angles(θ ) are with respect to the z-direction which is normal to the film surface. Figure 4 also allows us to characterize the surface. The point where the values start to show a monotonic increase (top part of the film) or decrease (bottom part of the film) can be termed as preferential onset of orientation. The surface is defined as the region beyond the preferential onset of orientation. We analyzed the preferential onset of orientation for all three tacticities and temperatures and found that utilizing a unique preferential onset of orientation as the starting point of the surface yields the same trends as utilizing the average of the top and bottom 1 nm for the film. Computation of orientation distributions for 1.5 nm and 2 nm, reveal recovery to bulk behavior with loss of the distinctions in chemical group orientation distributions. The recovery to bulk behaviour happens very quickly after 1 nm, which further validates our choice of interface. Figure S1 in the Supplementary Information shows the angular distribution in the middle of the film (at 500 K) which reflects bulk behavior, and beyond 1 nm we observed a similar distribution. The nature of the distribution is similar for 560 K, and shows isotropic behavior with no differentiation in chemical group orientation, as expected. 8 ACS Paragon Plus Environment

Page 8 of 28

Page 9 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Figure 3: Vectors used for computation of orientation distribution for PMMA film. Part of a PMMA polymer chain is shown. Angles are measured for the vectors with respect to the Zdirection that is normal to the surface.

Figure 4: Average cosθ values as a measure of the orientation for the three tacticities and two temperatures of 500 K and 560 K. Values are shown for methylene, α-methyl, ester-methyl and carbonyl groups.

9 ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The ester-methyl group shows consistently highest values of at the surface irrespective of the tacticity and temperature followed by α-methyl group (Figure 4). Also of interest is the methylene group, which shows least orientation of all the chemical groups (α-methyl, estermethyl and carbonyl), indicating that they are lying mostly along the horizontal plane. This will be revisited in the orientational analysis for the top 1 nm as a function of angle. For a relative comparison across tacticities we plotted the values for ester-methyl at two temperatures for the three tacticities (Figure 5). The results are very similar indicating invariance of orientation for the ester-methyl groups with tacticity and temperature.

Figure 5: Average cosθ for ester-methyl for the three tacticities at temperatures of 500 K and 560 K. To estimate the spatial orientation at the surface, the angular distribution is plotted for the top 1 nm of the film (Figure 6). We observe ester-methyl having the highest peaks, irrespective of tacticity. α-methyl has a comparable peak to ester-methyl for syn-PMMA, where we also observe a shift of the methylene group away from all the other chemical groups. A broad distribution of the methylene group is observed for all tacticities.

10 ACS Paragon Plus Environment

Page 10 of 28

Page 11 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Figure 6: Angular distribution for the top 1 nm of the film for three different tacticities and two temperatures of 500 K and 560 K. Values for α-methyl, ester-methyl, methylene, and carbonyl groups are shown. The errors are within 5 %.

To compare the peak for different chemical groups across tacticities we plotted the same distribution for each chemical group at two temperatures of 500 K and 560 K (Figure 7, and Figure S2 in Supplementary Information). For a given chemical group, Figure 7 is indicative of the effect of tacticity which could be an important design principle for acrylic surfaces. We observe synPMMA as having the highest peak of the three tacticities for α-methyl, while ester-methyl has the highest peaks for iso- and atac-PMMA. Carbonyl peaks are invariant for all tacticities, while there is a shift in the methylene peak to the right for syn-PMMA. The results from Figure 6 and Figure 7 lead us to conclude that the ester-methyl has the most well defined and narrow distribution while methylene orientation distributions are broad. The two methyl groups have similar distributions for syn-PMMA. There is also a shift towards the surface plane for the methylene group for syn-PMMA. Since the errors in orientation distribution are within 5 %, these distinctions are beyond the error bounds.

11 ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 28

Figure 7: Angular distribution for the top 1 nm of the film for three different tacticities and for αmethyl, ester-methyl, methylene, and carbonyl groups at 560 K. The errors are within 5 %. Similar distributions are observed for 500 K and are provided in the Supplementary Information, Figure S1.

The peaks are the most probable values. However, also useful are the average values for the angles, which were computed as a weighted average for the angular distribution. These values for the different groups are shown in Table 1. Table 1: Average Angle for Different Chemical Groups for Different Temperatures and Tacticities. Error is less than 0.1o for all values. Tacticity Isotactic

Temperature (K) 500 560

Ester-Methyl α-Methyl 67.6 80.4 68.0 82.2

Methylene 88.2 85.9

Carbonyl 76.1 78.8

Syndiotactic

500 560

73.9 74.1

77.9 76.6

89.5 91.0

79.2 80.3

Atactic

500 560

71.5 69.1

75.9 78.7

89.6 86.3

79.3 80.3

Table 1 quantifies the peak distribution behavior, through tilt angles with respect to the surface normal, seen in Figure 6 and Figure 7. We observe that the ester-methyl orientation follows the 12 ACS Paragon Plus Environment

Page 13 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

trend: syn>atac>iso, α methyl: iso>atac ∼ syn, methylene: syn>atac>iso, and carbonyl: syn ∼ atac>iso. The lower the angle, the more the group is oriented towards the surface normal. Methylene groups orient almost along the plane, with the orientation being above the plane for atac- and iso-PMMA by 2 − 4o and below the plane by 1 − 2o for syn-PMMA. Only the iso-PMMA has greater α-methyl angle than carbonyl angle, implying that the carbonyl has the lowest angle to the surface normal for iso-PMMA. This shows the interplay between surface number density and orientation. Syndiotactic PMMA has higher relative carbonyl surface number density as will be discussed in the next section, but iso-PMMA has the carbonyl group oriented more towards the normal. Interplay of orientation and number density between chemical groups determines the surface behavior of a polymer which can be utilized as design principle for functional interfaces.

Number Density at the Surface In order to know how different chemical groups come to the surface with tacticity and temperature, it is instructive to look at the number densities. These were computed by using a slab length of 1 Å in a manner similar to computation of mass densities, by taking the number count instead of atomic masses. As seen in all distributions in Figure 8, at a certain point the number density of ester-methyl becomes greater than the number density of the α-methyl and methylene units (we call this crossover regime). This crossover regime (Figure 8) is similar to preferential onset of orientation (Figure 4) which has been defined in the section on surface orientation. Hence, average of 1 nm (top and bottom) is representative of both the crossover point and the preferential onset of orientation. The similarity between the crossover regime and the preferential onset may be specific to PMMA and needs to be tested for other polymeric systems in the future.

13 ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 8: Number density versus tacticity and temperature for the upper part of the PMMA films for 500 K and 560 K. The crossover regime, shown by a broken line in top left distribution, is defined as the point where the number density of ester-methyl group exceeds that of the α-methyl units. It gives a picture of relative values of α-methyl to ester-methyl (and methylene) on the surface. The question arises as to how best to quantify relative number density effects. One way to look at it is through visualization of trajectories. Figure 9 shows how concentration of chemical groups change as we travel from the surface to the more interior region of the film. There is uneven distribution of chemical groups around 0.8-1 nm. For example, in panel B for iso-PMMA, the spatial distribution of chemical groups can most clearly be visualized as uneven at 0.8 nm. As we go inside and approach bulk, we observe a more uniform distribution. This is consistent with our definition of surface from both surface orientation and number density distributions. The uneven distributions are a reflection of the surface roughness consisting of crests and troughs. It is to be stressed that the uneven distribution observed on the surface is not a consequence of finite size effects, as we have observed the same nature of distribution of chemical groups even when we double the system size in our simulations. We also clearly observe that the ester-methyl group is in excess at the surface compared to α-methyl. The excess is apparent in the 0.8-1 nm region, beyond which there are similar number of chemical groups for both esterand α-methyl. In the discussion of ratios that follows, the recovery to similar distribution in bulk of different chemical groups that may be depleted or in excess at the surface becomes evident. Figure 9 provides clear visualization of this change in chemical group distribution. 14 ACS Paragon Plus Environment

Page 14 of 28

Page 15 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Figure 9: Distribution of ester-methyl (in blue) and α-methyl (in yellow) for different depths at the top interface at 560 K. The carbon for the chemical groups are highlighted. A shows atac-PMMA, B shows iso-PMMA, and C shows syn-PMMA depth profiles.

Another method to quantify relative number densities is to evaluate ratios of chemical groups as a function of distance from the surface. While computing these ratios, it is to be noted that the number densities for the top or bottom 0.5 nm are extremely small and ratios of chemical groups in these regions can be misleading and unphysical. For example, we can see in the case of atacPMMA 500 K in Figure 8 that surface number densities in the 43-48 Å range are too small to yield any measurable surface characteristics. Hence, our definition of the top of the film is after this region that is statistically significant in terms of number densities. We compute the ratios of the chemical groups from the sum of number densities starting from 15 ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

the top of the surface to a given point in the film. The ratios of number density of ester-methyl groups to α-methyl groups as a function of tacticity, temperature, and distance from the top of the film are shown in Figure 10 (500 K and 560 K). In all cases, the ratio is greater than 1 at the surface, indicating that ester-methyl is in excess at the surface compared to α-methyl. The relative values of the ratio are also similar, with a decrease as much as 7% with temperature.

Figure 10: Ratio of ester-methyl to α-methyl plotted with distance from the top of the film as a function of temperature and tacticity.

Figure 11 shows the surface number density of carbonyl groups compared to ester-methyl groups for the top part of the film as a function of tacticity at 560 K. The differences can be observed to a greater extent in examining the ratios of number density of ester-methyl groups to carbonyl groups at 500 and 560 K (Figure 12). In all cases the values are higher than 1 at the surface, indicating that ester-methyl groups are present at the surface in larger extent than carbonyl groups. The overall trend is similar to what we observed for ratios of ester-methyl to α-methyl groups (Figure 10), with the ester-methyl groups dominant irrespective of temperature and tacticity. It is interesting to note that the carbonyl groups are depleted near the surface and this ratio may change for PMMA surface in contact with water. The carbonyl group attracts the most participation

16 ACS Paragon Plus Environment

Page 16 of 28

Page 17 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

from water, and the PMMA chain is said to re-orient itself to maximize water-carbonyl oxygen interaction. 44 We discussed the carbonyl group separately from the other three chemical groups (ester-methyl, α-methyl, methylene), since its hydrophilic character is greater and would impact the behavior of PMMA surfaces towards polar media and its surface concentration may play an important role in determining the interfacial surface energy.

Figure 11: Carbonyl and ester-methyl number density in the upper part of the film for different tacticities at 560 K.

Figure 12: Ratio of ester-methyl to carbonyl plotted with distance from the top of the film as a function of temperature and tacticity.

Tacticity Effects on Conformation Previous work on the stereoregularity of PMMA and its effect on the bulk glass transition 11 suggests that the mobility of the side chain, specifically for the iso-PMMA leads to increased mobility 17 ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

of the backbone and a decreased glass transition. Another theory states that syn-PMMA has increased dipole-dipole interactions between the side chains 45,46 which restrict mobility leading to higher glass transition. Higher non-bond energy for the syn-PMMA has also been reported. 47 It has been hypothesized that syn-PMMA possesses lower intra-molecular energies whereas isoPMMA has lower inter-molecular energies leading to better packing. 5 Analyses of free volume by Lu et al. 7 indicate iso-PMMA as the better packed polymer. We propose that a more favorable steric distribution of chemical groups for iso-PMMA leads to better packing and may lead to faster dynamics. Our mass density computations also reveal a denser iso-PMMA. Higher surface activity, seen in our simulations through syn-PMMA being more planar on the surface, and the relative higher concentration of α-methyl for syn-PMMA is consistent with contact angle measurements. 48,49 The glide-plane structure at the surface yields a more surface active conformation 48 which is not possible for a helical conformation that iso-PMMA prefers even in a melt state. In Figure 13 we observe that the surface orientation and number density of the chemical groups would also depend on the tacticity which is related to the packing of the chains. We observe for chains 1-6 for both polymers, that syn-PMMA tends to coil and space itself on the surface. Atactic PMMA is in between these two behaviors.

Figure 13: Orientation of chains at the surface and bulk for iso-PMMA and syn-PMMA. Only a few representative chains have been highlighted to bring out the differences.

18 ACS Paragon Plus Environment

Page 18 of 28

Page 19 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Since the orientation at the surface is markedly different than at the bulk, the patterns of the bulk may be transferred to the surface and vice versa leading to oscillations in density observed in distribution plots. Coupled with control offered by stereoregularity discussed in the current work, these could lead to tunable optical and lithographic properties as has been explored in LangmuirBlodgett films of PMMA for quite some time. 50,51 Having observed chains with varying conformations at the surface, we have quantified the anisotropy in chain alignment by calculating diagonal elements of the conformation tensor for chains with any segments that are part of the top 1 nm of the film (Table 2). Similar behavior is observed for the bottom 1 nm. The conformation tensor is defined as, 52 Cαβ = 3h

Reeα Reeβ i hR2ee i0

where

Reeα and Reeβ are the end-to-end distance components in three cartesian directions (x,y, and z) for chains with segments in the top 1 nm of the film, and hR2ee i0 is the mean squared end-to-end distance of the system in bulk at the corresponding temperature. We observe that the anisotropy in the x-y plane (Cαα 6= 1) is the highest for the syn-PMMA where both Cxx and Cyy >1, implying a larger contribution to a planar conformation in the x-y (surface) plane as observed in Figure 13. We also observe that the atac-PMMA has the largest mean-square end-to-end distance in all directions, while iso-PMMA at 560 K appears to be closest to isotropic distribution. Roughness computed for the interface gave similar values for both isoPMMA and syn-PMMA of 0.5±0.02 nm. However, it has to be noted that the roughness does not take into account the spatial distribution of the chains. Table 2: Diagonal elements of conformation tensor for chains in the top 1 nm of the film and the mean squared end-to-end distance of the system in the bulk at the corresponding temperature. Deviation for the elements is less than 5 percent. Tacticity Isotactic

Temperature (K) Cxx Cyy Czz hR2ee i0 (Å2 ) 500 0.76 1.14 0.90 1492 ± 10 560 0.92 1.14 1.01 1178 ± 12

Syndiotactic

500 560

1.15 1.33

1.51 1.87

0.73 0.64

1117 ± 9 1403 ± 12

Atactic

500 560

1.19 1.19

0.76 0.76

0.71 0.64

2996 ± 10 3809 ± 12

19 ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

It must be emphasized that all observations through all-atom simulations show subtle differences in orientation and ordering which would be very hard to characterize through empirical approaches. As expected, the spectra for syn-PMMA and iso-PMMA match at the polymer-air interface for PMMA, discussed in the first part of the joint publication. 36 Coupled with insight provided by spectroscopy, our simulations have attempted to analyse the minimal differentials in surface orientation and number densities of chemical groups, along-with their interplay, that could lead to changes in molecular structure at the surface of PMMA.

Conclusions Through a combined investigation of the effect of stereoregularity on relative number densities it was observed that the ester-methyl groups come to the surface irrespective of the tacticity and temperature. We propose a preferential onset of orientation that is very close to the crossover regime to define the surface, where quantifiables in surface orientation and number densities paint a picture of the structure of PMMA. In this regime, through analyses of ratios of ester-methyl to αmethyl and carbonyl, it is observed that the iso-PMMA has a greater degree of ester-methyl excess at the surface. In part I of the joint publication, we showed the MD results could quantitatively match the SFG spectra observed in experiments. Even though the MD and SFG results show that ester-methyl groups are present in excess, SFG was not able to resolve the subtle differences as a function of tacticity. The differences in molecular hyperpolarization in addition to the values of orientation and number density are needed to explain the SFG results. Visualization of the trajectories show a planar conformation of syn-PMMA chains at the surface. A model for PMMA at the surface is presented that shows ester-methyl having the highest orientation and backbone methylene the lowest orientation irrespective of tacticity, with the other two chemical groups (αmethyl and carbonyl) variously oriented per tacticity and temperature. Combining the surface orientation and number densities obtained from all-atom MD with SFG provides a detailed picture of the surface structure of PMMA and highlights the need for MD in combination with SFG to

20 ACS Paragon Plus Environment

Page 20 of 28

Page 21 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

understand the number density and orientation of molecules at polymer surfaces and interfaces.

Acknowledgements This work was supported by NSF Grant Nos. DMR-0847580 (M.T.), DMR-1105370 (A.D.), and DMR-1006764 (A.D.). The authors thank Dr. Ram Bhatta and Alankar Rastogi for frutiful discussions and The University of Akron for additional support.

Supporting Information Available Angular distribution for the middle of the film for three different tacticities at 500 K which reflects bulk behavior. Angular distribution for the top 1 nm of the film for three different tacticities for α-methyl, ester-methyl,methylene, and carbonyl groups at 500 K. This material is available free of charge via the Internet at http://pubs.acs.org/.

References (1) De Santis, R.; Mollica, F.; Ambrosio, L.; Nicolais, L.; Ronca, D. Dynamic mechanical behavior of PMMA based bone cements in wet environment. J. Mater. Sci. - Mater. Med. 2003, 14, 583–594. (2) Tambe, N. S.; Bhushan, B. Micro/nanotribological characterization of PDMS and PMMA used for BioMEMS/NEMS applications. Ultramicroscopy 2005, 105, 238–247. (3) Izzo, L.; Griffiths, P. C.; Nilmini, R.; King, S. M.; Wallom, K.-L.; Ferguson, E. L.; Duncan, R. Impact of polymer tacticity on the physico-chemical behaviour of polymers proposed as therapeutics. Int. J. Pharm. 2011, 408, 213–222. (4) Matsuno, H.; Nagasaka, Y.; Kurita, K.; Serizawa, T. Superior activities of enzymes physically immobilized on structurally regular poly (methyl methacrylate) surfaces. Chem. Mater. 2007, 19, 2174–2179. 21 ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(5) Soldera, A. Comparison between the glass transition temperatures of the two PMMA tacticities: A molecular dynamics simulation point of view. Macromol Symp. 1998, 133, 21–32. (6) Soldera, A. Energetic analysis of the two PMMA chain tacticities and PMA through molecular dynamics simulations. Polymer 2002, 43, 4269–4275. (7) Lu, K.-T.; Tung, K.-L. Molecular dynamics simulation study of the effect of PMMA tacticity on free volume morphology in membranes. Korean J. Chem. Eng 2005, 22, 512–520. (8) Hartmann, L.; Gorbatschow, W.; Hauwede, J.; Kremer, F. Molecular dynamics in thin films of isotactic poly(methyl methacrylate). Eur. Phys. J. E 2002, 8, 145–154. (9) Tung, K.-L.; Lu, K.-T. Effect of tacticity of {PMMA} on gas transport through membranes: {MD} and {MC} simulation studies. J. Membr. Sci. 2006, 272, 37–49. (10) Soldera, A.; Grohens, Y. Local dynamics of stereoregular PMMAs using molecular simulation. Macromolecules 2002, 35, 722–726. (11) Soldera, A.; Grohens, Y. Cooperativity in stereoregular PMMAs observed by molecular simulation. Polymer 2004, 45, 1307–1311. (12) Genix, A.-C.; Arbe, A.; Alvarez, F.; Colmenero, J.; Willner, L.; Richter, D. Dynamics of poly(ethylene oxide) in a blend with poly(methyl methacrylate): A quasielastic neutron scattering and molecular dynamics simulations study. Phys. Rev. E 2005, 72, 031808. (13) Tung, K.-L.; Lu, K.-T.; Ruaan, R.-C.; Lai, J.-Y. {MD} and {MC} simulation analyses on the effect of solvent types on accessible free volume and gas sorption in {PMMA} membranes. Desalination 2006, 192, 391–400. (14) Erber, M.; Tress, M.; Mapesa, E. U.; Serghei, A.; Eichhorn, K.-J.; Voit, B.; Kremer, F. Glassy Dynamics and Glass Transition in Thin Polymer Layers of PMMA Deposited on Different Substrates. Macromolecules 2010, 43, 7729–7733.

22 ACS Paragon Plus Environment

Page 22 of 28

Page 23 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

(15) Chen, C.; Maranas, J. K.; García-Sakai, V. Local dynamics of syndiotactic poly (methyl methacrylate) using molecular dynamics simulation. Macromolecules 2006, 39, 9630–9640. (16) Gautam, K.; Schwab, A.; Dhinojwala, A.; Zhang, D.; Dougal, S.; Yeganeh, M. Molecular structure of polystyrene at air/polymer and solid/polymer interfaces. Phys. Rev. Lett. 2000, 85, 3854. (17) Harp, G. P.; Rangwalla, H.; Yeganeh, M. S.; Dhinojwala, A. Infrared-visible sum frequency generation spectroscopic study of molecular orientation at polystyrene/comb-polymer interfaces. J. Am. Chem. Soc. 2003, 125, 11283–11290. (18) Wang, J.; Paszti, Z.; Even, M. A.; Chen, Z. Measuring Polymer Surface Ordering Differences in Air and Water by Sum Frequency Generation Vibrational Spectroscopy. J. Am. Chem. Soc. 2002, 124, 7016–7023. (19) Liu, Y.; Messmer, M. C. Molecular Orientation at the Interface of Polystyrene/Poly(methyl ˘ ’ Evidence from Sum-Frequency Spectroscopy. J. Am. Chem. Soc. methacrylate) Blends:âAL 2002, 124, 9714–9715. (20) Liu, Y.; Messmer, M. C. Surface Structures and Segregation of Polystyrene/Poly(methyl methacrylate) Blends Studied by Sum-Frequency (SF) Spectroscopy. J. Phys. Chem. B 2003, 107, 9774–9779. (21) Miyamae, T.; Nozoye, H. Morphology and chemical structure of poly(methyl methacrylate) surfaces and interfaces: restructuring behavior induced by the deposition of SiO2. Surf. Sci. 2003, 532, 1045–1050. (22) Kweskin, S. J.; Komvopoulos, K.; Somorjai, G. A. Molecular Restructuring at Poly(n-butyl methacrylate) and Poly(methyl methacrylate) Surfaces Due to Compression by a Sapphire Prism Studied by Infraredâ´LŠVisible Sum Frequency Generation Vibrational Spectroscopy. Langmuir 2005, 21, 3647–3652.

23 ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(23) Clarke, M. L.; Chen, C.; Wang, J.; Chen, Z. Molecular Level Structures of Poly(n-alkyl methacrylate)s with Different Side Chain Lengths at the Polymer/Air and Polymer/Water Interfaces. Langmuir 2006, 22, 8800–8806. (24) Li, Q.; Hua, R.; Cheah, I. J.; Chou, K. C. Surface Structure Relaxation of Poly(methyl methacrylate). J. Phys. Chem. B 2008, 112, 694–697. (25) Wang, X.; Ni, H.; Xue, D.; Wang, X.; Feng, R.-r.; Wang, H.-f. Solvent effect on the film formation and the stability of the surface properties of poly(methyl methacrylate) end-capped with fluorinated units. J. Colloid Interface Sci. 2008, 321, 373–383. (26) Lu, X.; Shephard, N.; Han, J.; Xue, G.; Chen, Z. Probing Molecular Structures of Polymer/Metal Interfaces by Sum Frequency Generation Vibrational Spectroscopy. Macromolecules 2008, 41, 8770–8777. (27) Tateishi, Y.; Kai, N.; Noguchi, H.; Uosaki, K.; Nagamura, T.; Tanaka, K. Local conformation of poly(methyl methacrylate) at nitrogen and water interfaces. Polym. Chem. 2010, 1, 303– 311. (28) Jena, K. C.; Covert, P. A.; Hall, S. A.; Hore, D. K. Absolute Orientation of Ester Side Chains on the PMMA Surface. J. Phys. Chem. C 2011, 115, 15570–15574. (29) Horinouchi, A.; Atarashi, H.; Fujii, Y.; Tanaka, K. Dynamics of Water-Induced Surface Reorganization in Poly(methyl methacrylate) Films. Macromolecules 2012, 45, 4638–4642. (30) Wang, J.; Chen, C.; Buck, S. M.; Chen, Z. Molecular chemical structure on poly (methyl methacrylate)(PMMA) surface studied by sum frequency generation (SFG) vibrational spectroscopy. J. Phys. Chem. B 2001, 105, 12118–12125. (31) Dirlikov, S.; Koenig, J. Assignment of the carbon-hydrogen stretching and bending vibrations of poly (methyl methacrylate) by selective deuteration. Appl. Spectrosc. 1979, 33, 555–561.

24 ACS Paragon Plus Environment

Page 24 of 28

Page 25 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

(32) Nagai, H. Infrared spectra of stereoregular polymethyl methacrylate. J. Appl. Polym. Sci. 1963, 7, 1697–1714. (33) Schneider, B.; tokr, J. Ã.; Schmidt, P.; Mihailov, M.; Dirlikov, S.; Peeva, N. Stretching and deformation vibrations of CH2, C(CH3) and O(CH3) groups of poly(methyl methacrylate). Polymer 1979, 20, 705–712. (34) Willis, H.; Zichy, V.; Hendra, The laser-Raman and infra-red spectra of poly (methyl methacrylate). Polymer 1969, 10, 737–746. (35) Lipschitz, I. The Vibrational Spectrum of Poly (Methyl Methacrylate): A Review. Polym.Plast. Technol. 1982, 19, 53–106. (36) Zhu, H.; Jha, K. C.; Bhatta, R. S.; Tsige, M.; Dhinojwala, A. Molecular structure of poly (methyl methacrylate) surface I: Combination of interface-sensitive infrared-visible sum frequency generation, molecular dynamics simulations, and ab initio calculations. Langmuir 2014, 30, 11609–11618. (37) Chen, C.; Depa, P.; Maranas, J. K.; Sakai, V. G. Comparison of explicit atom, united atom, and coarse-grained simulations of poly (methyl methacrylate). J. Chem. Phys. 2008, 128, 124906. (38) Rissanou, A. N.; Harmandaris, V. Structure and dynamics of poly (methyl methacrylate)/graphene systems through atomistic molecular dynamics simulations. J. Nanopart. Res. 2013, 15, 1–14. (39) Amorphous Cell from Discover Suite of Materials Studio 6.0 Accelrys Software Inc. San Diego (40) Tatek, Y. B.; Tsige, M. Structural properties of atactic polystyrene adsorbed onto solid surfaces. J. Chem. Phys 2011, 135, 174708–174708.

25 ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(41) Bekele, S.; Tsige, M. Interfacial Properties of Oxidized Polystyrene and Its Interaction with Water. Langmuir 2013, 29, 13230–13238. (42) Plimpton, S. Fast parallel algorithms for short-range molecular dynamics. J. Comput. Phys. 1995, 117, 1–19. (43) Wunderlich, W. Polymer handbook 3rd ed.Brandrup, J. and Immergut, E.H. eds.; Springer, 1989; p 77. (44) Lee, W.-J.; Chang, J.-G.; Ju, S.-P. Hydrogen-bond structure at the interfaces between water/poly (methyl methacrylate), water/poly (methacrylic acid), and water/poly (2aminoethylmethacrylamide). Langmuir 2010, 26, 12640–12647. (45) Fujii, Y.; Akabori, K.-i.; Tanaka, K.; Nagamura, T. Chain conformation effects on molecular motions at the surface of poly (methyl methacrylate) films. Polym. J. 2007, 39, 928–934. (46) Tretinnikov, O. N.; Ohta, K. Conformation-sensitive infrared bands and conformational characteristics of stereoregular poly (methyl methacrylate) s by variable-temperature FTIR spectroscopy. Macromolecules 2002, 35, 7343–7353. (47) Soldera, A.; Metatla, N. Study of the glass transition temperatures of stereoregular PMMAs using different force fields. Internet Electron J. Mol. Des. 2005, 4, 721–736. (48) Tretinnikov, O. N. Selective accumulation of functional groups at the film surfaces of stereoregular poly (methyl methacrylate) s. Langmuir 1997, 13, 2988–2992. (49) Tretinnikov, O. N.; Ohta, K. Surface segregation in stereochemically asymmetric polymer blends. Langmuir 1998, 14, 915–920. (50) Kim, J.-J.; Jung, S.-D.; Roh, H.-S.; Ha, J.-S. Molecular configuration of isotactic {PMMA} Langmuir-Blodgett films observed by scanning tunnelling microscopy. Thin Solid Films 1994, 244, 700–704.

26 ACS Paragon Plus Environment

Page 26 of 28

Page 27 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

(51) Kim, J.-J.; Jung, S.-D.; Hwang, W.-Y. Molecular conformation and application of stereoregular PMMA Langmuir-Blodgett films. ETRI J 1996, 18, 195–206. (52) Harmandaris, V. A.; Mavrantzas, V. G.; Theodorou, D. N. Atomistic molecular dynamics simulation of stress relaxation upon cessation of steady-state uniaxial elongational flow. Macromolecules 2000, 33, 8062–8076.

27 ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Table of Contents Kshitij C. Jha, He Zhu, Ali Dhinojwala, and Mesfin Tsige

28 ACS Paragon Plus Environment

Page 28 of 28