02.10.2022 Views

Main_manuscript_JCP_template

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Sample title

Understanding the Formation of Surface Relief Gratings in Azopolymers:

A Combined Molecular Dynamics and Experimental Study

Milena Merkel, 1 Amala Elizabeth, 2, 3 Marcus Böckmann, 4 Harry Mönig, 2, 3 Cornelia Denz, 1 and Nikos L.

Doltsinis 4

1) Institut für Angewandte Physik, Westfälische Wilhelms-Universität Münster, Corrensstr. 2/4, 48149 Münster,

Germany.

2) Physikalisches Institut, Westfälische Wilhelms-Universität Münster, Wilhelm-Klemm-Str. 10, 48149 Münster,

Germany.

3) Center for Nanotechnology, Heisenbergstr. 11, Münster 48149, Germany

4) Institut für Festkörpertheorie, Westfälische Wilhelms-Universität Münster and Center for Multiscale Theory & Computation,

Wilhelm-Klemm-Str. 10, 48149 Münster, Germany.

(Dated: 30 September 2022)

The formation of surface relief gratings (SRGs) in thin films of poly-disperse-orange3-methyl-methacrylate is investigated

using atomistic molecular dynamics simulations and compared to experimental results. For this purpose, the film

is illuminated with a light pattern of alternating bright and dark stripes in both cases. The simulations use a molecular

mechanics switching potential to explicitly describe the photoisomerization dynamics between the E and Z isomers of

the azo-units and take into account the orientation of the transition dipole moment with respect to the light polarisation.

Local heating and elevation of the illuminated regions with subsequent movement of molecules into the neighboring

dark regions is observed. This leads to the formation of valleys in the bright areas after re-cooling and is independent of

the polarization direction. To verify these observations experimentally, the azopolymer film is illuminated with bright

stripes of varying distance using a spatial light modulator. Atomic force microscopy images confirm that the elevated

areas correspond to the previously dark areas. In the experiment, the polarization of the incident light makes only a

small difference, since tiny bead-like structures form in the valley only when the polarization is parallel to the stripes.

I. INTRODUCTION

The light-induced modification of azo-containing polymer

films has been a vivid research field since the first reports

describing the phenomenon in the late 1990’s 1–7 . As a consequence

of the E/Z photo-isomerisation of the azobenzene

(AB) photochromic unit, a large variety of macroscopic structural

changes in azopolymer materials, such as bending of

free-standing films, genuine solid-to-liquid transformations

(’liquefication’) or the formation of surface relief gratings

(SRGs) can occur 4,8–11 .

In the latter phenomenon, trenches and hills form on the

surface of a polymer film due to effective mass transport induced

by inhomogeneous irradiation 7,12 . For this purpose,

mostly periodic interference patterns are used, but also more

sophisticated patterns, such as optical vortices or differently

polarized Bessel beams 4,12–15 . Interestingly, the resulting

SRGs are extremely diverse in terms of shape, depth or stability.

Even the position of the peaks, whether in the low or

high intensity regions of the illumination pattern, varies 7,16 .

While in most applications mass transport is seen to be directed

away from the illuminated (bright) regions, and in the

direction of the incident light polarisation, several exceptions

can be found in literature. For example, Holme et al. 17 observed

the formation of hills in the bright regions for a liquid

crystalline polyester in contrast to an amorphous peptide

oligomer showing trenches. Furthermore, a more rigid backbone

of the polyester leads to the formation of trenches instead

of hills. Yadavalli et al. 18 in turn observed a formation of

hills in either the bright or the dark regions, depending on the

strength of the interaction between the azobenzene-containing

side chains and the backbone.

Also the dependence of the direction of material movement

on polarization is controversial. For example, Karageorgiev

et al. 19 prepared two orthogonal trenches on a thin film surface

prior to illumination and found that only the one parallel

to the polarisation vector was erased. Meanwhile, previous

publications have suggested that SRG formation is independent

of polarization under high-intensity cw and pulsed

irradiation 7,20,21 .

Furthermore, while the majority of applications have the

azo-chromophore either covalently linked (with spacer units

of variant length) to the polymer backbone or incorporated

into it, the azo-unit may also be present only as an additive

to the polymer composition, with surprisingly low concentrations

down to 1% 22 . In some cases, pulsed laser experiments

have produced SRGs even in samples with absorbing, but nonisomerizing

chromophores, suggesting a thermal origin of the

structures 7,20,21,23–26 .

All things considered, these varying and partially contradicting

results may be due to the wide range of different

experimental conditions in terms of material, illumination

intensity, exposure time, wavelength, light polarization or

pattern 10,11,16 . Accordingly, a large number of different theories

about the physical mechanism of SRG formation exist,

such as the isomerization pressure model, the asymmetric diffusion

model or the fluid dynamics model 13,27–30 . However,

none of the models can explain all of the observed features.

In this contribution, we investigate the response of a

thin film of the azo-polymer poly-disperse-orange3-methylmethacrylate

(PDO3M) to photo-stimulus both experimentally

and computationally. In the experiment, the polymer film

was illuminated with a pattern of dark and bright stripes of

varying distance generated by a spatial light modulator and


Sample title 2

the surface profile was read out using atom force microscopy.

The computational treatment comprised atomistic molecular

dynamics simulations explicitly taking into account the photoisomerization

dynamics and the light polarization 31 . This

combined approach allows for new insight into the mechanistic

aspects of the surface relief forming process. A particular

focus lies on the role of the light polarization direction relative

to the stripe pattern.

II.

COMPUTATIONAL DETAILS

Our atomistic computational model consists of a 20 nm

thick quasi-infinite periodic slab with a unit cell of size

19.3×32.1×50 nm 3 (z being the plane normal direction) containing

1944 16-meric PDO3M molecules (i.e., a total of

995 328 atoms and 31 104 AB chromophore units) that can

be partitioned into a bright (irradiated) half and a dark (not

irradiated) half, as shown in Fig.1. To maintain the shape of

the slab, the polymer backbone atoms at the bottom (i.e., with

z ≤0.9 nm) were fixed by constraints.

Laser-induced E ↔ Z photoswitching was then simulated

by repeatedly applying our molecular mechanics switch 31–33

(that we extended to include the Z → E direction (cf. SI)) every

50 ps to all chromophore units in the active region whose

transition dipole moment, D, was sufficiently aligned with

the polarisation vector, P, of the incident light. We chose

the two polarisation directions P = (1, 0, 0) (x-direction) and

P = (0, 1, 0) (y-direction).

The overall heating of the slab due to the energy uptake

of ≈ 2.8 eV per photoactivated chromophore was controlled

by a moderate heat bath of Nosé-Hoover type (τ = 80 ps,

T = 300 K) applied separately to the bright and dark regions,

respectively, in order to prevent the extreme temperatures observed

in our previous studies 31 .

In order to further probe orientation and finite size effects,

we carried out simulations with two different partitionings of

dark and bright regions as shown in Fig. 1. In the first partitioning,

the slab was divided into two halves along the y-

direction. With the two polarisations considered, this resulted

in the setups ’yy’ and ’yx’ (Fig. 1). The second partitioning

consisted in dividing the system in two halves in x-direction,

resulting in the ’xy’ and ’xx’ setups.

The molecular dynamics simulations were performed using

an in-house modified version Gromacs ?

III.

RESULTS AND DISCUSSION

In this section we describe the observed surface modifactions.

We first present the results from theory and experiment

in two separate sections and then discuss the implications on

the mechanistic aspects.

FIG. 1. Schematic representation of the different simulation setups.

A. Molecular dynamics simulations

The computer simulations mimick, due to the periodic

boundary conditions, alternating bright and dark stripes on

the thin film surface with the stripes oriented either parallel

to the x (’vertical’) or the y axis (’horizontal’), respectively

(cf. Fig. 1).

This general setup of the periodic MD box is visualised in

Fig. 2 together with the resulting surface profile.

FIG. 2. View of the periodic simulation box. Top row: complete box

with periodic continuation displayed for the first top layers (highlighted

via colour scheme) and ’bright’ area indicated by a yellow

rectangle. Bottom row: resulting final surface profile obtained after

photostimulation and re-cooling (see text).

The time evolution of the emerging height profiles for the

four different setups is shown in Fig. 3, where we have plotted

the maximum z-value of the molecules with a given position

along the axis perpendicular to the bright- dark separation,

or in other words, the profile seen when viewing the sample

parallel to the stripes.

Focussing first on Fig. 3a), it is seen that the photoirradiation

within the first 1 ns leads to a significant increase in height


Sample title 3

especially in the centre of the bright region (blue curve) due

to the energy input by repeated photoexcitation (note that the

centre of the dark region around y = 8 nm is nearly unaffected

throughout). In the time window 1–2 ns, the increase in the

centre comes to a halt and the elevation begins to shift to the

bright/dark boundary (red curve). This latter effect is dominant

in the remaining illumination time window (2–5 ns) and

is accompanied by a drop in height in the central bright region

(orange, black, yellow). After re-cooling to 300 K, a trench of

≈ 1 nm depth is clearly visible in the centre of the bright region

together with two hills (≈1 nm) formed at the bright/dark

boundary (green curve). The effect is similar when changing

the direction of polarisation as seen from Fig.3c), but is

significantly less pronounced upon changing the bright/dark

separation to ’horizontal’ (cf. Fig.3b), d)).

T / K

T / K

900

800

700

600

500

400

300

0 1 2 3 4 5 6

900

800

700

600

500

400

300

0 1 2 3 4 5 6

0 1 2 3 4 5 6 7 8 9 10 11

0 1 2 3 4 5 6 7 8 9 10 11

all

bright

dark

t / ps

t / ns

z / nm

20

18

FIG. 4. Time evoultion of global (bulk, bright and dark region) temperature

for the four different MD setups.

16

14

20

-15 -10 -5 0 5 10 15

-10 -5 0 5 10

0 ns

0.5 ns

1.0 ns

2.0 ns

3.0 ns

4.0 ns

5.0 ns

re-cooled

ter 1 ns. While the temperature bath manages to maintain a

bulk average temperature of about 600 K (see above), a hot

’bubble’ appearing in the bright region is clearly visible, with

peak temperatures of about 1400 K in the ’vertical’ setup and

roughly 200 K lower for the ’horizontal’ setup.

z / nm

18

16

14

-15 -10 -5 0 5 10 15

y / nm

-10 -5 0 5 10

x / nm

FIG. 3. Time evolution of the slab height profiles observed for the

four different MD setups.

The characteristics of the thermal energy uptake due to photoexcitation

is shown in Fig. 4, which presents the time evolution

of the overall bulk system temperature together with the

average values for the bright and dark regions, respectively.

In the first 1 ns, the system is seen to heat up from the initial

value of 300 K to a global average of roughyl 600 K in all

cases. Afterwards, the temperature remains stable for the remaining

duration. The pronounced fluctuations are due to the

photoexcitation cycles every 50 ps. A closer look at this analysis

reveals some subtle differences between the global average

temperatures of the ’vertical’ and ’horizontal’ setups. While

the overall system temperature of the ’vertical’ setups plateaus

at about 630 K, it converges to about 580 K for the ’horizontal’

setup. Furthermore, it is seen that the average temperature in

the bright region is ca. 100 K higher in the ’vertical’ setup as

compared to the ’horizontal’ partitioning (800 K vs. 700 K).

At the same time, the average temperature of the dark region is

lower for the ’vertical’ than for the ’horizontal’ setup. This is

attributed to the different widths of the bright and dark stripes

in the two setups, the narrower horizontal stripes allowing for

a more rapid exchange of heat between the two regions.

Going beyond temperature averages, Fig. 5 presents the

temperature distributions across the slab profile as reached af-

z / nm z / nm

FIG. 5. Temperature distribution of slab profile observed for the four

different MD setups after 1 ns.

The lower overall temperature reached in the two ’horizontal’

setups corresponds to a slightly lower (≈ 10 %) number of

activated chromophore units as can be seen from Fig. 6, where

the number of photoactivated E → Z and Z → E azo chromophore

units is shown, respectively, as a function of time

together with their sum (blue line). Here, it is seen that –

starting from an all-E situation – in each of the four setups a

kind of ’steady state’ is reached with about 58 % E → Z and

42 % Z → E excitation.

Next, we turn to the drift motion and the associated mass

transport (cf. Fig. ??), which is ultimately responsible for the

formation of a surface pattern. We begin our detailed analysis

of the molecular motion involved with Fig. 7, which presents

the average root mean square displacements (total rmsd and

cartesian components x, y and z) of the molecules represented


Sample title 4

no. of active NN units

no. of active NN units

4000

3000

2000

1000

4000

3000

2000

1000

0

0 1 2 3 4 5

0 1 2 3 4 5 6 7 8 9 10

E2Z

Z2E

total

rmsd / nm

rmsd / nm

1.6

1.2

0.8

0.4

0.0

-0.4

-0.8

0 1 2 3 4 5 6

1.6

1.2

0.8

0.4

0.0

-0.4

0 1 2 3 4 5 6 7 8 9 10 11

x

y

z

tot

x

y

z

tot

0

0 1 2 3 4 5

t / ns

0 1 2 3 4 5 6 7 8 9 10

t / ns

-0.8

0 1 2 3 4 5 6

t / ns

0 1 2 3 4 5 6 7 8 9 10 11

t / ns

FIG. 6. Time evoultion of photoexcitation of azo chromophore units

for the four different MD setups.

by their center of mass (COM) in the ’bright’ and ’dark’ region

for the four different setups. Generally, it is seen that

initially the rmsd in the bright region is dominated by the z-

component (blue line) for all four setups, implying an initial

height increase. Later on, the growth in z slows down, while

x and y-components rise more rapidly. the In the case of the

’yx’ and ’yy’ setups, movement starts to be dominated by the

y-component from about 1 ns onwards.

For the horizontal setups ’xx’ and ’xy’ (Fig. 7b,d), the

molecules are seen to travel much more slowly, which is in

line with their reduced temperature (Fig. 4). The general behaviour

of the x, y, and z components is similar to the vertical

setups, except for the fact that the x and y components never

diverge meaning that there is no preferred direction of mass

transport.

All four setups have in common that mass transport occurs

from the bright to the dark region. However, in contrast to

our expectation and seemingly at odds with many previous

works, molecular displacement is practically independent of

the polarisation direction. As it is known that the influence of

polarisation depends on many factors including the chemical

nature of the material and the light intensity, we aimed to verify

our computational observations by performing matching

experiments using the same material (see below).

B. Experimental

In order to experimentally verify the computational results,

we illuminated a PDO3MA azopolymer film with different

intensity patterns. These consisted of bright stripes of varying

distance, while their width remained the same. This way, the

position of the peaks of the SRG, whether in the illuminated

or dark regions, can be determined. Moreover, we varied the

polarization of the light to investigate its impact on the SRG

formation.

For fabrication of the azopolymer samples, 3 wt% of

FIG. 7. Time evolution of average COM root mean square displacement

(RMSD) split into individual cartesian components and

dark/bright regions for the four different MD setups. The curves for

the dark region are plotted with negative RMSD for better distinction.

poly-disperse-orange3-methyl-methacrylate (PDO3M)

(Sigma-Aldrich) were added to tetrahydrofuran (THF) and

stirred for 2 hours. In the meantime, glass substrates were

cleaned via ultrasonication in acetone and isopropyl alcohol

for 10 min each and subsequently treated in UV/ozone to

improve wettability. The azopolymer solution was spincoated

at 1000 rpm for 10 s on the glass substrates and finally

annealed on a hotplate at 110 ◦ C for 1 h.

To generate stripe patterns with different distances of the

bright stripes we utilized the setup shown in Fig. 8. Light

from a laser diode with a wavelength of 405 nm is modified by

a half-wave plate to ensure a polarization parallel to the long

side of the employed LETO phase-only spatial light modulator

(SLM) (HOLOEYE). The SLM is based on a reflective

LCOS microdisplay with a full high definition (1920 × 1080

pixel) resolution and a pixel pitch of 6.4 µm. It is used to tailor

the transverse intensity distribution of the reflected wave.

Therefore a blazed grating consisting of a linear phase ramp

between 0 and 2π with a grating period of ten pixels and a tilt

angle of 51 ◦ is addressed to the SLM 34 . The reflected beam is

focused with a lens on a pinhole to separate the first diffraction

order containing the desired pattern from residual light and

imaged with a second lens on the sample, resulting in a 10×

demagnification of the pattern. Using an infinity-corrected objective

and a lens, the generated light pattern can be observed

with a CMOS camera. After illumination, the azopolymer

sample surfaces were examined with the atomic force microscope

(AFM) NaioAFM from Nanosurf with the Naio Control

Software (version 3.6), operated at ambient conditions. Using

the sharp tip in the AFM, high resolution topography maps

can be obtained, see Fig. 9. On the left, the different illumination

patterns used to inscribe the SRGs in the films. Next to

them, images of the respective AFM measurements of the topographies

and the corresponding line profiles are shown. In

each case, the bright stripes have a width of 12.8 µm. The dis-


Sample title 5

FIG. 8. Scheme of the experimental setup. λ/2: half-wave plate,

SLM: phase-only spatial light modulator, M: mirror, L: lens, P: pinhole,

S: sample, MO: infinity-corrected microscope objective and

Cam: camera.

tance between the stripes varies between 25.6 µm (Fig. 9a),

31.4 µm (Fig. 9b) and 37.2 µm (Fig. 9c). For all of them,

the polarization of the light is perpendicular to the stripes and

the illumination time is 20 h. The intensity varies slightly between

158 mWcm −2 (Fig. 9a), 164 mWcm −2 (Fig. 9b) and

151 mWcm −2 (Fig. 9c).

The line profiles extracted from the topography images can

provide insight into the roughness of the surface, which can be

directly correlated to the polymer motion. Line profiles for the

sample in Fig. 9a show a valley of approximately 12 µm wide

and a plateau which is 26 µm wide. The average depth/height

of these features is between 60 nm and 100 nm. The line profiles

of the other two samples show valleys of the same width,

while the width of the plateaus varies. Thus, comparing the

dimensions of the topographical features in these images to

the respective illumination patterns, we can conclude that the

azopolymer has the tendency to move away from the bright

regions in the illumination pattern and accumulate in the dark

regions.

Moreover, the influence of the polarization on the SRG formation

is investigated. While in Fig. 9 the polarization is perpendicular

to the stripes, in Fig. 10 it is parallel. Otherwise

the illumination parameters are the same as in Fig. 9a with an

illumination time of 20 h and an intensity of 158 mWcm −2 .

Again the width of the valleys correspond to the width of the

bright stripes and the plateaus to the dark regions in between.

Also the depth of the structures is again in the range of 60 nm

to 100 nm as for the illumination with perpendicular polarization.

The only difference is the appearance of small bead-like

structures of 10 nm to 20 nm height in the valleys.

FIG. 9. Illumination patterns with different distances between the

bright stripes and the respective AFM topography images and line

profiles: (a) 25.6 µm, (b) 31.4 µm and (c) 37.2 µm. The line profiles

are extracted across the whole width of the respective scan image at

points indicated by the colored arrows. Laser light polarized perpendicular

to the striped pattern is used.

C. Merging theory and experiment

The simulation predicts that during illumination maxima

form in the exposed regions and then migrate from the sides

into the dark regions, as has been observed similarly earlier

by Yadavalli et al. 35 . Double-peak substructures similar to

those formed in our simulation after cooling have also been

FIG. 10. Illumination pattern with 12.8 µm wide bright stripes with

25.6 µm in between and respective AFM topography images and corresponding

line profiles of the illuminated azopolymer sample. The

line profiles are extracted across the whole width of the scan image at

points indicated by the colored arrows. The polarization of the laser

beam is parallel to the illumination pattern.


Sample title 6

observed in other publications under pulsed illumination or

depending on the material 23,36,37 . In the experiment, the dark

regions are identical with the elevated areas, which, when

compared with the simulation, equates to a full movement of

the material out of the illuminated regions. This is consistent

with the significantly longer illumination times of 20 h in the

experiment, compared to only a few picoseconds in the simulation.

We conclude that very short cw or pulsed illumination

can induce a double peak with maxima at the interfaces of illuminated

and dark regions, while a longer illumination leads

to a maximum in the dark region.

In both the simulation and in the experiment, the polarization

of the incident light has no or only a minor impact on

the surface modulation. In the experiment, small bead-like

structures of 10 nm to 20 nm height emerge in the valleys.

These are not predicted in the simulation. This difference

could again be due to the different time and length scales of

simulation and experiment. The illumination times of only picoseconds

and stripe sizes of a few nanometres investigated

in the simulation cannot be realized in these experiments and

vice versa.

The reason for this rare (virtual) independence of SRG

formation on polarization could be the high fluence used in

these experiments. While in most SRG experiments, the

azopolymer films are illuminated for a few seconds to several

minutes with intensities in the range of tens to hundreds

of mWcm −2 , here the films are illuminated for 20 hour

with over 150 mWcm −2 and thus with a significantly higher

fluence 4,7,16 . Also in our simulations, the fluence is very

high with ... photons per .. Our experiment and simulation

are thus consistent with publications claiming polarizationindependent

SRG formation under high-intensity cw and

pulsed irradiation 7,20,21,23–26 .

With regard to the cause of photomigration, the simulations

clearly show that the conversion of light-energy into local heat

is a prerequisite for this phenomenon to occur. In the fluence

regime studied here, mass transport appears to take place in

the direction of the negative temperature gradient and can thus

be described my the theory of thermal diffusion 38 .

IV.

CONCLUSIONS

We have carried out a combined experimental and theoretical

study of the surface relief grating formation in the azopolymer

poly-disperse-orange3-methyl-methacrylate. On the experimental

side, azopolymer polymer films were illuminated

with light patterns of bright and dark stripes created by a spatial

light modulator. The surface profile was subsequently

characterized using atom force microscopy. Thus is could be

established that mass transport occurred away from the bright

stripes into the dark stripes. Furthermore, the polarisation direction

was seen to have a minor effect. These experimental

findings were corroborated by atomistic molecular dynamics

simulations that explicitly modelled the photoisomerisation

dynamics and light polarisation. The simulations further

showed that local heating due to light-absorption is necessary

for mass transport to occur. In the high-fluence regime investigated

in this work, mass transport appears to be dominated by

the temperature gradient rather than any more subtle effects

related to the photoisomerisation dynamics.

All AIP journals require that the initial citation of figures or

tables be in numerical order. LATEX’s automatic numbering of

floats is your friend here: just put each figure environment

immediately following its first reference (\ref), as we have

done in this example file.

ACKNOWLEDGMENTS

We wish to acknowledge the support of the author community

in using REVTEX, offering suggestions and encouragement,

testing new versions, . . . .

DATA AVAILABILITY STATEMENT

The data that support the findings of this study are available

from the corresponding author upon reasonable request.

Appendix A: A little more on appendixes

1. A subsection in an appendix

1 H. Zhang, “Reprocessable photodeformable azobenzene polymers,”

Molecules 26, 4455–4484 (2021).

2 P. Weis, W. Tian, and S. Wu, “Reprocessable photodeformable azobenzene

polymers,” Chem. Eur. J. 24, 6494–6505 (2021).

3 S. Lee, H. S. Kang, and J.-K. Park, “Directional photofluidization lithography:

Micro/nanostructural evolution by photofluidic motions of azobenzene

materials,” Adv. Mater. 24, 2069–2103 (2012).

4 C. J. Barrett, J. Mamiya, K. G. Yager, and T. Ikeda, “Photo-mechanical

effects in azobenzene-containing soft materials,” Soft Matter 3, 1249–1261

(2007).

5 K. G. Yager and C. J. Barrett, “Photomechanical surface patterning in azopolymer

materials,” Macromolecules 39, 9320–9326 (2006).

6 A. Natansohn and P. L. Rochon, “Photoinduced motions in azo-containing

polymers,” Chem. Rev. 102, 4139–4175 (2002).

7 N. K. Viswanathan, D. Y. Kim, S. Bian, J. Williams, W. Liu, L. Li,

L. Samuelson, J. Kumar, and S. K. Tripathy, “Surface relief structures on

azo polymer films,” J. Mater. Chem. 9, 1941–1955 (1999).

8 W.-C. Xu, S. Sun, and S. Wu, “Photoinduced reversible solid-to-liquid transitions

for photoswitchable materials,” Angew. Chem. Int. Ed. 58, 9712–

9740 (2019).

9 J. Konieczkowska, K. Bujak, and E. Schab-Balcerzak, “A short review

of the photomechanical effect in azo-containing amorphous (glassy) polymers,”

eXPRESS Polymer Letters 15, 459–472 (2021).

10 S. De Martino, F. Mauro, and P. Netti, “Photonic applications of azobenzene

molecules embedded in amorphous polymer,” La Rivista del Nuovo

Cimento 43, 599–629 (2020).

11 M. Hendrikx, A. P. Schenning, M. G. Debije, and D. J. Broer, “Lighttriggered

formation of surface topographies in azo polymers,” Crystals 7,

231 (2017).

12 T. Fukuda, H. Matsuda, T. Shiraga, T. Kimura, M. Kato, N. K. Viswanathan,

J. Kumar, and S. K. Tripathy, “Photofabrication of surface relief grating on

films of azobenzene polymer with different dye functionalization,” Macromolecules

33, 4220–4225 (2000).

13 C. J. Barrett, A. L. Natansohn, and P. L. Rochon, “Mechanism of optically

inscribed high-efficiency diffraction gratings in azo polymer films,” The

Journal of Physical Chemistry 100, 8836–8842 (1996).


Sample title 7

14 A. Ambrosio, L. Marrucci, F. Borbone, A. Roviello, and P. Maddalena,

“Light-induced spiral mass transport in azo-polymer films under vortexbeam

illumination,” Nature communications 3, 1–9 (2012).

15 T. Grosjean and D. Courjon, “Photopolymers as vectorial sensors of the

electric field,” Optics Express 14, 2203–2210 (2006).

16 V. M. Kryshenik, Y. M. Azhniuk, and V. S. Kovtunenko, “All-optical patterning

in azobenzene polymers and amorphous chalcogenides,” Journal of

Non-Crystalline Solids 512, 112–131 (2019).

17 N. C. R. Holme, L. Nikolova, S. Hvilsted, P. H. Rasmussen, R. H. Berg, and

P. S. Ramanujam, “Optically induced surface relief phenomena in azobenzene

polymers,” Appl. Phys. Lett. 74, 519–521 (1999).

18 N. S. Yadavalli, T. König, and S. Santer, “Selective mass transport of

azobenzene-containing photosensitive films towards or away from the light

intensity,” Journal of the Society for Information Display 23, 154–162

(2015).

19 P. Karageorgiev, D. Neher, B. Schulz, B. Stiller, U. Pietsch, M. Giersig, and

L. Brehmer, “From anisotropic photo-fluidity towards nanomanipulation in

the optical near-field,” Nature Materials 4, 699–703 (2005).

20 K. G. Yager and C. J. Barrett, “All-optical patterning of azo polymer films,”

Current opinion in solid state and materials science 5, 487–494 (2001).

21 A. Leopold, J. Wolff, O. Baldus, M. Huber, T. Bieringer, and S. Zilker,

“Thermally induced surface relief gratings in azobenzene polymers,” The

Journal of Chemical Physics 113, 833–837 (2000).

22 A. Priimagi and A. Shevchenko, “Azopolymer-based micro-and nanopatterning

for photonic applications,” Journal of polymer science Part B: Polymer

physics 52, 163–182 (2014).

23 O. Baldus, A. Leopold, R. Hagen, T. Bieringer, and S. c. Zilker, “Surface

relief gratings generated by pulsed holography: A simple way to polymer

nanostructures without isomerizing side-chains,” The Journal of Chemical

Physics 114, 1344–1349 (2001).

24 O. N. Oliveira Jr, D. S. dos Santos Jr, D. T. Balogh, V. Zucolotto, and C. R.

Mendonça, “Optical storage and surface-relief gratings in azobenzenecontaining

nanostructured films,” Advances in colloid and interface science

116, 179–192 (2005).

25 T. Ubukata, Y. Moriya, and Y. Yokoyama, “Facile one-step photopatterning

of polystyrene films,” Polymer journal 44, 966–972 (2012).

26 K. G. Yager and C. J. Barrett, “Temperature modeling of laser-irradiated

azo-polymer thin films,” The Journal of chemical physics 120, 1089–1096

(2004).

27 C. J. Barrett, P. L. Rochon, and A. L. Natansohn, “Model of laser-driven

mass transport in thin films of dye-functionalized polymers,” The Journal

of chemical physics 109, 1505–1516 (1998).

28 P. Lefin, C. Fiorini, and J.-M. Nunzi, “Anisotropy of the photo-induced

translation diffusion of azobenzene dyes in polymer matrices,” Pure and

Applied Optics: Journal of the European Optical Society Part A 7, 71

(1998).

29 K. Sumaru, T. Yamanaka, T. Fukuda, and H. Matsuda, “Photoinduced surface

relief gratings on azopolymer films: Analysis by a fluid mechanics

model,” Applied physics letters 75, 1878–1880 (1999).

30 D. Bublitz, B. Fleck, and L. Wenke, “A model for surface-relief formation

in azobenzene polymers,” Applied Physics B 72, 931–936 (2001).

31 M. B"ockmann and N. L. Doltsinis, “Towards understanding photomigration:

Insights from atomistic simulations of azopolymer films explicitly

including light-induced isomerization dynamics,” J. Chem. Phys. 145,

154701 (2016).

32 M. Böckmann, S. Braun, N. L. Doltsinis, and D. Marx, “Mimicking

photoisomerisation of azo-materials by a force field switch derived from

nonadiabatic ab initio simulations: Application to photoswitchable helical

foldamers in solution,” J. Chem. Phys. 139, 084108 (2013).

33 M. Böckmann, D. Marx, C. Peter, L. D. Site, K. Kremer, and N. L. Doltsinis,

“Multiscale modelling of mesoscopic phenomena triggered by quantum

events: light-driven azo-materials and beyond,” Phys. Chem. Chem. Phys.

13, 7604–7621 (2011).

34 J. A. Davis, D. M. Cottrell, J. Campos, M. J. Yzuel, and I. Moreno, “Encoding

amplitude information onto phase-only filters,” Applied Optics 38,

5004–5013 (1999).

35 N. S. Yadavalli, M. Saphiannikova, and S. Santer, “Photosensitive response

of azobenzene containing films towards pure intensity or polarization interference

patterns,” Appl. Phys. Lett. 105, 051601 (2014).

36 V. Damian, E. Resmerita, I. Stoica, C. Ibanescu, L. Sacarescu, L. Rocha,

and N. Hurduc, “Surface relief gratings induced by pulsed laser irradiation

in low glass-transition temperature azopolysiloxanes,” Journal of Applied

Polymer Science 131 (2014).

37 E. Schab-Balcerzak, M. Siwy, M. Kawalec, A. Sobolewska, A. Chamera,

and A. Miniewicz, “Synthesis, characterization, and study of photoinduced

optical anisotropy in polyimides containing side azobenzene units,” The

Journal of Physical Chemistry A 113, 8765–8780 (2009).

38 M. Gillan, “Diffusion in a temperature gradient,” in Mass transport in

solids, edited by F. Béniére and C. Catlow (Springer, New York, 1983).

39 J. Vapaavouri, A. Laventure, C. Bazuin, O. Lebel, and C. Pellerin,

“Submolecular plasticization induced by photons in azobezene materials,”

J. Am. Chem. Soc. 137, 13510–13517 (2015).

40 N. S. Yadavalli, T. König, and S. Santer, “Selective mass transport of

azobenzene-containing photosensitive films towards or away from the light

intensity,” J. Soc. Inf. Disp. xx, xxx–xxx (2015).

41 N. S. Yadavalli, D. Korolkov, J. Moulin, M. Krutyeva, and S. Santer, “Probing

opto-mechanical stresses within azobenzene-containing photosensitive

polymer films by a thin metal film placed above,” ACS Appl. Mat. & Interf.

6, 11333–11340 (2014).

42 J.-B. Accary and V. Teboul, “How does the isomerization rate affect

the photoisomerization-induced transport properties of a doped molecular

glass-former?” J. Chem. Phys. 139, 034501 (2013).

43 V. Teboul, M. Saiddine, and J.-M. Nunzi, “Isomerization-induced dynamic

heterogeneity in a glass former below and above t g ,” Phys. Rev. Lett. 103,

265701 (2009).

44 V. Toshchevikov, M. Saphiannikova, and G. Heinrich, “Microscopic theory

of light-induced deformation in amorphous side-chain azobenzene polymers,”

J. Phys. Chem. B 113, 5032–5045 (2009).

45 E. Merino and M. Ribagorda, “Control over molecular motion using the

cis–trans photoisomerization of the azo group,” Beilstein journal of organic

chemistry 8, 1071–1090 (2012).

46 S. Golghasemi Sorkhabi, S. Ahmadi-Kandjani, F. Cousseau,

M. Loumaigne, S. Zielinska, E. Ortyl, and R. Barille, “Surface quasi

periodic and random structures based on nanomotor lithography for light

trapping,” Journal of Applied Physics 122, 015303 (2017).

47 C. D. Abernethy, G. M. Codd, M. D. Spicer, and M. K. Taylor, “A highly

stable N-heterocyclic carbene complex of trichloro-oxo-vanadium(V) displaying

novel Cl—C(carbene) bonding interactions,” J. Am. Chem. Soc.

125, 1128–1129 (2003).

48 (2007).

49 A. J. Arduengo, III, H. V. R. Dias, R. L. Harlow, and M. Kline, “Electronic

stabilization of nucleophilic carbenes,” J. Am. Chem. Soc. 114, 5530–5534

(1992).

50 A. J. Arduengo, III, S. F. Gamper, J. C. Calabrese, and F. Davidson, “Lowcoordinate

carbene complexes of nickel(0) and platinum(0),” check: jacsat

116, 4391–4394 (1994).

51 L. N. Appelhans, D. Zuccaccia, A. Kovacevic, A. R. Chianese, J. R.

Miecznikowski, A. Macchioni, E. Clot, O. Eisenstein, and R. H. Crabtree,

“An anion-dependent switch in selectivity results from a change of

C—H activation mechanism in the reaction of an imidazolium salt with

IrH5(PPh3)2,” J. Am. Chem. Soc. 127, 16299–16311 (2005).

52 A. M. Coghill and L. R. Garson, eds., The ACS Style Guide, 3rd ed. (Oxford

University Press, Inc. and The American Chemical Society, New York,

2006).

53 F. A. Cotton, G. Wilkinson, C. A. Murillio, and M. Bochmann, Advanced

Inorganic Chemistry, 6th ed. (Wiley, Chichester, United Kingdom, 1999).

54 “Communication from the european commission to the european council

and the european parliament: 20 20 by 2020: Europe’s climate change opportunity,”

Tech. Rep. (Brussels, Belgium, 2008).

55 E. Friedman-Hill, “Writing rules in jess,” in Jess in Action: Java Rulebased

Systems (Manning Publications Co., Greenwich, CT, USA, 2003) 1st

ed.

56 A. L. Johnson, “1-(alkylsubstituted phenyl)imidazoles useful in acth reverse

assay,” (1972).

57 N. S. Yadavalli, M. Saphiannikova, and S. Santer, “Photosensitive response

of azobenzene containing films towards pure intensity or polarization interference

patterns,” Applied Physics Letters 105, 051601 (2014).

58 M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R.

Cheeseman, Montgomery, Jr., J. A., T. Vreven, K. N. Kudin, J. C. Burant,


Sample title 8

J. M. Millam, S. S. Iyengar, J. Tomasi, V. Barone, B. Mennucci, M. Cossi,

G. Scalmani, N. Rega, G. A. Petersson, H. Nakatsuji, M. Hada, M. Ehara,

K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda,

O. Kitao, H. Nakai, M. Klene, X. Li, J. E. Knox, H. P. Hratchian, J. B.

Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann,

O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, P. Y. Ayala,

K. Morokuma, G. A. Voth, P. Salvador, J. J. Dannenberg, V. G. Zakrzewski,

S. Dapprich, A. D. Daniels, M. C. Strain, O. Farkas, D. K. Malick, A. D.

Rabuck, K. Raghavachari, J. B. Foresman, J. V. Ortiz, Q. Cui, A. G. Baboul,

S. Clifford, J. Cioslowski, B. B. Stefanov, G. Liu, A. Liashenko, P. Piskorz,

I. Komaromi, R. L. Martin, D. J. Fox, T. Keith, M. A. Al-Laham, C. Y. Peng,

A. Nanayakkara, M. Challacombe, P. M. W. Gill, B. Johnson, W. Chen,

M. W. Wong, C. Gonzalez, and J. A. Pople, Gaussian 03, Gaussian, Inc.,

Wallingford, CT (2004).

59 A. Abarca, P. Gómez-Sal, A. Martín, M. Mena, J. M. Poblet, and

C. Yélamos, “Ammonolysis of mono(pentamethylcyclopentadienyl) titanium(IV)

derivatives,” Inorg. Chem. 39, 642–651 (2000).

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!