Synthesis and Study of The Optical Properties of PMMA

Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

polymers

Article
Synthesis and Study of the Optical Properties of PMMA
Microspheres and Opals
Mayra Matamoros-Ambrocio 1 , Enrique Sánchez-Mora 2, * , Estela Gómez-Barojas 1
and José Alberto Luna-López 1

1 Centro de Investigación en Dispositivos Semiconductores, Instituto de Ciencias,


Benemérita Universidad Autónoma de Puebla, P.O. Box 196, 72000 Puebla, Mexico;
[email protected] (M.M.-A.); [email protected] (E.G.-B.);
[email protected] (J.A.L.-L.)
2 Instituto de Física, Benemérita Universidad Autónoma de Puebla, Eco campus Valsequillo,
Independencia O 2 sur No. 50, San Pedro Zacachimalpa, P.O. Box J-48, 72960 Puebla, Mexico
* Correspondence: [email protected]; Tel.: +52-222-229-5610 (ext. 1383)

Abstract: Polymethylmethacrylate (PMMA) microspheres were synthesized by surfactant-free emul-


sion polymerization. These microspheres were used to obtain opals by the self-assembly method.
Monomer and initiator quantities were varied systematically to monitor the size of PMMA micro-
spheres. From SEM and DLS measurements, a trend was observed showing as the monomer and
initiator amounts increased the average diameter of PMMA microspheres increased except when a
minimum monomer amount was reached, for which the size of the microspheres remained practically
constant. Diffuse reflectance spectra were processed by the Kubelka–Munk treatment to estimate the
energy band gap (Eg ) of the PMMA microspheres. It was found that PMMA microspheres present

 an indirect transition. From SEM micrographs, it is seen that PMMA opals photonic crystals are
formed by microspheres in a uniform periodic face-centered cubic (fcc) array. Variable-angle specular
Citation: Matamoros-Ambrocio, M.;
Sánchez-Mora, E.; Gómez-Barojas, E.;
reflectance spectra show that the opals possess a pseudo photonic band gap (PBG) in the visible
Luna-López, J.A. Synthesis and Study and near-IR regions. Furthermore, it was found that PBGs shift towards larger wavelengths as the
of the Optical Properties of PMMA average diameter of the PMMA microspheres increases.
Microspheres and Opals. Polymers
2021, 13, 2171. https://doi.org/ Keywords: PMMA microspheres; opals photonics crystals; optical properties
10.3390/polym13132171

Academic Editor: Jem-Kun Chen


1. Introduction
Received: 2 June 2021
Photonic crystals are ordered structures whose dielectric constant is modulated with
Accepted: 7 June 2021
periods on the wavelength scale of visible light [1]. In nature, they are present in some
Published: 30 June 2021
butterfly wings, bird feathers, and opal gemstones. In these structures, periodicity in
the refraction index gives rise to diffraction effects of photons on dielectric lattice planes,
Publisher’s Note: MDPI stays neutral
resulting in a photonic band gap (PBG) or a stop band that inhibits a range of wavelengths
with regard to jurisdictional claims in
from propagating through the crystal [2].
published maps and institutional affil-
iations.
The photonic crystals with complete photonic band gaps have been studied with
great interest because of their applications in different areas of science and technology.
For example, optical sensors have been coupled to a transduction device that identifies
the optical diffraction that occurs once the external medium changes [3]. Optical waveg-
uides [2], switches [4], light-emitting diodes [5], and templates to fabricate devices with
Copyright: © 2021 by the authors.
interconnected porous matrices have also all been examined [6,7]. In order for the gamma
Licensee MDPI, Basel, Switzerland.
of applications to be successful, improved methods for the fabrication of highly ordered
This article is an open access article
3D photonic crystals must be developed.
distributed under the terms and
conditions of the Creative Commons
Opals are distinguished as a class of photonic crystals mainly obtained by a self-
Attribution (CC BY) license (https://
assembly approach, where particles are organized spontaneously to form ordered three-
creativecommons.org/licenses/by/ dimensional structures [8]. The most common methods used are: gravity sedimentation,
4.0/). centrifugation, vertical deposition, and spin coating [9]. The materials most widely used

Polymers 2021, 13, 2171. https://doi.org/10.3390/polym13132171 https://www.mdpi.com/journal/polymers


Polymers 2021, 13, 2171 2 of 16

for opals are silica [10] and polymers as PS [11] and PMMA [12]. On the other hand,
block copolymers (BCPs) provide another route for creating photonic crystal structures,
where self-assembly of high-molecular-weight block copolymers yields periodic photonic
crystal. For example, polystyrene–polyisoprene diblock copolymer demonstrates that block
copolymers can form self-assembly optical structures for the construction of lightweight
and flexible photonic devices [13]. Additionally, the system polystyrene-b-quaternized poly
(2-vinyl pyridine) (PS-b-QP2VP) lamellar films provide photonic gels with exceptionally
large stop-band tunability across the ultraviolet-visible and near-infrared regions [14]. This
expected system will lead to many novel applications including colorimetric sensors, active
components of simple display devices, electrically controlled tunable optically pumped
lasers, photonic switches, and multiband filters. Regarding the recent use of new polymeric
materials, inspired by structural coloration arising from assembled melanosomes in avian
feathers, they have been prepared and assembled using synthetic melanin nanoparticles
that offer numerous new opportunities towards multifunctional photonics [15].
The opals obtained using polymer monodisperse particles have proved to be an
efficient approach because the preparative process is simple, easy, and cheap [11]. Among
the polymers, the most studied is PMMA because it has good thermal stability, interesting
optical properties, and a high Young’s modulus of around 3 MPa, making it thermoplastic
with high resistance to scratching [16].
The PMMA synthesis method most used is the surfactant-free emulsion polymer-
ization (SFEP) because the obtained colloidal spheres have good monodispersity and
clean surface features [12]. In this process, the monomer is dispersed in water and, in
the presence of a free radical initiator, the polymerization is started. The oligomers act
as surfactants and form micelles as soon as their concentration exceeds critical micellar
concentration. The polymerization reaction proceeds within micelles analogously to the
classical emulsion polymerization. The stability of the colloidal dispersion is reached due
to stabilizing groups originating from the initiator molecules that are covalently bonded to
the surface of spheres [17].
Using the SFEP method, synthesis conditions including temperature, stirring speed,
and amounts of monomer and initiator have been proven to be key factors influencing
particle size. In general, if we want to obtain small-size particles, we should increase the
temperature, use a small amount of initiator, or decrease the amount of monomer. In
the other case, to obtain a large particle size, the amount of monomer must be increased,
the amount of initiator must be decreased, or the reaction must be carried out at low
temperature [1,6,18–21]. Although these results are well established in the literature, a
careful review of the synthesis of PMMA and other polymers using the SFEP method
shows inconsistencies when the amount of initiator is varied [6]. Thus, in this research
work we investigate the effect of the amount of monomer and the amount of initiator used
to obtain PMMA microspheres by the SFEP method.
On the other hand, the optical properties of PMMA material have been widely studied
and reported [22–24]. Although a relationship among optical properties, morphology, and
size of PMMA microspheres has not been elucidated, it has been widely accepted in the
literature to report the calculation of both direct and indirect energy band gaps from UV-Vis
spectra. By the same token, many of the reported studies do not justify with certainty the
type of transition that takes place in this material [25].
The success of making opals with good reproducibility is highly dependent on the
ability to preparate monodisperse colloidal spheres with controlled size on the optimization
of a synthesis reaction and experimental conditions so that the spheres self-assemble into
ordered matrices with a defined structure and good reproducibility.
The aim of this work is to study the effect on the variation of the monomer and the
initiator quantity on the size of the PMMA microspheres that constitute the opals. Fur-
thermore, the PMMA opals’ photonic crystals are characterized to determine the influence
of size on their optical properties. The results can provide insight into the properties
of PMMA.
Polymers 2021, 13, 2171 3 of 16

2. Materials and Methods


2.1. Materials
Methylmetacrylate (MMA; C5 H8 O2 ; 99%) and 2,20 -azobis (2-methylpropionamidine)
dihydrochloride (V50, C8 H20 Cl2 N6 ; 97%) were purchased from Sigma-Aldrich (St. Louis,
MO, USA) and used without further purification. Isopropyl alcohol (C3 H8 O; 99%) was
purchased from J. T. Baker (Phillipsburg, NJ, USA). Deionized water (DI; 18.2 MΩ cm;
Thermo Fisher Scientific, Marietta, OH, USA) was used in all experiments.

2.2. Synthesis of PMMA Microspheres


PMMA microspheres were synthesized by means of surfactant-free emulsion polymer-
ization method reported by E. Sánchez et al. [26]. Briefly, in a three-necked round bottom
flask (250 mL) equipped with a reflux system, 160 mL of DI water and methylmethacrylate
were added, and the system was kept under vigorous magnetic stirring (~400 rpm) and
constant N2 bubbling (0.2 cm3 s−1 ) flux. At this moment, the temperature in the system was
raised to 75 ◦ C, then the V50 was added. Since the polymerization is an exothermic reaction,
the temperature of the system reached 95 ◦ C; this temperature was kept constant for 2 h.
Finally, the reaction was allowed to cool to room temperature and the rising procedure
of the PMMA microsphere was carried out by centrifugation at 6000 rpm (HERMLE Z36
HK) for 30 min. Then, the supernatant was taken away by decantation and the solid was
redispersed in a solution of DI water and isopropyl alcohol (50/50 v/v) in ultrasonic for
15 min. This procedure was repeated 3 times. In total, a set of 9 samples was prepared with
different monomer and initiator amounts, as summarized in Table 1.

Table 1. Synthesis conditions of PMMA microspheres, average diameter measured by SEM (Davg ),
hydrodynamic diameter measured by DLS system (Dh ), PDI and Zeta Potential values (ζ ).

Monomer Initiator
Sample Davg (nm) Dh (nm) PDI ζ (mV)
(mmol) (mmol)
PMMA-1 190 1.65 235 ± 9.0 292 ± 17.1 0.034 47.3 ± 5.3
PMMA-2 190 1.10 244 ± 11.2 301 ± 17.3 0.023 43.9 ± 5.2
PMMA-3 190 0.55 247 ± 15.1 303 ± 17.4 0.069 47.2 ± 5.5
PMMA-4 285 1.65 298 ± 8.0 390 ± 19.7 0.045 51.1 ± 5.4
PMMA-5 285 1.10 273 ± 7.0 380 ± 19.5 0.034 48.5 ± 5.7
PMMA-6 285 0.55 249 ± 7.0 360 ± 19.0 0.042 49.7 ± 5.4
PMMA-7 380 1.65 323 ± 9.0 413 ± 20.3 0.083 46.1 ± 4.9
PMMA-8 380 1.10 311 ± 8.1 398 ± 20.0 0.082 41.4 ± 4.4
PMMA-9 380 0.55 306 ± 11.0 379 ± 19.5 0.088 46.1 ± 4.8

2.3. Preparation of PMMA Opals


PMMA opals were obtained by adding 50 mL of the colloidal suspension in Petri
dishes and left alone at room temperature (25 ◦ C) for 3 weeks to evaporate the DI water
and isopropyl alcohol in which they were dispersed. The thickness of opals was about
1 mm.

2.4. Materials Characterization


The PMMA microspheres were characterized by Fourier Transform Infrared Spec-
troscopy (FT-IR; Spectrum One, Perkin Elmer, Waltham, MA, USA) in transmittance mode
with an ATR accessory and a scanning velocity of 8 nm/s, at room temperature in the range
of 650–4000 cm−1 . Micro-Raman spectroscopy (µ-RS; LabRAM HR-Olympus, Horiba Jobin
Yvon Inc. Edison, NJ, USA) with a He-Ne laser (λ = 632.8 nm) was used as an excitation
source, and an optical microscope at 10× was used to select the region of the interest. The
spectra were recorded from 200–2000 cm−1 range.
The surface morphology of both PMMA microspheres and opals was analyzed with a
scanning electron microscope system (SEM; JSM-7800F, JEOL, Tokyo, Japan). The colloidal
suspensions diluted in DI water were deposited on clean Si wafers. Once the solvent was
Polymers 2021, 13, 2171 4 of 16

evaporated from the wafers, the PMMA microspheres were observed under the SEM, and
their average diameter was determined from the SEM micrographs considering about 300
spheres and using the ImageJ software.
The optical characterization of the PMMA microspheres was performed using a
spectrometer UV-Vis-NIR (200–2500 nm) (Cary 5000 UV-Vis-NIR system, Aligent, Santa
Clara, CA, USA) equipped with a DRA-CA-30I and a Labsphere as a reference to record
diffuse reflectance spectra (DRS). The powder samples were compacted, and each sample
was set on the sample holder. Additionally, it was equipped with a variable angle specular
reflectance accessory (VASRA) and an Al mirror as a reference to measure the specular
reflectance at different angles in the range of 20–70◦ of the PMMA opals. The sample was
approximately 1 cm2 in area and it was set on the sample holder.
The hydrodynamic diameters (Dh ) and the polydispersity indices (PDI) were mea-
sured from the PMMA dilute aqueous solutions with a dynamic light scattering system
(DLS; Malvern Zetasizer Nano ZS, Malvern Instruments Worcestershire, UK) equipped
with an He-Ne laser (λ = 633 nm) with 4.0 mW power and backscattering mode. The PDI
data were obtained directly from the equipment software. Three consecutive measurements
at 25 ◦ C were taken, and each of them consisted of 10 individual runs. The reported values
are average values. The zeta potential measurements were carried out with the Zeta-Sizer
IV system (Malvern Instruments Worcestershire, UK) under the same conditions as the
ones used with the DLS system.

3. Results and Discussion


3.1. PMMA Microspheres
3.1.1. FT-IR and Micro-Raman Analysis
The success of the PMMA microspheres’ synthesis was confirmed by means of the
FT-IR and Raman spectra shown in Figure 1. The PMMA structure contains an ester group,
one methylene (CH2 ) group, and two methyl (CH3 ) groups. Therefore, the bands present
in the spectra correspond to vibrational modes of these groups.

PMMA-1

PMMA-2
Transmittance (a.u.)

PMMA-3

PMMA-4

PMMA-5

PMMA-6

PMMA-7

PMMA-8

PMMA-9
* *
δ (CH2) ν(C-C)
νa(CH2) δ(C-H)
ρ(CH3)

νS(C-O-C)
ν(C-C)

of CH3
νs(C-H) of CH3 ν(C-O)

νa(C-H) of CH3 ν (C=O)


νa(C-O-C)
ρ(CH2)

3200 3000 1800 1600 1400 1200 1000 800


Wavenumber (cm-1)
(a)
νS (C-O-C)
Figure 1. Cont. δ(C-H) δ(C-C)
νa(C-O-C)

ρ(CH3) of CC4
Raman Intensity (a.u.)

of CH3
β(C-C)
ρ(CH2)

δ(CH2) τ(CH2) ν(C-O) *


ν(C=O) ∗
∗ ν(C-C) * PMMA-1
∗ ∗ *
* PMMA-2

PMMA-3

PMMA-4

PMMA-5

PMMA-6

PMMA-7

PMMA-8

PMMA-9

1800 1600 1400 1200 1000 800 600 400 200


Raman shift (cm-1)
Trans
PMMA-9
* *
δ (CH2) ν(C-C)
νa(CH2) δ(C-H)

ρ(CH3)

νS(C-O-C)
ν(C-C)
of CH3
νs(C-H) of CH3 ν(C-O)

νa(C-H) of CH3 ν (C=O)


νa(C-O-C)
ρ(CH2)
Polymers 2021, 13, 2171 3200 3000 1800 1600 1400 1200 1000 800 5 of 16
Wavenumber (cm -1)

(a)
νS (C-O-C)
δ(C-C)

νa(C-O-C)
δ(C-H) ρ(CH3) of CC4

Raman Intensity (a.u.)


of CH3
β(C-C)

ρ(CH2)
δ(CH2) τ(CH2) ν(C-O) *
ν(C=O) ∗
∗ ν(C-C) * PMMA-1
∗ ∗ *
* PMMA-2

PMMA-3

PMMA-4

PMMA-5

PMMA-6

PMMA-7

PMMA-8

PMMA-9

1800 1600 1400 1200 1000 800 600 400 200


Raman shift (cm-1)
(b)

Figure 1. (a) FT-IR spectra and (b) micro-Raman spectra of the whole set of synthesized PMMA
microspheres.

The bands at 2999 and 2952 cm−1 in the FT-IR spectra are assigned to symmetrical
and asymmetrical stretching vibration modes of the CH3 group. The bands at 1448, 1436,
and 1387 cm−1 present in the FT-IR spectrum, and the ones at 1452 and 1390 cm−1 in the
Raman spectra, are assigned to the deformation modes of the C-H bond of the CH3 group.
In the case of the bands at 1123 and the one at 914 cm−1 in the Raman spectrum, and also
the one at 912 cm−1 in the FT-IR spectrum, these are assigned to the rocking modes of the
CH3 groups.
The CH2 group also presents a symmetric stretching vibrational mode at 2952 cm−1
and deformation modes at 1484 cm−1 in the FT-IR spectrum, and at 1481 cm−1 in the
Raman spectrum. The band at 1326 cm−1 appearing in the Raman spectrum is assigned to
the twisting mode of the C-H bond. Finally, the rocking vibrational mode corresponds to
the bands at 877 and 875 cm−1 present in the Raman spectrum, and to the one at 842 cm−1
in the FT-IR spectrum [27]. The band at 1727 cm−1 in the FT-IR spectrum and the one at
1728 cm−1 in the Raman spectrum correspond to the C=O stretching mode. For the C-O
bond, the stretching modes are located in the 1000 to 1400 cm−1 range in both FT-IR and
micro-Raman spectra, even though the bands appearing in the 900 to 1000 cm−1 range
in both spectra bend vibrational modes of the C-C bond are observed at 484, 558, and
601 cm−1 of the Raman spectrum.
The bands in the range 400 to 200 cm−1 of the Raman spectrum correspond to the
deformation modes of the C-C bonds. The remaining bands are assigned to symmetrical
and asymmetrical stretching vibration modes of the C-O-C bond located at 813, 1159, and
1185 cm−1 of the Raman spectrum, while they are observed at 810, 1242 and 1271 cm−1 in
1 the FT-IR spectrum [28].

3.1.2. SEM and DLS Analysis


PMMA microspheres were prepared, keeping the amount of monomer fixed at
285 mmol while varying the amounts of initiator at 0.55, 1.10, and 1.65 mmol. Figure 2
shows SEM micrographs at ×5000 of these prepared PMMA spheres. A corresponding size
distribution histogram appears as an inset in each micrograph. Thus, 0.55 mmol of initiator
gives PMMA spheres had average diameters (Davg ) of 249 nm (σ = 7 nm) for 1.10 mmol,
Davg = 273 nm (σ = 7 nm) and, for 1.65 mmol of initiator, Davg = 298 nm (σ = 7 nm). Thus,
based on these SEM micrographs, we infer that the influence of the initiator is found on
the size of the PMMA microspheres; moreover, the bigger the amount of initiator, the
greater the diameter of the spheres. In SEM micrographs at ×10,000 in (b), (d), and (f), it
is observed with more detail that all PMMA particles have a spherical shape. This result
confirms the spherical term as it has been used in the text.
in the SEM micrographs (Figure 2). This difference is because the DLS measurements were
carried out with PMMA microspheres in a water solution. Then, the formation of a water
molecular layer on the PMMA microspheres’ surface must be considered. The formation
of this layer is due to hydrogen bridge bonds or the PMMA microspheres’ surface proto-
Polymers 2021, 13, 2171 nation, which gives rise to a positive surface charge. The corresponding polydisperse in- 6 of 16
dices (PDI) are also listed in Table 1. In all cases, the PDI values are less than 0.1, indicating
a uniform distribution of the microspheres [29].

70 70
Number of particles

Number of particles

Number of particles
Davg=298 nm Davg=273 nm 60 Davg=249 nm
60 60
𝝈 =8 nm 𝝈 =7 nm 50 𝝈 =7 nm
50 50
40
40 40
30 30
30
20 20 20
10 10 10
0 0 0
270 280 290 300 310 320 330 250 260 270 280 290 300 310 220 230 240 250 260 270 280 290
Size (nm) Size (nm) Size (nm)

(a) (c) (e)

(b) (d) (f)


Figure 2.
Figure 2. SEM
SEMmicrographs
micrographsofofPMMA
PMMAmicrospheres
microspheresshowing
showingthe
the formation
formation with
with 285285 mmol
mmol of monomer
of monomer and:
and: (a,b)
(a,b) 1.651.65
mmol
mmol of initiator, (c,d) 1.10 mmol of initiator, and (e,f) 0.55 mmol of initiator. The insets present the size distribution
of initiator, (c,d) 1.10 mmol of initiator, and (e,f) 0.55 mmol of initiator. The insets present the size distribution histograms of
histograms of the PMMA microspheres of each sample.
the PMMA microspheres of each sample.
For better visualization, data in Table 1 are plotted in Figure 3a shows the diameter
The PMMA microspheres were also analyzed by means of a DLS system. The hydro-
of the PMMA microspheres vs. the amount of monomer at a constant quantity of initiator.
dynamic diameter (Dh ) values are listed in Table 1. It is found that the PMMA microsphere
It is seen a nonlinear tendency of the PMMA microsphere diameters as the mmols of mon-
diameters measured by DLS are greater than the ones given by the distribution histograms
omer are increased; this tendency is followed by the SEM and DLS diameters’ measure-
in the SEM micrographs (Figure 2). This difference is because the DLS measurements were
ments. This behavior is in accord to the results obtained by Waterhouse et.al. [1], Nandi-
carried out with PMMA microspheres in a water solution. Then, the formation of a water
yanto et.al. [6], and Yohanala et.al. [21], even when they synthesized PS (Table 2). In (b),
molecular layer on the PMMA microspheres’ surface must be considered. The formation of
it is seen that, for fixed amounts of monomer (285 and 380 mmol) with varying amount of
this layer is due to hydrogen bridge bonds or the PMMA microspheres’ surface protonation,
the initiator (0.55, 1.10, and 1.65 mmol), the PMMA microspheres’ diameter increases lin-
which gives rise to a positive surface charge. The corresponding polydisperse indices (PDI)
early. However, when the monomer amount is fixed at 190 mmol with varying amounts
are also listed in Table 1. In all cases, the PDI values are less than 0.1, indicating a uniform
of initiator (0.55, 1.10, and 1.65 mmol), the PMMA microspheres’ diameter decreases line-
distribution of the microspheres [29].
arly. This result is unexpected, even though it is in accord with the tendencies reported by
For better visualization, data in Table 1 are plotted in Figure 3a shows the diameter
Yamamoto et al. [20] who, for monomer amounts fixed at 28.3 and 14.2 g and with varying
of the PMMA microspheres vs. the amount of monomer at a constant quantity of initia-
amounts of initiator (0.034, 0.061, and 0.186 g), the microspheres Davg tended to increase.
tor. It is seen a nonlinear tendency of the PMMA microsphere diameters as the mmols
However, for a fixed monomer amount of 2.83 g with the same increasing amounts of
of monomer are increased; this tendency is followed by the SEM and DLS diameters’
measurements. This behavior is in accord to the results obtained by Waterhouse et al. [1],
Nandiyanto et al. [6], and Yohanala et al. [21], even when they synthesized PS (Table 2).
In (b), it is seen that, for fixed amounts of monomer (285 and 380 mmol) with varying
amount of the initiator (0.55, 1.10, and 1.65 mmol), the PMMA microspheres’ diameter
increases linearly. However, when the monomer amount is fixed at 190 mmol with varying
amounts of initiator (0.55, 1.10, and 1.65 mmol), the PMMA microspheres’ diameter de-
creases linearly. This result is unexpected, even though it is in accord with the tendencies
reported by Yamamoto et al. [20] who, for monomer amounts fixed at 28.3 and 14.2 g and
with varying amounts of initiator (0.034, 0.061, and 0.186 g), the microspheres Davg tended
to increase. However, for a fixed monomer amount of 2.83 g with the same increasing
amounts of initiator, Davg tends to decrease. This unexpected result is in disagreement
with that reported by Waterhouse et al. [1], probably because they varied two parameters
Polymers 2021, 13, x FOR PEER REVIEW 7 of 16

Polymers 2021, 13, 2171 7 of 16

initiator, Davg tends to decrease. This unexpected result is in disagreement with that re-
ported by Waterhouse et al. [1], probably because they varied two parameters simultane-
simultaneously: the amount of monomer and temperature (Table 2). In this regard, we
ously: the amount of monomer and temperature (Table 2). In this regard, we emphasize
emphasize that our experiments are more reliable than the ones by Waterhouse et al. [1],
that our experiments are more reliable than the ones by Waterhouse et al. [1], since we
since we varied one synthesis parameter at a time.
varied one synthesis parameter at a time.

PMMA Microsphere Diameter (nm)


PMMA Microsphere Diameter (nm)

420 190 mmol *SEM


1.65 mmol 440
285 mmol
1.10 mmol
390 380 mmol
0.55 mmol 400
360
360
330
*
300 * 320 *
*
270
280 *
240
*SEM
240
*
180 210 240 270 300 330 360 390 0.6 0.8 1.0 1.2 1.4 1.6
Monomer mmol Initiator mmol
(a) (b)
Figure 3.
Figure 3. (a)
(a) A
Aplot
plotofofmicrosphere
microspherediameter
diametervs.vs.
MMA mmol
MMA andand
mmol (b) (b)
Microsphere diameter
Microsphere vs. initiator
diameter mmol.
vs. initiator The diam-
mmol. The
eters were measured by SEM (*) and DLS techniques.
diameters were measured by SEM (*) and DLS techniques.

In Table 1, the 𝜁 potential values obtained from microspheres of PMMA are listed;
Table 2. Effect of the variation of monomer and initiator in the synthesis of various polymers using the SFEP method.
they lie in the 41.4 to 51.1 mV range. Therefore, it is concluded that the dispersions are
very stable [29]. It is also important to Trend
Davg mention that no relation was found between the
Davg Trend
Polymer Initiator Monomer T (°C) Reference
reaction variables used in this study.Monomer
Yamamoto and Higashitani Initiator
[30] have suggested that
V-50 a positive 𝜁 potential for the PMMA microspheres is associated to the
It decreases withfunctional groups
Methylstyrene
PMS 4.23 mM in the initiator V50. If70this were the case, the bands corresponding to C-N and N-H
increasing initiator vibra-
[19]
424 mM
22.3 mM tions would be present in the FT-IR and Raman spectra. Since, concentration.
in the corresponding spec-
tra of our microspheres set, these Davg bands arefrom
increases not present
The (Figure 1), we infer that the 𝜁
effect of initiator
potential
Methyl
is instead positive due to
364 the
to 415protonation
nm when ofamount
the carboxyl group, as is illustrated
is not conclusive
V-50 schematically 4 and confirmed with the pH= 3.30 measured of
monomer amount because amount
PMMA methacrylate in Figure
70 to 80 from the colloidal
[1]
0.75 to 3.0 g PMMA microspheres. increases from 300 to monomer and
300 to 400 mL
400 mL at fixed 1.5 g of temperature is varied
initiator and T = 70 °C. simultaneously.
Table 2. Effect of the variation of monomer and initiator in the synthesis of various polymers using the SFEP method.
For fixed monomer
Davg Trend Davg Trend
amounts: 28.3, 14.2 g,
Polymer Initiator
V-50
Monomer
Methyl
T (℃) with increasing
Reference
Monomer Initiatorinitiator,
PMMA 0.034, 0.061, methacrylate 70 Davg increases. For fixed [20]
V-50 It decreases with increas-
0.186 g 28.3, 14.2, 2.83 g
Methylstyrene monomer amount: 2.83 g,
PMS 4.23 mM 70 ing initiator
with concentra-
increasing initiator, [19]
424 mM
22.3 mM tion.
Davg decreases.
DavgForincreases from of
each amount 364 Foreffect
each amount of
The of initiator
to initiator,
415 nm the when mono- monomer, the initiator
monomer
Methyl methac- to three amount was is not conclusive
V-50
AIBA Styrene merwas variedincreases
amount varied to four
PMMA rylate 70 to 80 quantities. A similar because amount of mon-
quantities. A similar [1]
PS 0.75 to
0.0008, 3.0 g
0.004, 0.40, 0.80, 55 to 90 from 300 to 400 mL at [6]
0.008, 0.04 wt%
3002.0towt%
400 ml trend was observed: omer and temperature
trend was observed: theis
fixed
the D1.5 g of
was initiator
increased D was increased as
avg varied
avg simultaneously.
and = 70 ℃.
T monomer
as the the initiator amount was
V-50 Methyl methac- amount was increased. increased.
For fixed monomer
PMMA 0.034, 0.061, rylate 70 amounts: 28.3, 14.2 g, [20]
0.186 g 28.3, 14.2, 2.83 g with increasing initiator,
Polymers 2021, 13, 2171 8 of 16
Polymers 2021, 13, x FOR PEER REVIEW 8 of 16

Table 2. Cont.
Davg increases. For fixed
Davg Trend monomer Damount:
avg Trend2.83 g,
Polymer Initiator Monomer T (°C) Reference
Monomer Initiator
with increasing initiator,
The monomer was DWhen
avg decreases.
the styrene
For eachvariedamount to three
of ini- monomer concentration
quantities. The Davg ofFor each amount of mon-
tiator, the monomer was set constant at 10%
Styrene polystyrene spheres omer, the initiator was
KPS
AIBA was varied to 223
three (v/v), the Davg of the PS
PS 5%, 10%, 14% 80 increased to [21]
0.05, 0.2 g Styrene varied to four
particles quantities.
decreased from
0.0008, 0.004, (v/v) 316 nm when
55 to 90 quantities. styrene
A similar
PS 0.40, 0.80, 2.0 249 nmtrend
A similar usingwas
0.05 g of
ob- [6]
0.008, 0.04 trend was monomer
observed: initiator, to 181 nm using
wt% concentration was served: the Davg was in-
wt% the Davg was increased
increased.
0.2 g of initiator.
creased as the initiator
as the monomer
For each amount of amount was increased.
amount was increased.
initiator, the monomer For fixed monomer
was varied to three When the styrene
amounts: 380, 285mono-
mmol,
Methyl The monomer was var-
V-50 quantities. In all cases a mer with increasing initiator,
concentration was
methacrylate ied tosimilar
three quantities. This
PMMA 0.55, 1.10, 75 non-linear set D avg increased.
constant at 10% For(v/v),
fixed
190, 285, work
1.65 mmol Styrene avg ofwas
The Dtrend polystyrene
observed: monomer amount: 190 g,
KPS 380 mmol
the Davg is increased
the Davg of the PS parti-
PS 5%, 10%, 14% 80 spheres increased 223as with increasing initiator, [21]
0.05, 0.2 g the monomer amount cles decreased
D avg
from 249
decreased.
(v/v) to 316 nm when sty-
is increased. nm using 0.05 g of initia-
rene monomer concen-
PMS: poly (methyl styrene); PS: poly (styrene); PMMA: poly (methyl tor, to 181 nm using 0.2 g
methacrylate).
tration was increased.
of initiator.
In Table 1, the ζ potential values obtained
For each amount of ini- from microspheres of PMMA are listed;
they lie in the 41.4 to 51.1 mV range. Therefore,
tiator, the monomer it is concluded
For fixed monomer that the dispersions are
very stable [29]. It is also important to mention that no
was varied to three amounts: 380, 285 mmol,relation was found between the
Methyl
reaction methac-
variables used in quantities.
this study.InYamamoto and Higashitani [30] have suggested
V-50 all cases with increasing initiator,
rylate
that a positive ζ potential for the PMMA microspheres is associated to the functional This
PMMA 0.55, 1.10, 75 a similar non-linear Davg increased. For fixed
190, 285,
groups in380
the initiator V50. If this were the case, the bands corresponding to C-Nwork and N-H
1.65 mmol trend was observed: monomer amount: 190 g,
vibrations
mmol would be present in the FT-IR and Raman spectra. Since, in the corresponding
spectra of our microspheres the Davg
set, is increased
these bands areasnot with increasing
present (Figureinitiator,
1), we infer that the ζ
the monomer amount D avg decreased.
potential is instead positive due to the protonation of the carboxyl group, as is illustrated
schematically in Figure 4 and confirmed is increased. with the pH = 3.30 measured from the colloidal
PMS: polyPMMA
(methyl microspheres.
styrene); PS: poly (styrene); PMMA: poly (methyl methacrylate).

Figure 4. The schematic protonation reaction of the carboxyl group that surrounds the PMMA
Figure 4. The schematic protonation reaction of the carboxyl group that surrounds the PMMA mi-
microsphere
crosphere surface.
surface.
3.1.3. DRS Analysis
3.1.3. DRS Analysis
Figure 5 shows diffuse reflectance spectra (expressed as absorbance) in the UV-Vis-NIR
Figure
range 5 shows
of the PMMA diffuse reflectance
microspheres spectra (expressed
in powder as absorbance)
form, synthesized in the UV-Vis-
with different reaction
NIR range of the PMMA microspheres in powder form, synthesized with different
conditions. All spectra in the UV region (200 to 400 nm) show a pronounced absorbance reac-
tion conditions. All spectra in the UV region (200 to 400 nm) show a pronounced absorb-
edge while, in the Vis region (400 to 800 nm), the absorption is small and almost constant.
ance edge while, in the Vis region (400 to 800 nm), the absorption is small and almost
Polymers 2021, 13, 2171 9 of 16

In the NIR region (800 to 2500 nm) seven bands related to the vibrations of the C-H and C-O
bonds are present. A detailed analysis of the bands in the NIR region is described below.

0.8 UV Visible Near Infrared


PMMA-1
0.7 PMMA-2
PMMA-3
0.6
Absorbance (u.a.)
PMMA-4
PMMA-5
0.5
PMMA-6
PMMA-7
0.4
PMMA-8
0.3 PMMA-9

0.2

0.1

0.0
200 400 600 800 1000 1200 1400 1600 1800 2000 2200 2400
Wavelength (nm)
Figure 5. UV-Vis-NIR diffuse reflectance spectra (shown as absorbance) of PMMA microspheres.

In the NIR region, the more intense bands are located at 1173, 1425, 1679, 1912, 2136,
2256, and 2377 nm. The band at 1173 nm is assigned to the second harmonics of the
fundamental stretching vibrations of the C-H, CH3 and CH2 bonds. In this region (800
to 1350 nm), the bands are of low intensity compared to the bands located at longer
wavelengths due to the characteristic harmonics of the involved vibrations [31]. The CH3
and CH2 groups are also responsible for the band located at 1425 nm, which is due to
combined stretching vibrations and deformation of the C-H bond. The bands at 1679 and
1697 nm are assigned to fundamental stretching vibrations of the C-H bond of the CH3
group, while the bands at 1717 and 1779 nm correspond to the first harmonic of the CH2
group [32]. The band located at 1912 has been assigned to the second overtone of the
stretching of the C=O bond, while the band located at 2133 nm is due to a combination
of stretching vibrations of the C-H bond of the OCH3 groups and to the stretching of
the C=O of the ester [31]. Finally, the bands located at 2256 and 2377 nm are due to a
combination of the stretching and flexing vibrations of the C-H bond in the CH3 and CH2
groups, respectively [33]. These results confirm that the C-N bonds are absent due to the
initiator and the positive charge indicated in the results of the ζ potential, which is itself
due to the protonation of the carboxyl group.
In the UV region, we can see a sharp absorption edge of about 270 nm due to elec-
tronic transitions from n level orbital (HOMO, highest occupied molecular orbital) to σ∗
level (LUMO, lowest unoccupied molecular orbital), according to the molecular orbitals
theory [25]. This absorption edge is also associated with the energy band gap that is the
difference in energy between the minimum of the conduction band and the maximum of
the valence band. The Eg values of the PMMA microspheres were determined using the
Kubelka–Munk (K-M) formalism [34] given by Equation (1):

[ F ( R∞ )hv]1/n = C hv − Eg ,

(1)

where h is the Planck constant, ν is the electromagnetic radiation frequency, C is a pro-


portionality constant, and the n value depends on the nature of the electronic transition
of the material. In the case of a compound with a direct energy band gap, n = 2, and for

1
for an indirect energy band gap, 𝑛 = 1/2. 𝐹(𝑅 ) is the the K–M function, which relates
the dispersion and absorption coefficients given by Equation (2):
Polymers 2021, 13, 2171 10 of 16
(1 − 𝑅 ) 𝐾
𝐹(𝑅 ) = = (2)
2𝑅 𝑆
where 𝑅 is the ratio between the reflectance of the sample to that of the reference:
an indirect energy band gap, n = 1/2. F ( R∞ ) is the the K–M function, which relates the
dispersion and absorption coefficients given by 𝑅 Equation (2):
𝑅 = (3)
𝑅
2
(1 − R ∞ ) K
It is widely accepted in the F ( R∞ ) = to determine
literature = the Eg values of PMMA consider- (2)
2R∞ S
ing both cases of direct and an indirect- band gap energy.
where TheR∞Eis g values were
the ratio determined
between by plotting
the reflectance 𝐹(𝑅sample
of the )ℎ𝜈 /to that
vs. ℎ𝜈, making
of the a linear fit
reference:
to the absorption edge, and the Eg value was obtained by the intersection to the ℎ𝜈 axis.
Figure 6 shows an 𝐹(𝑅 )ℎ𝜈 / vs. ℎ𝜈 plotRconsidering sample in (a-c) a direct transition, and
R∞ = (3)
in (d-f) an indirect transition is considered. The Rre fobtained
erence Eg values are listed in Table 2.

In Table 3, it is seen that the Eg values for a direct energy band gap are greater than
thoseItofisan
widely accepted
indirect in the literature
band structure. These to determine
values are close the to
Egthose
values ofof PMMA
Aziz et al.considering
[25], who
both cases of direct and an indirect- band gap energy.
reported 5.04 eV for a direct energy band gap and 4.8 eV for an indirect one in relation to
PMMA Thethin
Eg values were determined
films. However, by plotting
these values [ F ( Rthan
are smaller ∞ ) hν ]1/n
the vs. hν,
7.0996 eVmaking a linear
theoretical valuefit
to the absorption
calculated by Hazimedge,
et and the Eusing
al. [35] g value
thewas obtained
density by thetheory.
functional intersection to the hν axis.
The difference in
PMMA E values is due to fact 1/n
that PMMA microspheres are porous and present punctual
Figure 6 shows an [ F ( R∞ )hν]
g vs. hν plot considering in (a–c) a direct transition, and in
defects
(d–f) anin their structures
indirect transitionasisgrain frontiers,
considered. Theoxygen
obtainedvacancies, etc.,are
Eg values which
listedare
inthe same
Table 2. as
those presented by amorphous SiO2 microspheres [10].

180 180
180 180 180 180

150 150 150

150 120
150 120 150 120
[F(R∞) hν]2 (eV)2
[F(R∞) hν]2 (eV)2

90

[F(R∞) hν]2 (eV)2


90
90
60 60

120 30 120 60
120 30

0 30 0
2 3 4 5 6 2 3 4 5 6
0
90 90 2 3 4 5 6 90

60 60 60

PMMA-1 PMMA-4 PMMA-1


30 30 30
PMMA-2 PMMA-5 PMMA-2
PMMA-3 PMMA-6 PMMA-3
0 0 0
4.8 5.0 5.2 5.4 5.6 5.8 6.0 6.2 4.8 5.0 5.2 5.4 5.6 5.8 6.0 4.8 5.0 5.2 5.4 5.6 5.8 6.0 6.2
hν (eV) hν (eV) hν (eV)
(a) (b) (c)
4 4
4
3.5 3.5
3.5
3 3
3
3.0 3.0
[F(R∞) hν ]1/2 (eV)1/2
[F(R∞) hν ]1/2 (eV)1/2

[F(R∞) hν]1/2 (eV)1/2

3.0 2 2
2

2.5 1 2.5 2.5


1
1
0

2.0
3.0 3.5 4.0 4.5 5.0 5.5 6.0 2.0 0 2.0
2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0
0
2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0 6.5

1.5 1.5 1.5

1.0 1.0 1.0


PMMA-4 PMMA-7
PMMA-1
0.5 PMMA-5 0.5 PMMA-8
0.5 PMMA-2
PMMA-6 PMMA-9
PMMA-3
0.0 0.0 0.0
4.2 4.4 4.6 4.8 5.0 5.2 5.4 5.6 5.8 6.0 4.4 4.6 4.8 5.0 5.2 5.4 5.6 5.8 4.4 4.6 4.8 5.0 5.2 5.4 5.6 5.8
hν (eV) hν (eV) hν (eV)
(d) (e) (f)
Figure6.6.Kubelka–Munk
Figure Kubelka–Munk transform
transform reflectance
reflectance spectra
spectra of
of PMMA
PMMA microspheres
microspheres ininpowder:
powder:(a–c)
(a–c)considering
consideringa adirect
directband
band
structure, and (d–f) considering an indirect band structure. From these spectra, the Eg values of the PMMA were deter-
structure, and (d–f) considering an indirect band structure. From these spectra, the Eg values of the PMMA were determined.
mined.
In Table 3, it is seen that the Eg values for a direct energy band gap are greater than
thoseTo
of determine
an indirectwhich
band transition
structure. type
Theseoccurs in are
values the close
PMMA microspheres,
to those of Aziz etwe
al.used
[25],two
who
approaches, according to Gupta et al. [36]; when the absorption coefficient
reported 5.04 eV for a direct energy band gap and 4.8 eV for an indirect one in relation𝛼 value is
greater than 10 cm , electronic direct transition occurs, while, when 𝛼 <
to PMMA thin films. However, these values are smaller than the 7.0996 eV theoretical 10 cm , an
value calculated by Hazim et al. [35] using the density functional theory. The difference in
PMMA Eg values is due to fact that PMMA microspheres are porous and present punctual
defects in their structures as grain frontiers, oxygen vacancies, etc., which are the same as
those presented by amorphous SiO2 microspheres [10].
Polymers 2021, 13, 2171 11 of 16

Polymers 2021, 13, x FOR PEER REVIEW


Table 3. Direct energy band gap (Eg/D ), indirect energy band gap (Eg/I ), Urbach energy (Et ),11and
of 16
absorption constant (α0 ) of the obtained PMMA microspheres.

Sample E (eV) E (eV) Et (eV) α0 10−5 (cm−1 )


indirect transition takesg/Dplace. Figure 7 shows
g/I
the absorption coefficient vs. incident radi-
PMMA-1 5.23 ± 0.06 4.68 ± 0.01
ation. It is observed that the absorption coefficient is less0.484
than 10 cm in2.030 the energy
PMMA-2 5.22 ± 0.05 4.54 ± 0.03 0.509 2.870
range 4.5 to 5.7 eV; thus, we infer that the energy band transition in the PMMA under
PMMA-3 5.15 ± 0.06 4.36 ± 0.02 0.558 3.500
study is indirect. Additionally,
PMMA-4 5.19 ±0.07 the energy4.65 ±band
0.05 structure 0.522
of polymers has been
3.920 investi-
gated by calculating 5.22
PMMA-5 the Urbach
± 0.09 energy4.71using Equation (4):0.403
± 0.04 2.670
PMMA-6 5.23 ± 0.06 4.75 ± 0.03 ℎ𝜐 0.479 1.500
PMMA-7 5.21 ± 0.04 𝛼 =±
4.61 𝛼 0.02
𝑒𝑥𝑝 0.544 5.089 (4)
PMMA-8 5.34 ± 0.05 4.65 ± 0.03 E 0.537 4.992
PMMA-9 ± 0.06 tail.
5.35energy
where E is the Urbach 4.73 ± 0.03 0.403 0.155

Table
To3.determine
Direct energy band
which gap (Eg/D),type
transition indirect energy
occurs band
in the gap (Emicrospheres,
PMMA g/I), Urbach energy
we (E t), and
used ab-
two
sorption constant (𝛼 ) of the obtained PMMA microspheres.
approaches, according to Gupta et al. [36]; when the absorption coefficient α value is greater
than 104 cm−1 , electronic direct
Eg/D transition occurs,
Eg/I while, when𝐄α𝐭 < 104 cm−1 , 𝜶an indirect
𝟎 𝟏𝟎
𝟓
Sample
transition takes place. Figure 7 shows the absorption coefficient vs. incident radiation. It
(eV) (eV) (eV) (cm )−1

is observed that the absorption coefficient is less than 104 cm−1 in the energy range 4.5 to
PMMA-1 5.23 ± 0.06 4.68 ± 0.01 0.484 2.030
5.7 eV;PMMA-2
thus, we infer that the
5.22 energy band4.54
± 0.05 transition
± 0.03 in the PMMA0.509 under study 2.870
is indirect.
Additionally,
PMMA-3the energy band
5.15 structure of4.36
± 0.06 polymers
± 0.02 has been0.558investigated by calculating
3.500
the Urbach
PMMA-4energy using Equation
5.19 ±0.07 (4): 4.65 ± 0.05 0.522 3.920
PMMA-5 5.22 ± 0.09 4.71 ±  0.04  0.403 2.670
PMMA-6 5.23 ± 0.06 4.75 ± hυ
0.03 0.479 1.500 (4)
α = α0 exp
PMMA-7 5.21 ± 0.04 4.61 ± 0.02Et 0.544 5.089
PMMA-8 5.34 ± 0.05 4.65 ± 0.03 0.537 4.992
where Et is the Urbach energy tail.
PMMA-9 5.35 ± 0.06 4.73 ± 0.03 0.403 0.155

2.5

PMMA-1
Absorption Coefficient (cm-1)

PMMA-2
2.0 PMMA-3
PMMA-4
PMMA-5
PMMA-6
1.5
PMMA-7
PMMA-8
PMMA-9
1.0

0.5

0.0
2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0
hν (eV)
Figure7.7. Variation
Figure Variation of
of absorption coefficient(𝛼)
absorptioncoefficient (α) as
asaafunction
functionofofphoton
photonenergy (ℎ𝜈)
energy (hνof PMMA
) of mi-
PMMA
crospheres.
microspheres.

ItIthas
hasbeen
beendemonstrated
demonstratedthat thetheF (𝐹(𝑅
that )~𝛼
R∞ ) ∼ approximationisisconsidered
α approximation consideredprecise
precise
enoughtotoevaluate
enough evaluate the
the optical
optical properties
properties of materials
of materials via via specular
specular reflectance
reflectance [37];[37]; there-
therefore,
fore, Equation
Equation (4) is approximated
(4) is approximated to: to:
ℎ𝜐
𝐹(𝑅 ) = 𝛼 𝑒𝑥𝑝
 
hυ (5)
F ( R∞ ) = α0 exp E (5)
Et
where 𝐹(𝑅 ) is the optical absorption coefficient, 𝛼 is a constant, ℎ𝜈 is the incident
photon energy, and E is the Urbach energy that refers to the width of tails of the localized
states. E can be determined by taking the reciprocal slope of the straight line in the ln 𝛼
Polymers 2021, 13, 2171 12 of 16

where F ( R∞ ) is the optical absorption coefficient, α0 is a constant, hν is the incident photon


Polymers 2021, 13, x FOR PEER REVIEW
energy, and Et is the Urbach energy that refers to the width of tails of the localized states. 12 of 16 E
t
can be determined by taking the reciprocal slope of the straight line in the ln α vs. hν plot.
The corresponding results are listed in Table 3. In this table, it is seen that the Eg values
vs. ℎ𝜈 plot. The
considering either a direct or an
corresponding indirect
results energy
are listed inband
Tablegap
3. Indecrease
this table, asitEis
t values increase.
seen that the
Since
E theconsidering
g values Urbach energy eitherisathe band
direct tailindirect
or an of localized
energystates
band ingapthe energyasband
decrease gap, it
Et values
is expected
increase. that,
Since theinUrbach
the PMMA,
energyindirect transitions
is the band take place
tail of localized first;inthen,
states with the
the energy band help
of
gap,
 phonons,
it is the
expected
 transition
that, in theto the
PMMA,conduction
indirect band occurs
transitions (LUMO).
take place As
first; a matter
then, with of fact,
the
help of phonons, the transition to the
Eg/I + Et is approximately equal to Eg/D [38]. conduction band occurs (LUMO). As a matter of
fact, E / E is approximately equal to E / [38].
3.2. PMMA Opals
3.2. PMMA Opals
Figure 8 shows SEM micrographs of PMMA microspheres forming an opal or colloidal
Figure
crystal. In 8(a)
shows SEM micrographs
at ×25,000, we observe ofthe
PMMA microspheres forming
self-organization an and
properties opal anor colloi-
intrinsic
dal
disorder that is inevitable in the self-ensemble process of colloidal spheres. Inintrinsic
crystal. In (a) at ×25,000, we observe the self-organization properties and an (b), some
disorder that iscontinuous
regions with inevitable in the self-ensemble
periodic process
array can be seen of
incolloidal spheres.
more detail. In In
(c),(b),
onsome
a SEM
regions with continuous periodic array can be seen in more detail. In (c), on a SEM micro-
micrograph at ×100,000, a face center cubic (fcc) structure formed by the PMMA spheres
graph at ×100,000, a face center cubic (fcc) structure formed by the PMMA spheres with
with the (111) planes parallel to the subjacent substrate is delineated. This structure may
the (111) planes parallel to the subjacent substrate is delineated. This structure may not be
not be a pure fcc with a repeated ABC stacking sequence along the growth direction [39]
a pure fcc with a repeated ABC stacking sequence along the growth direction [39] since,
since, according to the studies of Versmold [40] and Verhaegh et al. [41], the most common
according to the studies of Versmold [40] and Verhaegh et al. [41], the most common type
type or structure in shear-ordered colloidal crystals is the close random stacking of close-
or structure in shear-ordered colloidal crystals is the close random stacking of close-
packed planes.
packed planes.

(a) (b) (c)

Figure8.8. SEM
Figure SEM images of PMMA
PMMA opals
opalsprepared
preparedwith
withPMMA_4
PMMA_4microspheres
microspheresshowing self-assembly
showing of the
self-assembly microspheres:
of the microspheres:
(a) 10,000×, (b) 25,000× and (c) 100,000×.
(a) 10,000×, (b) 25,000× and (c) 100,000×.

Figure
Figure9a–c
9a–cshows
showsspecular
specularreflectance
reflectancespectra
spectraofof thetheopal photonic
opal photonic crystals vs.vs.
crystals wave-
wave-
length, measured
measured at at different
different incident
incident angles
angles (20 70◦∆=
(20◦toto70°, , ∆10°) ◦ ) and
= 10and prepared
preparedwith single
with single
microspheres
microspheres labeled
labeledasasPMMA-7,
PMMA-7,PMMA-8,
PMMA-8,and andPMMA-9.
PMMA-9.Due Duetotothethe
high periodicity
high of of
periodicity
PMMA microspheres that constitute each opal, they exhibit narrow photonic
PMMA microspheres that constitute each opal, they exhibit narrow photonic band gaps band gaps
(PBGs)
(PBGs) in
in the
the (111)
(111)direction.
direction.
The stop band position of the opal photonic crystal satisfices the Bragg–Snell law
given by Equation (6), [1]:
/
𝜆 = 2(𝑑 /𝑚) 𝑛 − 𝑠𝑖𝑛 𝜃 (6)

where 𝜆 is the wavelength position of the Γ − 𝐿 stop bands, 𝑑 is the distance be-
tween the h, k, l planes, 𝑚 is the order of the Bragg diffraction, 𝑛 is the effective re-
fractive index of the material, and 𝜃 is the incident angle of light during measurements.
The effective refractive index is calculated using Equation (7):
𝑛 =𝜑 𝑛 (1 − 𝜑)𝑛 (7)

where 𝜑 is the solid volume fraction for fcc, the structure is 𝜑 = 0.74, 𝑛 , and 𝑛 ,
which are the refractive indices of the PMMA (1.492) and of air (1.000), then 𝑛 = 1.364.
In the case of first order diffraction from (111) planes, 𝑚 = 1 and 𝑑 = 𝑑 = 0.816𝐷,
where 𝐷 is the diameter of the PMMA microspheres. Then Equation (6) takes on the form
of:
From the 𝜆 vs. 𝑠𝑖𝑛 𝜃 plot shown in Figure 9d, it is seen that a Bragg diffraction
condition is satisfied because this is a straight-line result; moreover, the slope is equal to
−(2𝑑 ) and the 𝑦 −axis intercept is (2𝑑 ) 𝑛 . Then, from this relation, it was easy
to calculate the interplanar distance (𝑑 ), the effective refractive index (𝑛 ), the diam-
Polymers 2021, 13, 2171 eter of the PMMA microspheres (D), and the solid volume fraction (𝜑). These results 13 are
of 16
listed in Table 4.

35
(a) PMMA-7 (b) PMMA-8
35
20° 20°

Reflectance (%)
28 30°

Reflectance (%)
30°
28 40° 40°
50° 50°
21 60°
21 60°
70°
70°
14
14

7 7

0 0
400 500 600 700 800 900 1000 400 500 600 700 800 900 1000
Wavelength (nm) Wavelength (nm)

50 (c) PMMA-9 0.8 (d)


20°
Reflectance (%)

40 30°
0.7
40°
50°

λ2 (nm)
30 60°
70° 0.6

20
0.5
10 PMMA-7
PMMA-8
0.4
PMMA-9
0
400 500 600 700 800 900 1000
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
Wavelength (nm) sin2θ

Figure 9. Specular
Figure 9. reflectance spectra
Specular reflectance spectra of
of the
the opals’
opals’photonic
photoniccrystals
crystals(a)
(a)PMMA-7,
PMMA-7,(b)
(b)PMMA-8,
PMMA-8,and
and
(c) PMMA-9. In (d), the linear behavior of the stop bands’ angle dependent can be observed.
(c) PMMA-9. In (d), the linear behavior of the stop bands’ angle dependent can be observed.

The
Figurestop band position
9 shows the pseudoof the opal photonic
photonic band gapcrystal of eachsatisfices the Bragg–Snell
opal photonic law
crystal. These
given by Equation (6), [1]:
pseudo photonic band gaps are formed because the refractive indices ratio is less than 2.9,
a required value to obtain a complete stop-band  [1]. In astop-band,
1/2 all light is reflected
and no propagated mode in λ the =medium n2e f f −In
2(dhkl /mis) found. 2
sinFigure
θ 9a–c, it is seen that the posi-(6)
tions of the pseudo stop-bands are shifted towards shorter wavelengths as 𝜃 increases
where λ is to
according theEquation
wavelength(8). position of the Γ − L stop bands, dhkl is the distance between
the h, k, l planes, m is the order of the Bragg diffraction, ne f f is the effective refractive index
Table
of the 4. Wavelength
material, andposition
θ is theofincident
opals’ stop bandof(𝜆),
angle intercept
light during and slope fit data, interplanar
measurements. dis-
The effective
tance (d111).index
refractive diameter of the PMMA
is calculated microspheres
using Equation (D),
(7): and the solid volume fraction (𝜑) of the ob-
tained opals’ photonic crystals.

𝝀 (𝜽 = 20°) a ne f f = ϕ × dn111
PMMA
b +D(1b − ϕ)n air 𝝀𝒎𝒂𝒙 (𝜽 = 20°) b(7)
Sample Intercept Slope 𝒏𝒆𝒇𝒇 b 𝝋b
(nm) (nm) (nm) (nm)
where ϕ is the solid volume fraction for fcc, the structure is ϕ = 0.74, n PMMA , and n air ,
PMMA-7 883 0.8248 −0.4094 320 ± 22 392 ± 9 1.420 ± 0.02 0.852 880 ± 14
which are the refractive indices of the PMMA (1.492) and of air (1.000), then ne f f = 1.364. In
PMMA-8 814 0.7072 −0.3586 300 ± 18 367 ± 11 1.404 ± 0.01 0.820 816 ± 18
the case of first
PMMA-9
order diffraction
803
from (111) planes, m = 1 and dhkl =0.820
0.6861 −0.3482 295 ± 15 361 ± 7 1.403 ± 0.03
d111 = 0.816D,
803 ± 22
where
D is the diameter of the PMMA microspheres. Then Equation (6) takes on the form of:
p
λ = 1.632D 1.3642 − sin2 θ (8)

From the λ2 vs. sin2 θ plot shown in Figure 9d, it is seen that a Bragg diffraction
condition is satisfied because this is a straight-line result; moreover, the slope is equal to
−(2d111 )2 and the y-axis intercept is (2d111 )2 n2e f f . Then, from this relation, it was easy to
calculate the interplanar distance (d111 ), the effective refractive index (ne f f ), the diameter
of the PMMA microspheres (D), and the solid volume fraction (ϕ). These results are listed
in Table 4.
Polymers 2021, 13, 2171 14 of 16

Table 4. Wavelength position of opals’ stop band (λ), intercept and slope fit data, interplanar distance (d111 ). diameter of
the PMMA microspheres (D), and the solid volume fraction (ϕ) of the obtained opals’ photonic crystals.
◦ ◦
λ (θ = 20 ) a d111 b Db b λmax (θ = 20 ) b
Sample Intercept Slope neff ϕb
(nm) (nm) (nm) (nm)
PMMA-7 883 0.8248 −0.4094 320 ± 22 392 ± 9 1.420 ± 0.02 0.852 880 ± 14
PMMA-8 814 0.7072 −0.3586 300 ± 18 367 ± 11 1.404 ± 0.01 0.820 816 ± 18
PMMA-9 803 0.6861 −0.3482 295 ± 15 361 ± 7 1.403 ± 0.03 0.820 803 ± 22
a Data obtained directly from specular reflectance spectra. b Data calculated from Bragg–Snell law.

Figure 9 shows the pseudo photonic band gap of each opal photonic crystal. These
pseudo photonic band gaps are formed because the refractive indices ratio is less than 2.9, a
required value to obtain a complete stop-band [1]. In a stop-band, all light is reflected and
no propagated mode in the medium is found. In Figure 9a–c, it is seen that the positions of
the pseudo stop-bands are shifted towards shorter wavelengths as θ increases according to
Equation (8).
The position of the stop bands at a 20◦ incident angle of the opals’ photonic crystals
are: PMMA-7 at 883 nm, PMMA-8, 814 nm, and PMMA-9 at 803 nm. It is observed that the
largest stop-band position corresponds to the sample with the largest diameter, PMMA-
7 (323 nm). These results agree with those shown in Table 1, confirming an increment
in size of the PMMA microspheres as the initiator amount increases. Furthermore, we
can infer from these experimental results that the PBGs of the opals’ photonic crystals
can be designed in accordance with the size of the PMMA microspheres throughout the
optimization of initiator and monomer optimal amounts.
From the data in Table 4, we note that the calculated solid volume fractions exceed the
theoretical value of 0.74 for an fcc sphere from hard materials. This is probably because
PMMA spheres are considered soft material spheres [42].

4. Conclusions
PMMA microspheres were synthesized by means of the surfactant-free emulsion
polymerization method by systematically varying the amount of initiator and the amount
of monomer. The proportion of these precursors influences the diameter of PMMA micro-
spheres. As the mmol of monomer and initiator increases, the microspheres’ size increases,
except when a minimum quantity of monomer (190 mmol) is used, since the monomer acts
as a limiting reactant. For this particular monomer quantity, the PMMA microspheres’ size
remains practically constant. The obtained colloidal microspheres are highly stable and
present a surface with positive charge due to the protonation of the carboxyl group of the
PMMA. The microspheres under study present an energy band gap less than the theoretical
one reported in the literature, since PMMA microspheres are porous and present punctual
and structural defects. The Eg values were calculated considering both direct and indirect
energy band structures. The energy band gap and the Urbach energy values of PMMA
allowed us to conclude that that PMMA energy band gap structure was indirect.
All synthesized opal photonic crystals present narrow stop bands since the refractive
index contrast of the two media is less than the minimum required (2.9). The stop bands
shift towards larger wavelengths as the microsphere size increases. It was also observed
that PBGs shifted towards shorter wavelengths and the intensity decreased as the incident
angle of irradiation increased. The Bragg–Snell law is satisfied since the diffraction and the
reflection effects take place in the opal photonic crystal.

Author Contributions: E.S.-M., M.M.-A., E.G.-B. and J.A.L.-L. developed the methodology for
obtaining the samples and carried out the characterization of the samples. All the authors contributed
to the analysis and discussion of results, and to the writing of the manuscript. All authors have read
and agreed to the published version of the manuscript.
Polymers 2021, 13, 2171 15 of 16

Funding: This research was funded by Programa para el Desarrollo Profesional Docente de la
Secretaría de Educación Pública (PRODEP-SEP), and Vicerrectoría de Investigación y Estudios de
Posgrado-Benemérita Universidad Autónoma de Puebla (VIEP-BUAP).
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: The data presented in this study are available on request from the
corresponding author.
Acknowledgments: M.M.-A. acknowledges to Consejo Nacional de Ciencia y Tecnología, México
(CONACyT) for the Graduate Student Scholarship, Grant No. 712521.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Waterhouse, G.I.N.; Waterland, M.R. Opal and inverse opal photonic crystals: Fabrication and characterization. Polyhedron 2007,
26, 356–368. [CrossRef]
2. Aguirre, C.I.; Reguera, E.; Stein, A. Tunable colors in opals and inverse opal photonic crystals. Adv. Funct. Mater. 2010, 20,
2565–2578. [CrossRef]
3. Fudouzi, H. Optical properties caused by periodical array structure with colloidal particles and their applications. Adv. Powder
Technol. 2009, 20, 502–508. [CrossRef]
4. Liu, Y.; Qin, F.; Wei, Z.Y.; Meng, Q.B.; Zhang, D.Z.; Li, Z.Y. 10 Fs Ultrafast All-Optical Switching in Polystyrene Nonlinear Photonic
Crystals. Appl. Phys. Lett. 2009, 95, 1–4. [CrossRef]
5. Yablonovitch, E. Photonic band-gap structures. J. Opt. Soc. Am. B. 1993, 10, 283–295. [CrossRef]
6. Nandiyanto, A.B.D.; Suhendi, A.; Ogi, T.; Iwaki, T.; Okuyama, K. Synthesis of additive-free cationic polystyrene particles with
controllable size for hollow template applications. Colloids Surf. A Physicochem. Eng. Asp. 2012, 396, 96–105. [CrossRef]
7. Schroden, R.C.; Al-Daous, M.; Blanford, C.F.; Stein, A. Optical properties of inverse opal photonic crystals. Chem. Mater. 2002, 14,
3305–3315. [CrossRef]
8. Zhang, J.; Sun, Z.; Yang, B. Self-assembly of photonic crystals from polymer colloids. Curr. Opin. Colloid Interface Sci. 2009, 14,
103–114. [CrossRef]
9. Li, Z.; Wang, J.; Song, Y. Self-assembly of latex particles for colloidal crystals. Particuology 2011, 9, 559–565. [CrossRef]
10. Carmona-Carmona, A.J.; Palomino-Ovando, M.A.; Hernández-Cristobal, O.; Sánchez-Mora, E.; Toledo-Solano, M. Synthesis and
characterization of magnetic opal/Fe3 O4 colloidal crystal. J. Cryst Growth 2017, 462, 6–11. [CrossRef]
11. Fang, J.; Xuan, Y.; Li, Q. Preparation of polystyrene spheres in different particle sizes and assembly of the PS colloidal crystals. Sci.
China Technol. Sci. 2010, 53, 3088–3093. [CrossRef]
12. Gu, Z.Z.; Chen, H.; Zhan, S.; Sun, L.; Xie, Z.; Ge, Y. Rapid synthesis of monodisperse polymer spheres for self-assembled photonic
crystals. Colloids Surf. A Physicochem. Eng. Asp. 2007, 302, 312–319. [CrossRef]
13. Deng, T.; Chen, C.; Honoker, C.; Thomas, E.L. Two-dimensional block copolymer photonic crystals. Polymer 2003, 44, 6549–6553.
[CrossRef]
14. Kang, Y.; Walish, J.J.; Gorishnyy, T.; Tomas, E.L. Broad-wavelength-range chemically tunable block-copolymer photonic gels. Nat.
Mater. 2007, 6, 2007. [CrossRef]
15. Xiao, M.; Li, Y.; Allen, M.C.; Deheyn, D.D.; Yue, X.; Zhao, J.; Gianneschi, N.C.; Shawkey, M.D.; Dhinojwala, A. Bio-inspired
structural colors produced via self-assembly of synthetic melanin nanoparticles. ACS Nano 2015, 9, 5454–5460. [CrossRef]
16. Ali, U.; Karim, K.J.B.A.; Buang, N.A. A Review of the Properties and Applications of Poly (Methyl Methacrylate) (PMMA). Polym.
Rev. 2015, 55, 678–705. [CrossRef]
17. Camli, S.T.; Buyukserin, F.; Balci, O.; Budak, G.G. Size controlled synthesis of sub-100 nm monodisperse poly(methylmethacrylate)
nanoparticles using surfactant-free emulsion polymerization. J. Colloid Interface Sci. 2010, 344, 528–532. [CrossRef]
18. Gorsd, M.N.; Blanco, M.N.; Pizzio, L.R. Synthesis of Polystyrene Microspheres to be Used as Template in the Preparation of
Hollow Spherical Materials: Study of the Operative Variables. Procedia Mater. Sci. 2012, 1, 432–438. [CrossRef]
19. Tanrisever, T.; Okay, O.; Sönmezoǧlu, I.C. Kinetics of emulsifier-free emulsion polymerization of methyl methacrylate. J. Appl.
Polym. Sci. 1996, 61, 485–493. [CrossRef]
20. Yamamoto, T.; Higashitani, K. Growth processes of poly methylmethacrylate particles investigated by atomic force microscopy.
Adv. Powder Technol. 2007, 18, 567–577. [CrossRef]
21. Yohanala, P.T.F.; Mulaya Dewa, R.; Quarta, K.; Widiyastuti; Winardi, S. Preparation of Polystyrene Spheres Using Surfactant-Free
Emulsion Polymerization. Mod. Appl. Sci. 2015, 9, 121–126. [CrossRef]
22. Zidan, H.M.; Abu-Elnader, M. Structural and optical properties of pure PMMA and metal chloride-doped PMMA films. Phys. B
Condens. Matter. 2005, 355, 308–317. [CrossRef]
23. Mergen, Ö.B.; Arda, E.; Kara, S.; Pekcan, Ö. Effects of GNP addition on optical properties and band gap energies of PMMA films.
Polym. Compos. 2019, 40, 1862–1869. [CrossRef]
Polymers 2021, 13, 2171 16 of 16

24. Trejo-García, P.M.; Palomino-Merino, R.; De la Cruz, J.; Espinosa, J.E.; Aceves, R.; Moreno-Barbosa, E.; Portillo-Moreno, O.
Luminescent properties of Eu3+ -doped hybrid SiO2 -PMMA material for photonic applications. Micromachines 2018, 9, 441.
[CrossRef] [PubMed]
25. Aziz, S.B.; Abdullah, O.G.; Hussein, A.M.; Ahmed, H.M. From insulating PMMA polymer to conjugated double bond behavior:
Green chemistry as a novel approach to fabricate small band gap polymers. Polymers 2017, 9, 626. [CrossRef]
26. Sánchez-Mora, E.; Fernádez-Candelario, M.; Gómez-Barojas, E.; Pérez-Rodríguez, F. Influence of Fe ions on the optical properties
of Fe-ZnO inverse opals. J. Supercond. Nov. Magn. 2013, 26, 2447–2449. [CrossRef]
27. Kaczmarczyk, B.; Morejko-Buz, B.; Stolarzewicz, A. Investigation of infrared calibration methods for application to the study of
methyl methacrylate polymerization. Anal. Bioanal. Chem. 2001, 370, 899–903. [CrossRef]
28. Mas Haris, M.R.; Kathiresan, S.; Mohan, S. FT-IR and FT-Raman Spectra and Normal Coordinate Analysis of Poly methyl
methacrylate. Der. Pharma Chem. 2010, 2316–2323.
29. Bhattacharjee, S. DLS and zeta potential—What they are and what they are not? J. Control. Release 2016, 235, 337–351. [CrossRef]
30. Yamamoto, T.; Higashitani, K. Size control of polymeric particle in soap-free emulsion polymerization. KONA Powder Part J. 2018,
2018, 66–79. [CrossRef]
31. Li, X.; King, T.A. Microstructure and optical properties of PMMA/gel silica glass composites. J. Sol. Gel. Sci. Technol. 1995, 4,
75–82. [CrossRef]
32. Wheeler, O.H. Near Infrared Spectra Of Organic Compounds. Chem. Rev. 1959, 59, 629–666. [CrossRef]
33. Siesler, H.W.; Ozaki, Y.; Kawata, S.; Heise, H.M. Near-Infrared Spectroscopy: Principles, Instruments, Applications, 1st ed.; Wiley-VCH:
Weinheim, Germany, 2002; pp. 11–39.
34. Escobedo Morales, A.; Sánchez-Mora, E.; Pal, U. Use of diffuse reflectance spectroscopy for optical characterization of un-
supported nanostructures. Rev. Mex. Física 2007, 53, 8–22.
35. Hazim, A.; Abduljalil, H.M.; Hashim, A. First Principles Calculations of Electronic, Structural and Optical Properties of (PMMA–
ZrO2–Au) and (PMMA–Al2O3–Au) Nanocomposites for Optoelectronics Applications. Trans. Electr. Electron. Mater. 2021, 22,
185–203. [CrossRef]
36. Gupta, A.K.; Bafna, M.; Vijay, Y.K. Study of optical properties of potassium permanganate (KMnO4 ) doped poly(methylmethacrylate)
(PMMA) composite films. Bull. Mater. Sci. 2018, 41, 1–7. [CrossRef]
37. Gesesse, G.D.; Gomis-Berenguer, A.; Barthe, M.F.; Ania, C.O. On the analysis of diffuse reflectance measurements to estimate the
optical properties of amorphous porous carbons and semiconductor/carbon catalysts. J. Photochem. Photobiol. A Chem. 2020, 398,
112622. [CrossRef]
38. Bhat, V.S.; Kapatkar, S.B.; Ayachit, N.H.; Naik, I.; Murari, M.S. Doping-induced modulation of optical properties of PFO/PMMA
composite films. Polym. Bull. 2020, 0123456789. [CrossRef]
39. Wang, Z.L.; Chan, C.T.; Zhang, W.Y.; Chen, Z.; Ming, N.B. Optical properties of inverted opal photonic band gap crystals with
stacking disorder. Phys. Rev. E 2003, 67, 1–10. [CrossRef]
40. Versmold, H. Neutron diffraction from shear ordered colloidal dispersions. Phys. Rev. Lett. 1995, 75, 763–766. [CrossRef]
41. Verhaegh, N.A.M.; Van Duijneveldt, J.S.; Van Blaaderen, A.; Lekkerkerker, H.N.W. Direct observation of stacking disorder in a
colloidal crystal. J. Chem. Phys. 1995, 102, 1416–1421. [CrossRef]
42. Zhang:, R.; Qiang, Z.; Wang, M. Integration of Polymer Synthesis and Self-Assembly for Controlled Periodicity and Photonic
Properties, Adv. Funct. Mater. 2021, 31, 1–8. [CrossRef]

You might also like