Optimization and Modeling of CO Photoconversion Using A Response Surface Methodology With Porphyrin-Based Metal Organic Framework

Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

Reac Kinet Mech Cat

https://doi.org/10.1007/s11144-018-1407-z

Optimization and modeling of CO2 photoconversion


using a response surface methodology with
porphyrin-based metal organic framework

Nasrin Sadeghi1 • Shahram Sharifnia1 • Trong-On DO2

Received: 29 November 2017 / Accepted: 13 April 2018


Ó Akadémiai Kiadó, Budapest, Hungary 2018

Abstract In the current study, porphyrin-based metal organic framework


(M(metal)/PMOF) was used to photoreduction conversion of CO2 under UV/Visible
light irradiation. The metal and linker selection in the M/PMOF synthesis influences
its structure and properties. In this regards, M/PMOFs were prepared using three
different metals of Zn, Al, and Co. The prepared photocatalysts were characterized
by powder X-ray diffraction (XRD), N2 adsorption BET surface area, scanning
electron microscopy (SEM), Energy dispersive X-ray spectrometer (EDX), Fourier
transform infrared (FTIR) and UV–Vis spectroscopy. Also, the photoluminescent
properties and photoreactor tests, for these three photocatalysts were investigated.
The obtained results demonstrated that Al/PMOF has high CO2 photoreduction
conversion, 4.3%. The operating conditions were optimized to find the best con-
ditions which are lead to high CO2 photoreduction conversion over Al/PMOF. For
this purpose, the experiments were conducted based on central composite design
(CCD) and analyzed using response surface methodology (RSM). The ANOVA
analysis revealed that the maximum photoreduction conversion of CO2, as 10.63%,
can be obtained at the optimum conditions (catalyst amount of 297.24 mg, total feed
pressure of 1.4 atm and methanol as sacrificial agent). Finally, a verification
experiment was performed and results confirmed the validity and adequacy of the
predicted model for simulating the CO2 photocatalytic conversion.

Electronic supplementary material The online version of this article (https://doi.org/10.1007/s11144-


018-1407-z) contains supplementary material, which is available to authorized users.

& Shahram Sharifnia


[email protected]
1
Catalyst Research Center, Chemical Engineering Department, Razi University, Kermanshah,
Iran
2
Department of Chemical Engineering, Laval University, Quebec G1V0A8, Canada

123
Reac Kinet Mech Cat

Keywords Photocatalyst  Metal organic framework  Optimization  Response


surface methodology  CO2 photoreduction

Introduction

Carbon dioxide (CO2) as a primary greenhouse gas plays a significant role in global
warming. In this respect, there have been a variety of programs of policy and action
in relation to the environment and among them CO2 photoreduction process is a
popular technology used for reducing CO2 in last decades, however, few reports
dealing with the application of statistical design of experiments towards the
operational optimization of this process. The photocatalytic reduction of CO2 is a
complex process and involved complicated chain reactions, which generally require
a number of assumptions to solve the equations derived on the basis of physical and
concepts [1–3]. In the CO2 photoreduction process, possible variables that affect the
photocatalytic reduction efficiency could be initial CO2 pressure, catalyst amount
and/or characteristics, illumination time, light irradiance (power and/or intensity),
temperature and electron acceptors [4, 5]. The traditional ‘‘trial-and-error’’ and one-
factor-at-a-time (OFAT) optimization approaches can be applied to optimize CO2
photoreduction process [2, 4]. The major disadvantage of OFAT approach is that it
does not include interactive effects among the variables and, finally, it does not
describe the complete effects of the parameters on the process [6, 7]. Another
disadvantage is a large number of experiments needed which means require
additional time and also additional expense due to reagent costs [2, 8, 9]. Another
less frequently used approach is artificial neural networks (ANNs). The neural
networks behave as ‘black boxes’ and can be simply used to estimate various
parameters as a function of different variables [10–13]. The most significant
advantages of these modeling techniques are the computational efficiency, the
possibility to apply it on complex non-linear processes, the ease of manipulation
[2, 14, 15]. Nevertheless, the disadvantages of the ANNs approach are the selection
of regression variables, the necessity of obtaining a perfect neural network model
based on experimental or operational history data [2, 13, 14, 16]. In general, the
main aim of optimization is to improve the systems performance and to enhance the
yield of the processes while not increasing the cost [6, 17]. In this respect, the
number of experiments has a crucial role in operational and capital costs. More
experiments require more material, more analysis and spent more time [18].
In order to overcome these problems and optimize all of the parameters, the
statistical experimental design was applied by response surface methodology (RSM)
[19, 20]. RSM is an efficient mathematical model for the optimization of a
multivariable system with fewer experimental trials [21–23]. For instance, the
central composite design (CCD), an experimental design, was used by RSM to fit a
model by least squares technique [24, 25]. Adequacy of the suggested model is then
evaluated using the diagnostic checking tests provided by analysis of variance
(ANOVA). The response surface plots can be applied to consider the surfaces and
locate the optimum. In recent years, RSM is commonly used to evaluate the results

123
Reac Kinet Mech Cat

and operations efficiency even in many industrial processes [26–30]. In 2016,


Mosleh et al. [31] synthesized BiPO4/Bi2S3-HKUST-1-MOF and applied in a
catalytic rotating packed bed reactor for photocatalytic degradation of toluidine blue
(TB) and auramine-O (AO). CCD was applied to optimize operational parameters:
irradiation time, pH, photocatalyst dosage, rotational speed, solution flow rate,
aeration flow rate and TB and AO concentration. Totally 52 experiments were
performed to find the optimum values for each parameter. At optimum condition,
the photocatalytic degradation percentages of TB and AO were 99.37 and 98.44%,
respectively. In another study, Mosleh et al. [32] used Ag3PO4/Bi2S3-HKUST-1-
MOF to sonophotocatalytic degradation of trypan blue (TB) and vesuvine (VS). The
effect of operational parameters including the initial TB and VS concentration, flow
rate, irradiation and sonication time, pH and photocatalyst dosage were investigated
and optimized using CCD combined with desirability function (DF). Under the
optimized conditions, the hybrid system was found to have higher efficiency
compared with sum of the individual processes. According to our knowledge, no
prior work regarding optimizing the operational parameters to obtain high CO2
photoreduction conversion over MOF has been reported to date, which is important
for scale-up and commercialization purposes have been made.
This present study follows on from our previous work [1]. Previously we reported
Zn/PMOF to photoreduction of CO2 in the presence of the H2O vapor as an electron
donor under UV/Visible light. Forasmuch as the choice of metal and linker dictates
the structure and therefore properties of MOFs, it can be easy to construct MOFs
with desirable properties for specific application in CO2 reduction [33]. In this
respect, different metals are chosen to prepare M/PMOF [M: metal (Al and Co)] and
are used for photocatalytic reduction of CO2 under same reaction conditions of our
previous study [1]. The results are compared and the best photocatalyst with high
CO2 photoreduction conversion is chosen. Thereafter, the effects of several
operating parameters (amount of the catalyst, total pressure of feed and sacrificial
agent) are investigated. For this purpose, the central composite design (CCD) along
with RSM has been applied to the modeling and optimization of CO2 photoreduc-
tion conversion. Finally, the best operating condition which is lead to high CO2
photoreduction conversion is obtained.

Experimental

Materials

Pyrrole and 4-carboxybenzaldehyde were purchased from the Aldrich Chemical Co.
CH3Cl, acetone, methanol, Acetonitrile (MeCN), propionic acid, DMF, Co(NO3)2-
6H2O and Al(NO3)39H2O were supplied by Merck. All chemicals were used as
received without further purification. The CO2 gas with 99% purity was used as the
feed gas. The deionized water was used to prepare the catalyst.

123
Reac Kinet Mech Cat

Preparation of catalyst

Tetrakis (4-carboxy phenyl) porphyrin (TCPP) was synthesized according to the


method provided in our previous paper [1]. To prepare the catalyst, 0.39 g TCPP
was introduced into 0.13 L DMF and dissolved on a stirring hot plate. The reaction
was allowed to proceed for 20 min and then followed by the addition of 2.017 mmol
metal (0.75 g Al(NO3)39H2O and 0.6 g Co(NO3)26H2O). The solution was
allowed to react on the stirring hot plate for about 30 min and then transferred into a
250 mL Teflon-lined autoclave and heated at 180 °C for 24 h. After the solution
was cooled down to room temperature, the solid was centrifuged and washed 3
times with deionized water. Finally, the obtained powder was dried at 60 °C.

Photocatalyst characterization

Powder XRD patterns were recorded on a Bruker SMART APEXII X-ray


diffractometer equipped with a CuKa radiation source (k=1.5418 Å). N2 adsorp-
tion–desorption isotherms were obtained at 77 K using Quantachrome Autosorb-1
MP analyzer. Before the measurements, the samples were outgassed under vacuum
for 6 h at 150 °C. Scanning electron microscopic (SEM) images were captured
using a JOEL JEM 1230 operated at 120 kV. The elemental composition of the
catalyst was analyzed by energy dispersive X-ray spectrometer (EDX). The Fourier
transform infrared spectroscopy (FTIR) was recorded with an ALPHA (Bruker,
Germany) FTIR spectrophotometer. Samples of 1–2 mg were mixed with 100 mg
KBr and pressed into translucent disks at room temperature. All spectra were taken
in the range 400–4000 cm-1 at a resolution of 4 cm-1. UV–Vis spectra were
recorded on a Cary 100 Scan UV–Vis spectrophotometer (Varian, USA) at room
temperature (23–25 °C). The photoluminescence spectra of samples were per-
formed with a JASCO spectrophotometer Model FP-6200 equipped with a
thermostat bath, using a 1.0 cm quartz cell. The diffuse reflectance UV–Vis spectra
of the Al/PMOF and Co/PMOF powders were obtained by using Avantes
spectrometer (Avaspec-2048-TEC) with a reflectance sphere in the range,
300–900 nm.

Photoreactor system

The photoreactor system and the light source were selected similar to our previous
study [1]. The CO2 photoreduction was investigated in a batch reactor in the gaseous
phase. The photoreactor was made from stainless steel and built in the laboratory.
The effective volume of the reaction system was about 1 L. The photocatalyst
powders were ground to obtain fine particles and then distributed at the bottom of
the reactor. A 300 W high-pressure mercury (200 \ k \ 800) was placed in the
center of the reactor to direct irradiation on the catalyst surface. At the next step, the
reactor was filled with pure CO2 with 120 mL/min flow rate and then vacuumed
several times to ensure all the impurities and trapped air were completely removed
and then it was filled with high purity CO2 with 120 mL/min flow rate as feed gas to
1.4, 2.7, 4.1 atm (20, 40, 60 psia) of absolute total pressure. Different amount of

123
Reac Kinet Mech Cat

sacrificial agent (0.6, 2.96 and 4.52 mL for H2O, MeCN, and methanol) was
subsequently injected into the reactor as a reducer. Before starting the reaction, the
gas feed concentration was analyzed by using an on-line commercial gas
chromatograph GC-CGCA-1 apparatus. Then the reactor was illuminated for 4 h
continuously. During the reaction, the temperature was kept at 100 °C by a jacket
around the reactor equipped with a circulating cold water bath. After 4 h irradiation,
the products were analyzed by using on-line GC. Decreasing (conversion (%)) of
CO2 was used for evaluation the efficiency of photocatalytic activity (Eq. (1)):
 
Mole of COi2  Mole of CO2
Conversion of CO2 ð%Þ ¼  100 ð1Þ
Mole of COi2

where COi2 and CO2 are the concentration of CO2 at the beginning and the end of
the reaction time, respectively.

Experimental design and optimization by RSM

In this study, central composite design (CCD), which is widely used form of RSM,
was employed to optimize operating parameters that maximize the CO2 photore-
duction conversion. For this purpose, Design Expert software was applied. The
RSM fundamental assumptions and more detailed information have been described
elsewhere [4, 9]. Forasmuch as there has not been a great deal of study done on the
pressure effects on the MOF’s behavior [34], the effect of the pressure on the CO2
photoreduction process is studied. The studied photoreactor was tested to operate at
6 atm, higher pressure is possible, however, it was not tested and safe. So,
optimization was done at 1.4, 2.7 and 4.1 atm, which are available and safe. On the
other hand, electron donor can effect the products yield. Studies [35, 36] were
carried out to investigate the effects of the electron donors on the CO2
photoreduction over different MOFs. With these in mind, different electron donors
(methanol where is accessible, H2O that is safe, non-toxic and extensively available
and MeCN that is commonly used in CO2 photoreduction over MOFs) were selected
as effective parameters. On the other hand, catalyst amount is one of the main
parameters that affected the CO2 reduction but from the economic viewpoint
applying more catalyst is not affordable. Consequently, 100, 200, 300 mg catalyst
was applied for optimization in order to save money. The experimental ranges and
levels of the independent variables for photoreduction of CO2 are given in Table 1.

Table 1 Experimental ranges


Variables Ranges and levels
and levels of the independent
variables -1 0 ?1

Amount of catalyst, mg (X1) 100 200 300


Total pressure of feed, atm (X2)* 1.4 2.7 4.1
*
Total pressure of feed (X2): 20, Sacrificial agent (X3) H2 O MeCN Methanol
40 and 60 psia

123
Reac Kinet Mech Cat

Totally 39 experiments with three main factors were performed. The complete
experimental design matrix and the responses based on experiments proposed by
CCD for photoreduction of CO2 are given in Table 2. The following second-order
(quadratic) polynomial response equation was used to correlate the dependent and
independent variables:
X XX X
Yð%Þ ¼ b0 þ bi x i þ bij xi xj þ bii x2i ð2Þ
i i j i

Where Y is the response variable of conversion efficiency, b0 is the constant


coefficient, bi ;bii and bij present the regression coefficients for linear, quadratic
effects and the coefficients of the interaction parameters, respectively, and xi are the
independent variables studied.
Analysis of variance (ANOVA) for final predictive equation was used to obtain
the interaction between the process variables and the responses. The quality of the
fit second-order polynomial equation was expressed by the correlation coefficient
(R2) and its statistical significance was determined by the Fisher’s F-test in the same
program. The model’s terms were evaluated according to p-value (probability).
Three-dimensional plots and their respective contour plots were obtained for CO2
photoreduction conversion based on the effect of the three independent variables
(catalyst amount, pressure and sacrificial agent) at three levels. Moreover, the
optimum region was recognized based on the main parameters in the desirability
plot.

Results and discussion

Catalyst characterization

XRD analysis was used to investigate the crystal structure of the as-prepared
samples. As shown in Fig. 1a, all diffraction peaks for Al/PMOF are indexed to an
orthorhombic structure with Cmmm symmetry. The cell parameters have been
known with a = 31.978 Å, b = 6.5812 Å and c = 16.862 Å. As a result, the main
peaks obtained at 2h = 7.43° and 13.67° could be indexed as the (201) and (110)
reflections, respectively. In Al/PMOF, each oxygen in the carboxylate group is
coordinated with one Al3?, the aluminum coordination consists of four carboxylate-
derived oxygen atoms in the equatorial plane and two l2 axial OH- bridging
adjacent Al3? centers to form an infinite Al(OH)O4 chain which is a common motif
for M3? frameworks with carboxylate ligands (Fig. 1b) [37, 38]. According to all
diffraction peaks for Al/PMOF and Co/PMOF no impurity peaks are observed,
Fig. 1a. The relatively strong diffraction intensity demonstrates good crystallinity of
both MOF structures.
Fig. 2 demonstrates the nitrogen adsorption–desorption isotherms measured at
liquid nitrogen temperature. Both Al/POMF and Co/PMOF exhibit a typical II N2
adsorption isotherm which is observed in nanoporous or macropores materials or
open voids [38]. The BET surface areas were 1142 and 957 cm2/g for Al/PMOF and

123
Reac Kinet Mech Cat

Table 2 Experimental design matrix and experimental results with predicted values
Run Point type Independent variables Photoconversion (%)

Catalyst amount (mg) Pressure (atm) Sacrificial agent Observed Predicted

1 Axial 200 1.4 MeCN 9.70 10.10


2 Axial 200 1.4 Methanol 10.50 10.15
3 Center 200 2.7 MeCN 6.30 6.17
4 Fact 300 4.1 MeCN 1.20 1.40
5 Axial 300 2.7 MeCN 6.30 5.97
6 Axial 200 4.1 MeCN 2.05 2.23
7 Fact 100 1.4 MeCN 8.60 8.59
8 Fact 100 4.1 Methanol 3.90 2.71
9 Center 200 2.7 H2O 5.20 4.93
10 Fact 100 1.4 Methanol 9.40 8.64
11 Axial 200 4.1 Methanol 3.00 2.86
12 Center 200 2.7 MeCN 6.80 6.17
13 Center 200 2.7 H2O 6.10 4.93
14 Axial 100 2.7 Methanol 5.10 5.67
15 Fact 100 4.1 H2O 5.09 5.44
16 Fact 300 4.1 H2O 4.10 4.71
17 Axial 100 2.7 H2O 3.97 4.11
18 Fact 300 1.4 H2O 4.30 4.79
19 Axial 100 2.7 MeCN 4.64 5.34
20 Center 200 2.7 Methanol 6.37 6.50
21 Center 200 2.7 H2O 4.60 4.93
22 Center 200 2.7 Methanol 5.30 6.50
23 Fact 300 1.4 MeCN 10.35 10.58
24 Axial 300 2.7 H2O 5.40 4.74
25 Center 200 2.7 MeCN 7.27 6.16
26 Fact 100 1.4 H2O 2.88 2.78
27 Center 200 2.7 Methanol 6.20 6.50
28 Center 200 2.7 H2O 5.04 4.93
29 Fact 100 4.1 MeCN 1.80 2.08
30 Center 200 2.7 Methanol 5.90 6.50
31 Axial 200 1.4 H2O 3.80 4.28
32 Axial 200 4.1 H2O 5.70 5.59
33 Fact 300 4.1 Methanol 2.17 1.99
34 Fact 300 1.4 Methanol 11.02 10.62
35 Center 200 2.7 H2O 4.92 4.93
36 Center 200 2.7 Methanol 6.40 6.50
37 Center 200 2.7 MeCN 6.00 6.16
38 Axial 300 2.7 Methanol 6.20 6.31
39 Center 200 2.7 MeCN 6.03 6.16

123
Reac Kinet Mech Cat

(a)

Intensity Co/PMOF

Al/PMOF

5 10 15 20 25

(b)

Fig. 1 a XRD patterns of Al/PMOF and Co/PMOF, b structure of Al/PMOF. Atoms are colored as
follows: dark grey Al, light grey C, dark black N, and light black O

Co/POMF, respectively. The pore volume is 0.606 and 0.348 cm3/g for Al/PMOF
and Co/PMOF, respectively.
The morphology of the Al/PMOF and Co/PMOF were examined by SEM. As can
be seen in Fig. S1 (supplementary material), the both Al/PMOF and Co/PMOF are
composed by nanoplates. EDX analyses were done and confirmed the presence of
the Al in Al/PMOF and Co in Co/PMOF structures. In Fig. S2 (supplementary
material), the EDX patterns show a clear signal of the Al and Co which indicate Al
and Co were successfully incorporated into the Al/PMOF and Co/PMOF
frameworks, respectively.

123
Reac Kinet Mech Cat

600

500
Volume adsorbed (cm3/gcat.)

400 Co/PMOF

300

200

Al/PMOF
100

0
0.0 0.2 0.4 0.6 0.8 1.0
(P/Po)

Fig. 2 N2 adsorption–desorption isotherms of Al/PMOF and Co/PMOF

The FTIR spectra of TCPP, Co/POMF, and Al/PMOF were shown in Fig. S3
(supplementary material). The typical symmetric and asymmetric stretching bands
in the range of the 700–1700 cm-1 are assigned to the pyrrole ring of m (N–H), (C–
H), (C=C) and (C=N) of TCPP ligand. The band appeared about 3400 cm-1 in all
three structures is related to the OH vibration of adsorbed water. The FTIR spectra
show bands at 1715 and 1179 cm-1 which are due to m (C=O) stretch and m (C–O)
stretch of the carboxylic acid groups, respectively. The band appearing at
1500 cm-1 is assigned to the m (C=C) stretching vibration in pyrrole ring. It is
noticeable that the N–H vibration at 962 cm-1 is disappeared and the new peak at
1008 and 1002 cm-1 is appeared in the FTIR spectra of Co/POMF and Al/PMOF,
respectively. It is due to the replacement of hydrogen by metal cation [1, 39–42].
The UV–Vis spectra of TCPP, Co/PMOF, and Al/PMOF are demonstrated in the
Fig. 3. A typical strong Soret band with kmax at 418, 424 and 420 nm is shown for
TCPP, Co/PMOF, and Al/PMOF, respectively. Due to the metal (Co, Al) insertion
in the TCPP structure, a slight shift in the soret band is observed. Furthermore, two
weak split Q bands with kmax at 515 and 553 nm can see for TCPP, respectively.
Similar results were reported in the previous studies [37, 39]. The diffuse reflectance
UV–Vis spectra of the Al/PMOF and Co/PMOF powders were measured to
calculate the band gap energy. Band gap energy was estimated through Tauc
relation:
 
ðahmÞn ¼ B hm  Eg ð3Þ
where ht is the photon energy, a is the absorption coefficient, B is a proportionality

123
Reac Kinet Mech Cat

TCPP
Al/PMOF
Co/PMOF

Intensity (a.u.)

300 400 500 600 700


Wavelength (nm)

Fig. 3 UV–Vis spectra of TCPP, Al/PMOF and Co/PMOF in DMF solution

constant, Eg is the band gap energy, and the exponent n is a value that depends on
the nature of the transition (2 for a direct allowed transition, 2/3 for direct forbidden
transition and 1/2 for indirect allowed transition). MOFs display direct band tran-
sition, therefore n=2 [43, 44]. From Tauc plots, Fig. S4 (a) and (b) (supplementary
material), the band gaps of Co/PMOF and Al/PMOF are estimated to be 1.92 and
1.89 eV, respectively.

Photoreactor tests

One of the purposes of the current study was to determine the best M/PMOF with
desirable properties in CO2 photoreduction process. In this respect, the photolu-
minescent properties and photoreactor tests of the three photocatalysts, Zn/PMOF
[1], Al/PMOF and Co/PMOF, were investigated.
The photoluminescent properties of Zn/PMOF, Al/PMOF, and Co/PMOF were
investigated with an excitation wavelength 420 nm. The emission peaks of
photocatalysts are illustrated in Fig. 4. Zn/PMOF exhibits an intense emission
between 600 and 730 nm with kmax at 647 nm. The photoluminescent of Co/PMOF
was slightly quenched. Also, the emission intensity of Al/PMOF was quenched by
ca. 70%, indicating that lifetime of photogenerated charge carriers in Al/PMOF is
longer than those of Zn/PMOF and Co/PMOF. In other words, it implies that the
photogenerated charge carries can be separated efficiently due to the TCPP excited
electron transferring to the conduction band of metal, the mechanism of CO2
photoreduction over M/PMOF discussed in our previous work [1]. Also, this
behavior can be validated by photoreactor tests illustrated in Fig. 5.
It should be noted that as the different sacrificial agents, such as H2O, methanol
and MeCN, are utilized for process optimization (discussed in the following section)
and potentially can produce different product(s). For instance, for sacrificial water,
it seems methane is the main product and also CO and CH3OH can produce
[1, 45–49]. As another example, for sacrificial methanol, HCOOH and HCHO are
expected to be the main products of CO2 reduction. It is also plausible that by
proceeding the reaction over the time, some of these carbonyl compounds

123
Reac Kinet Mech Cat

Co/PMOF
Al/PMOF
Zn/PMOF
Intensity (a.u.)

550 600 650 700 750 800


Wavelenght (nm)

Fig. 4 Photoluminescence spectra of Zn/PMOF, Al/PMOF and Co/PMOF in DMF solution

mineralized into CO2 and H2O; and also CO2 may undergo methanation by the
excited electrons [45, 50, 51]. Consequently, due to the wide variety of the
producible products, the CO2 conversion was selected as an evaluation measure in
the current study.
Fig. 5 demonstrated the photoconversion of CO2 under UV/Visible light in the
presence of water vapor as electron donor after 4 h. The reaction condition is similar
to our previous study [1], 0.3 g of photocatalyst was loaded into the reactor and the
pressure of the CO2 was 1.4 atm. As it can be seen, the CO2 photoconversion
obtained by Al/PMOF is 4.3%. Using Zn/PMOF and Co/PMOF as a photocatalyst
under similar reaction conditions gave 1.3 and 1.8%, respectively. It can be
concluded that the conversion yield obtained by Al/PMOF is about 3.3 and 2.4 times
greater than the Zn/PMOF and Co/PMOF, respectively. Overall, the results of these
investigations show that Al/PMOF is the best photocatalyst with high CO2
photoreduction conversion.

5
CO2 Photoconversion

4.3
4

2 1.8
1.3
1

0
Zn/PMOF Co/PMOF Al/PMOF

Fig. 5 Comparison of the CO2 photoconversion over Zn/PMOF [1], Al/PMOF and Co/PMOF

123
Reac Kinet Mech Cat

In the following, the operating parameters (amount of the catalyst, total pressure
of feed and sacrificial agent) are optimized to obtain high CO2 photoreduction
conversion over Al/PMOF.

Model fitting and statistical analysis

After fitting the obtained data for each parameter of the full second-order model
(model containing all two parameter interactions), ANOVA for the model was
carried out and the full second-order model significance was examined. Forasmuch
as the full second-order model was not fulfilled the determined conditions, it was
improved by elimination of model terms.
While using the obtained results, an empirical relationship between the CO2
photoconversion, response, and the process variables, independent variables, was
achieved and expressed by the following second-order polynomial equations for
each sacrificial agent:
For H2O:
% Photoconversion
¼ 1:811 þ 0:037273  catal:amount þ 1:48395
ð4Þ
 pressure  5:00617  103  catal:amount  pressure
 5:09048  105  catal:amount2

For MeCN:
% Photoconversion
¼ þ 8:75497 þ 0:037273  catal:amount  1:91235
ð5Þ
 pressure  5:00617  103  catal:amount  pressure
 5:09048  105  catal:amount2

For methanol:
%Photoconversion
¼ þ 8:50083 þ 0:037273  catal:amount  1:6963
ð6Þ
 pressure  5:00617  103  catal:amount  pressure
 5:09048  105  catal:amount2

From the ANOVA of the reduced quadratic model (Table 3), the F value for the
model is 74.42, which implies that the model is significant. There is only a 0.01%
chance that the ‘‘model F value’’ could occur due to noise. Furthermore, the model
p-value is \ 0.0001, indicating that the model is significant. In this study, the
independent variables, including the pressure (X2), sacrificial agent (X3) and
interactions between pressure and sacrificial agent (X2X3) are highly significant
parameters with p \ 0.0001. Moreover, the catalyst amount (X1), interactions
between catalyst amount and pressure (X1X2) and the second-order effect of catalyst
amount (X12 ) are significant at p \ 0.05. On the other hand, lack of fit of the model is
not significant.

123
Reac Kinet Mech Cat

Table 3 ANOVA results for the response surface reduced quadratic model
Source Sum of Freedom Mean F- p-value
squares degree square value

Model 202.39 8 25.30 74.42 \ 0.0001 Significant


X1 -catalyst 1.78 1 1.78 5.23 0.0295
amount
X2-pressure 95.80 1 95.80 281.80 \ 0.0001
X3-sacrificial 17.63 2 8.82 25.94 \ 0.0001
agent
X1 X2 5.50 1 5.50 16.18 0.0004
X2 X3 79.17 2 39.58 116.44 \ 0.0001
X12 2.51 1 2.51 7.39 0.0108
Residual 10.20 30 0.34
Lack of fit 6.91 18 0.38 1.40 0.2798 Not
significant
Pure error 3.29 12 0.27

The important part of the data analysis procedure is the model validity checking.
The residual plots were examined by the proposed model. Residuals are the
difference between experimental values and the calculated values for each point by
the model. Residuals demonstrate how well the model satisfies the assumptions of
the ANOVA by detecting some outliers among total data [9, 52, 53]. The normal
probability versus studentized residuals plot for CO2 photoconversion is shown in
Fig. S5a (supplementary material). The results in Fig. S5a (supplementary material)
indicate the points fall near a straight line which means the residuals follow a
normal distribution. Furthermore, the S-shaped curve was not formed in Fig. S5a
(supplementary material), indicating that no response transformation was required
and also there was not an apparent problem with normality. The plot of studentized
residuals versus predicted responses is shown in Fig. S5b (supplementary material).
All points of experimental runs are dispersed randomly above and below the
horizontal lines between ± 3, which confirm the adequacy of the proposed model
and also the constant variance assumption was verified [54]. The ‘‘pred R-squared’’’’
of 0.9141 is in reasonable agreement with the ‘‘adj R-squared’’ of 0.9392 that
confirm good predictability of the model. The accuracy of the model is
demonstrated in Fig. S5c (supplementary material), which compares the actual
(measured) values against the predicted responses of the model for the CO2
photoreduction conversion. As observed in Fig. S5c (supplementary material), there
are tendencies in the linear regression fit, and the model could be used in the
experimental range of studied adequately. The fitted regression equation presented a
good fit of the model. Also, the experimental results and the predicted values
obtained from the models (Eqs. (4–6)) are presented in Table 2. These results
demonstrated good agreements between the experimental and predicted values of
CO2 photoconversion with R2 = 0.9520. The correlation coefficient (R2) implies
that 95.20% of the variations for percent CO2 photoconversion are explained by the

123
Reac Kinet Mech Cat

independent variables, in other words, this means that the model does not explain
only 4.80% of the variation.

Process optimization and model verification

In this study three-dimensional (3D) and contour (2D) plots were utilized to
demonstrate the effect of the one factor on the response value at different levels of
other factors. Notably, the 3D plots were obtained based on the proposed model and
widely used to realize the interactions between variables within the range
considered [54, 55].
Since there was 1 categorical factor (X3) in the experimental design, 3 plots can
be determined for each level (H2O, MeCN, and methanol). Effects of the interaction
between pressure and catalyst amount on the photoreduction conversion of CO2 are
shown in Fig. 6, also, Fig. S6 (supplementary material).
Fig. 6a shows the effects of the pressure and catalyst amount on the CO2
photoconversion for H2O as a sacrificial agent. As can be seen from Fig. 6a, the
CO2 photoconversion percentage increases with increasing pressure. On the other
hand, the CO2 photoconversion decreases with increasing the catalyst amount at the
high pressure, 4.1 atm. The influence of pressure on the photoconversion of CO2 is
significant for MeCN as shown in Fig. 6b. Surprisingly, the photoconversion
percentage of CO2 decreases drastically as the pressure increases. It is difficult to
explain this result, but it might be related to the deactivation of the photocatalyst
surface. There is the tendency toward the accumulation of reaction products on the
photocatalyst surface which also resulted in the photocatalyst deactivation after 4 h
as reported and discussed in previous studies [1, 2, 5, 56, 57]. Also, this result may
be explained by the Le Chatelier&s principle. According to this principle, with
increasing the system pressure, the reaction favors in the reverse order, and
conversion of products to reactants [2, 58, 59]. At low pressure, the CO2
photoconversion increasing proportionally to the catalyst amount, as expected,
confirming the positive influence of the catalyst amount on photocatalytic
performance. At high pressure, a slight decrease in CO2 photoconversion was
observed with increasing catalyst amount (from 200 to 300 mg). Similar results
were observed for methanol as shown in Fig. 6c.
The main purpose of the experimental design is the optimization of the process
which is lead to saving effort, time and process costs that are desirable from
industrial standpoint. The program uses several possible goals to conduct
desirability indices: maximize, minimize, target, within range, none (for response
only) and set to an exact value (factors only). The numerical optimization seeks to
maximize the desirability function. Simply put, desirability is an objective function,
which ranges from zero outside of the limits to one at the goal, to find the optimum
condition [53–55]. In this research, the goal was to reach maximum CO2
photoconversion. So, all three variables were set at ‘‘within range’’ and the
‘‘maximum value’’ was selected for the response, as shown in Table 4. The result of
the optimization is presented in Table 5. From Table 5, the result of the desirability
showed that the proposed model equations could adequately be used to describe the
CO2 photoconversion. According to the possible reasons discussed above and also

123
Reac Kinet Mech Cat

Conversion
B: Pressure

A: Catal. amount
(a)
Conversion
B: Pressure

A: Catal. amount
(b)
Conversion
B: Pressure

A: Catal. amount
(c)
Fig. 6 The response surface and contour plots of CO2 photoconversion (%) over Al/PMOF as the
function of pressure (atm) and catalyst amount (mg) for a H2O, bMeCN and c Methanol as a sacrificial
agent

shown in Table 5, the optimum pressure 1.4 atm was found to be having the highest
efficiency under the studied experimental conditions. Also, as shown in Table 5 the
best sacrificial agent is methanol. This result can be explained by the fact that the
pKa value of methanol, water and MeCN are 15.5, 15.7 and 25, respectively
[60, 61]. It is worthy to note that pKa values are a convenient way to compare acid
strengths, base strengths, Gibbs free energy changes, etc. The low pKa values point
to stronger acids while high pKa values indicate weak acids [62, 63]. So, it can be
concluded that methanol has the lowest pKa, so it is more acidity than the other
sacrificial agents (MeCN and water).
Besides this matter, for methanol and methoxide anion, the primary interaction (I,
Fig. 7) takes place between a filled lone-pair orbital (n) on oxygen and an

123
Reac Kinet Mech Cat

Table 4 Optimization of the individual responses (di) to obtain the overall desirability response (D)
Name Goal Lower Upper Lower Upper Importance
limit limit weight weight

Catal. amount In the 100 300 1 1 3


range
Pressure In the 1.4 4.1 1 1 3
range
Solvent In the H2O Methanol 1 1 3
range
Photoconversion Maximize 1.2 11.02 1 1 3

Table 5 Comparison between optimized CO2 photoconversion calculated from central composite design
and experimental study
Catalyst Pressure Sacrificial Photoconversion Desirability
amount (atm) agent (%)
(mg)

Optimized CO2 photoconversion 297.24 1.4 Methanol 10.63 0.96


calculated from CCD
(predicted value)
Confirmation study of optimized 297 1.4 Methanol 10.98
CO2 photoconversion
(experimental value)
Experimental valuePredicted value 3.18
Errorð%Þ ¼ Experimental value  100

unoccupied pMe orbital of the methyl group leading to a two-electron stabilizing


effect. Deprotonation of OH (giving rise to O-) results in an increase in the energy
of the n orbitals on oxygen and a declined energy separation between the interacting
orbitals n and pMe and thus a greater stabilizing interaction (II, Fig. 7). Therefore
methyl is more effective in stabilizing the OH- anion in comparison with the neutral
water molecule, and as a result, methanol is more acidic (in the gas phase) than
water [64].
In order to validate the optimum point generated by CCD, an experimental run
was performed at the optimum conditions. As shown in Table 5, the experimental
value (CO2 photoconversion: 10.98%) was found to be in good agreement with the
predicted value (CO2 photoconversion: 10.63%). Furthermore, two blank tests were
performed: (1) empty reactor (2) catalyst loaded in the reactor in the dark, under
same experimental condition. No Co2 conversion was observed in blank tests. For
case (1) no matter the light was on or off. These blank tests prove that the
conversion only occurred when catalyst present together with light irradiation. In
other words, it can be concluded that the reaction is photocatalytic. Also, The TiO2
was applied as photocatalyst at the optimum condition for photocatalytic reduction

123
Reac Kinet Mech Cat

Fig. 7 Energy diagram demonstrating the interaction of pMe and pMe orbitals of methyl with a lone-pair
orbital (n) of the OH and O- groups

of CO2. After 4 h illumination, the obtained CO2 photoconversion was 3.4% where
is about 3.3 times less than Al/PMOF. Based on the results, it can conclude that Al/
PMOF demonstrates the superiority photocatalytic activity toward CO2
photoreduction.

Conclusion

The photocatalytic reduction of CO2 over M/PMOF under UV/Visible light


irradiation was investigated and optimized in this study. In this respect, at first, three
different metals (Zn, Al and Co) were used to prepare M/PMOF. According to their
photoluminescent properties and photoreactor tests, it was found that Al/PMOF has
high CO2 photoreduction conversion among other two metals (Zn and Co). The
second aim of this study was to optimize the operating conditions which mainly
influence the photocatalytic reduction of CO2 as the crucial factor in this process
commercialization. Since different employed solvents were used for process
optimization and wide variety of different products may be produced, only on the
CO2 photoconversion was selected as an evaluation measure of the process.
Analysis of variance results for the reduced quadratic model demonstrated a high

123
Reac Kinet Mech Cat

coefficient determination (R2 = 0.9520 and Adj-R2 = 0.9392), thus ensuring a


satisfactory adjustment of the second-order regression model with the experimental
data. Optimizing experimental conditions showed that the maximum photoconver-
sion of CO2 was achieved in catalyst amount of 297.24 mg, total feed pressure of
1.4 atm and methanol as a sacrificial agent. Under optimized condition, the
experimentally measured value of CO2 photocatalytic reduction was in good
agreement with the predicted value implied that the optimization using RSM based
on central composite design can save time and effort by the estimation of the
optimum conditions of the maximum photoreduction conversion of CO2.

Acknowledgements This work was partly supported by the Natural Science and Engineering Research
Council of Canada (NSERC) through the Discovery Grant. The authors are deeply appreciated Professor
Trong-On DO for his helpful comments. A portion of this work was performed in his laboratory.

References
1. Sadeghi N, Sharifnia S, Sheikh Arabi M (2016) A porphyrin-based metal organic framework for high
rate photoreduction of CO2 to CH4 in gas phase. J CO2 Util 16:450–457. https://doi.org/10.1016/j.
jcou.2016.10.006
2. Torabi Merajin M, Sharifnia S, Mansouri AM (2014) Process modeling and optimization of simul-
taneous direct conversion of CO2 and CH4 greenhouse gas mixture over TiO2/webnet photocatalyst.
J Taiwan Inst Chem Eng 45(3):869–879. https://doi.org/10.1016/j.jtice.2013.09.013
3. Gharaie M, Zhang N, Jobson M, Smith R, Panjeshahi MH (2013) Simultaneous optimization of CO2
emissions reduction strategies for effective carbon control in the process industries. Chem Eng Res
Des 91(8):1483–1498. https://doi.org/10.1016/j.cherd.2013.06.006
4. Sakkas VA, Islam MA, Stalikas C, Albanis TA (2010) Photocatalytic degradation using design of
experiments: a review and example of the Congo red degradation. J Hazard Mater 175(1–3):33–44.
https://doi.org/10.1016/j.jhazmat.2009.10.050
5. Akhter P, Hussain M, Saracco G, Russo N (2015) Novel nanostructured-TiO2 materials for the
photocatalytic reduction of CO2 greenhouse gas to hydrocarbons and syngas. Fuel 149:55–65. https://
doi.org/10.1016/j.fuel.2014.09.079
6. Baş D, Boyacı İH (2007) Modeling and optimization I: usability of response surface methodology.
J Food Eng 78(3):836–845. https://doi.org/10.1016/j.jfoodeng.2005.11.024
7. Bezerra MA, Santelli RE, Oliveira EP, Villar LS, Escaleira LA (2008) Response surface method-
ology (RSM) as a tool for optimization in analytical chemistry. Talanta 76(5):965–977. https://doi.
org/10.1016/j.talanta.2008.05.019
8. Ansari F, Ghaedi M, Taghdiri M, Asfaram A (2016) Application of ZnO nanorods loaded on acti-
vated carbon for ultrasonic assisted dyes removal: experimental design and derivative spectropho-
tometry method. Ultrason Sonochem 33:197–209. https://doi.org/10.1016/j.ultsonch.2016.05.004
9. Amdoun R, Khelifi L, Khelifi-Slaoui M, Amroune S, Asch M, Assaf-Ducrocq C, Gontier E (2010)
Optimization of the culture medium composition to improve the production of hyoscyamine in
elicited Datura stramonium L. hairy roots using the response surface methodology (RSM). Int J Mol
Sci. https://doi.org/10.3390/ijms11114726
10. Desai KM, Vaidya BK, Singhal RS, Bhagwat SS (2005) Use of an artificial neural network in
modeling yeast biomass and yield of b-glucan. Process Biochem 40(5):1617–1626. https://doi.org/10.
1016/j.procbio.2004.06.015
11. Mjalli FS, Al-Asheh S, Alfadala HE (2007) Use of artificial neural network black-box modeling for
the prediction of wastewater treatment plants performance. J Environ Manage 83(3):329–338. https://
doi.org/10.1016/j.jenvman.2006.03.004
12. Dreiseitl S, Ohno-Machado L (2002) Logistic regression and artificial neural network classification
models: a methodology review. J Biomed Inform 35(5–6):352–359. https://doi.org/10.1016/S1532-
0464(03)00034-0

123
Reac Kinet Mech Cat

13. Tu JV (1996) Advantages and disadvantages of using artificial neural networks versus logistic
regression for predicting medical outcomes. J Clin Epidemiol 49(11):1225–1231. https://doi.org/10.
1016/S0895-4356(96)00002-9
14. Ciprian-George Piuleac CS, Cañizares Pablo, Curteanu Silvia, Rodrigo Manuel Andrés (2012)
Hybrid model of a wastewater-treatment electrolytic process. Int J Electrochem Sci 7(7):6289–6301
15. Pal MP, Vaidya BK, Desai KM, Joshi RM, Nene SN, Kulkarni BD (2009) Media optimization for
biosurfactant production by Rhodococcus erythropolis MTCC 2794: artificial intelligence versus a
statistical approach. J Ind Microbiol Biotechnol 36(5):747–756. https://doi.org/10.1007/s10295-009-
0547-6
16. Curteanu S, Piuleac CG, Godini K, Azaryan G (2011) Modeling of electrolysis process in wastewater
treatment using different types of neural networks. Chem Eng J 172(1):267–276. https://doi.org/10.
1016/j.cej.2011.05.104
17. Oliveira R, Almeida MF, Santos L, Madeira LM (2006) Experimental design of 2,4-dichlorophenol
oxidation by Fenton’s reaction. Int J Electrochem Sci 45(4):1266–1276. https://doi.org/10.1021/
ie0509544
18. Mosleh S, Rahimi MR, Ghaedi M, Dashtian K, Hajati S (2016) Photocatalytic degradation of binary
mixture of toxic dyes by HKUST-1 MOF and HKUST-1-SBA-15 in a rotating packed bed reactor
under blue LED illumination: central composite design optimization. RSC Adv 6(21):17204–17214.
https://doi.org/10.1039/c5ra24564h
19. Ahmadi M, Vahabzadeh F, Bonakdarpour B, Mofarrah E, Mehranian M (2005) Application of the
central composite design and response surface methodology to the advanced treatment of olive oil
processing wastewater using Fenton’s peroxidation. J Hazard Mater 123(1–3):187–195. https://doi.
org/10.1016/j.jhazmat.2005.03.042
20. Mahmodi G, Sharifnia S, Rahimpour F, Hosseini SN (2013) Photocatalytic conversion of CO2 and
CH4 using ZnO coated mesh: effect of operational parameters and optimization. Sol Energ Mat Sol
Cells 111:31–40. https://doi.org/10.1016/j.solmat.2012.12.017
21. Sasikumar E, Viruthagiri T (2008) Optimization of process conditions using response surface
methodology (RSM) for ethanol production from pretreated sugarcane Bagasse: kinetics and mod-
eling. Bio Energy Res 1(3):239–247. https://doi.org/10.1007/s12155-008-9018-6
22. Zhang H, Li Y, Lu Z, Wu M, Shi R, Chen L (2017) Highly efficient synthesis of biodiesel catalyzed
by CF3SO3H-functionalized ionic liquids: experimental design and study with response surface
methodology. React Kinet Mech Catal 121(2):579–592. https://doi.org/10.1007/s11144-017-1171-5
23. Baziar A, Ghashang M (2016) Preparation of pyrano[3,2-c]chromene-3-carbonitriles using ZnO
nano-particles: a comparison between the Box-Behnken experimental design and traditional opti-
mization methods. React Kinet Mech Catal 118(2):463–479. https://doi.org/10.1007/s11144-016-
1013-x
24. Montgomery DC, Runger GC, Hubele NF (2001) Engineering statistics. Wiley, New Jersey
25. Vining GG (2003) Statistical methods for engineers, 3rd edn. Cengage Learning, Boston
26. Parajó JC, Alonso JL, Lage MA, Vázquez D (1992) Empirical modeling of eucalyptus wood pro-
cessing. Bioprocess Eng 8(3):129–136. https://doi.org/10.1007/bf01254228
27. Wen Z, Liao W, Chen S (2005) Production of cellulase by Trichoderma reesei from dairy manure.
Bioresour Technol 96(4):491–499. https://doi.org/10.1016/j.biortech.2004.05.021
28. Beg QK, Sahai V, Gupta R (2003) Statistical media optimization and alkaline protease production
from Bacillus mojavensis in a bioreactor. Process Biochem 39(2):203–209. https://doi.org/10.1016/
S0032-9592(03)00064-5
29. Mayerhoff ZDVL, Roberto IC, Franco TT (2004) Purification of xylose reductase from Candida
mogii in aqueous two-phase systems. Biochem Eng J 18(3):217–223. https://doi.org/10.1016/j.bej.
2003.09.003
30. Muthukumar M, Sargunamani D, Selvakumar N, Venkata Rao J (2004) Optimisation of ozone
treatment for colour and COD removal of acid dye effluent using central composite design experi-
ment. Dyes Pigm 63(2):127–134. https://doi.org/10.1016/j.dyepig.2004.02.003
31. Mosleh S, Rahimi MR, Ghaedi M, Dashtian K, Hajati S (2016) BiPO4/Bi2S3-HKUST-1-MOF as a
novel blue light-driven photocatalyst for simultaneous degradation of toluidine blue and auramine-O
dyes in a new rotating packed bed reactor: optimization and comparison to a conventional reactor.
RSC Adv 6(68):63667–63680. https://doi.org/10.1039/c6ra10385e
32. Mosleh S, Rahimi MR, Ghaedi M, Dashtian K (2016) Sonophotocatalytic degradation of trypan blue
and vesuvine dyes in the presence of blue light active photocatalyst of Ag3PO4/Bi2S3-HKUST-1-

123
Reac Kinet Mech Cat

MOF: central composite optimization and synergistic effect study. Ultrason Sonochem 32:387–397.
https://doi.org/10.1016/j.ultsonch.2016.04.007
33. Wang C-C, Zhang Y-Q, Li J, Wang P (2015) Photocatalytic CO2 reduction in metal–organic
frameworks: a mini review. J Mol Struct 1083:127–136. https://doi.org/10.1016/j.molstruc.2014.11.
036
34. Ryder MR, Tan J-C (2014) Nanoporous metal organic framework materials for smart applications.
Mater Sci Technol 30(13a):1598–1612. https://doi.org/10.1179/1743284714Y.0000000550
35. Shi L, Wang T, Zhang H, Chang K, Ye J (2015) Electrostatic self-assembly of nanosized carbon
nitride nanosheet onto a zirconium metal-organic framework for enhanced photocatalytic CO2
reduction. Adv Funct Mater 25(33):5360–5367. https://doi.org/10.1002/adfm.201502253
36. Wang S, Lin J, Wang X (2014) Semiconductor–redox catalysis promoted by metal–organic frame-
works for CO2 reduction. Phys Chem Chem Phys 16(28):14656–14660. https://doi.org/10.1039/
c4cp02173h
37. Fateeva A, Chater PA, Ireland CP, Tahir AA, Khimyak YZ, Wiper PV, Darwent JR, Rosseinsky MJ
(2012) A water-stable porphyrin-based metal–organic framework active for visible-light photo-
catalysis. Angew Chem Int Ed 51(30):7440–7444. https://doi.org/10.1002/anie.201202471
38. Liu Y, Yang Y, Sun Q, Wang Z, Huang B, Dai Y, Qin X, Zhang X (2013) Chemical adsorption
enhanced CO2 capture and photoreduction over a copper porphyrin based metal organic framework.
ACS Appl Mater Interf 5(15):7654–7658. https://doi.org/10.1021/am4019675
39. Yuan Y-J, Tu J-R, Ye Z-J, Lu H-W, Ji Z-G, Hu B, Li Y-H, Cao D-P, Yu Z-T, Zou Z-G (2015)
Visible-light-driven hydrogen production from water in a noble-metal-free system catalyzed by zinc
porphyrin sensitized MoS2/ZnO. Dyes Pigm 123:285–292. https://doi.org/10.1016/j.dyepig.2015.08.
014
40. Yuan Y, Lu H, Ji Z, Zhong J, Ding M, Chen D, Li Y, Tu W, Cao D, Yu Z, Zou Z (2015) Enhanced
visible-light-induced hydrogen evolution from water in a noble-metal-free system catalyzed by
ZnTCPP-MoS2/TiO2 assembly. Chem Eng J 275:8–16. https://doi.org/10.1016/j.cej.2015.04.015
41. Krishna MBM, Venkatramaiah N, Venkatesan R, Narayana Rao D (2012) Synthesis and structural,
spectroscopic and nonlinear optical measurements of graphene oxide and its composites with metal
and metal free porphyrins. J Mater Chem 22(7):3059–3068. https://doi.org/10.1039/c1jm14822b
42. Seoudi R, El-Bahy GS, El Sayed ZA (2005) FTIR, TGA and DC electrical conductivity studies of
phthalocyanine and its complexes. J Mol Struc 753(1):119–126. https://doi.org/10.1016/j.molstruc.
2005.06.003
43. Jo SW, Kwak BS, Kim KM, Do JY, Park N-K, Ryu SO, Ryu H-J, Baek J-I, Kang M (2015)
Effectively CO2 photoreduction to CH4 by the synergistic effects of Ca and Ti on Ca-loaded
TiSiMCM-41 mesoporous photocatalytic systems. Appl Surf Sci 355:891–901. https://doi.org/10.
1016/j.apsusc.2015.07.176
44. Ngo TT (2016) Photocatalytic reduction of CO2 with tunable bandgap and bandedge materials.
University of South Florida, Tampa
45. Karamian E, Sharifnia S (2016) On the general mechanism of photocatalytic reduction of CO2. J CO2
Util 16:194–203. https://doi.org/10.1016/j.jcou.2016.07.004
46. Ji Y, Luo Y (2016) Theoretical study on the mechanism of photoreduction of CO2 to CH4 on the
anatase TiO2(101) surface. ACS Catal 6(3):2018–2025. https://doi.org/10.1021/acscatal.5b02694
47. Murcia-Lopez S, Vaiano V, Hidalgo MC, Navio JA, Sannino D (2015) Photocatalytic reduction of
CO2 over platinised Bi2WO6-based materials. Photochem Photobiol Sci 14(4):678–685. https://doi.
org/10.1039/c4pp00407h
48. Fresno F, Jana P, Renones P, Coronado JM, Serrano DP, de la Pena O’Shea VA (2017) CO2
reduction over NaNbO3 and NaTaO3 perovskite photocatalysts. Photochem Photobiol Sci
16(1):17–23. https://doi.org/10.1039/c6pp00235h
49. Zhou S-S, Liu S-Q (2017) Photocatalytic reduction of CO2 based on a CeO2 photocatalyst loaded
with imidazole fabricated N-doped graphene and Cu(ii) as cocatalysts. Photochem Photobiol Sci
16(10):1563–1569. https://doi.org/10.1039/c7pp00211d
50. Chen J, Xin F, Qin S, Yin X (2013) Photocatalytically reducing CO2 to methyl formate in methanol
over ZnS and Ni-doped ZnS photocatalysts. Chem Eng J 230:506–512. https://doi.org/10.1016/j.cej.
2013.06.119
51. Dey GR, Pushpa KK (2006) Methane generated during photocatalytic redox reaction of alcohols on
TiO2 suspension in aqueous solutions. Res Chem Intermediat 32(8):725–736. https://doi.org/10.1163/
156856706778606462

123
Reac Kinet Mech Cat

52. Körbahti BK, Rauf MA (2008) Response surface methodology (RSM) analysis of photoinduced
decoloration of toludine blue. Chem Eng J 136(1):25–30. https://doi.org/10.1016/j.cej.2007.03.007
53. Shahrezaei F, Mansouri Y, Zinatizadeh AAL, Akhbari A (2012) Process modeling and kinetic
evaluation of petroleum refinery wastewater treatment in a photocatalytic reactor using TiO2
nanoparticles. Powder Technol 221:203–212. https://doi.org/10.1016/j.powtec.2012.01.003
54. Yaqubzadeh AR, Ahmadpour A, Bastami TR, Hataminia MR (2016) Low-cost preparation of silica
aerogel for optimized adsorptive removal of naphthalene from aqueous solution with central com-
posite design (CCD). J Non-Cryst Solid 447:307–314. https://doi.org/10.1016/j.jnoncrysol.2016.06.
022
55. Vaez M, Zarringhalam Moghaddam A, Alijani S (2012) Optimization and modeling of photocatalytic
degradation of Azo dye using a response surface methodology (RSM) based on the central composite
design with immobilized titania nanoparticles. Ind Eng Chem Res 51(11):4199–4207. https://doi.org/
10.1021/ie202809w
56. Sasirekha N, Basha SJS, Shanthi K (2006) Photocatalytic performance of Ru doped anatase mounted
on silica for reduction of carbon dioxide. Appl Catal B 62(1–2):169–180. https://doi.org/10.1016/j.
apcatb.2005.07.009
57. Ahmad Beigi A, Fatemi S, Salehi Z (2014) Synthesis of nanocomposite CdS/TiO2 and investigation
of its photocatalytic activity for CO2 reduction to CO and CH4 under visible light irradiation. J CO2
Util 7:23–29. https://doi.org/10.1016/j.jcou.2014.06.003
58. Joos P, Serrien G (1991) The principle of Braun—Le Châtelier at surfaces. J Colloid Interf Sci.
145(1):291–294. https://doi.org/10.1016/0021-9797(91)90123-P
59. Jenkins HDB (2008) Le Chatelier’s Principle. Chemical Thermodynamics at a Glance. Blackwell
Publishing Ltd, New Jersey, pp 160–163. https://doi.org/10.1002/9780470697733.ch49
60. Brown W, Foote C, Iverson B, Anslyn E (2008) Organic chemistry. Cengage Learning, Boston
61. van Gunsteren WF, Weiner PK, Wilkinson T, Wilkinson AJ (1997) Computer simulation of
biomolecular systems: theoretical and experimental applications. Springer, New York
62. Shields GC, Seybold PG (2013) Computational approaches for the prediction of pKa values. CRC
Press, Boca Raton
63. Menger FM (1980) Electronic interpretation of organic chemistry: a problems-oriented text. Springer,
New York
64. Pross A, Radom L (1978) Does a methyl substituent stabilize or destabilize anions? Am Chem Soc
100(21):6572–6575. https://doi.org/10.1021/ja00489a005

123

You might also like