Next Article in Journal
Biocatalytic Oxidations of Substrates through Soluble Methane Monooxygenase from Methylosinus sporium 5
Next Article in Special Issue
Thermal Stability and Potential Cycling Durability of Nitrogen-Doped Graphene Modified by Metal-Organic Framework for Oxygen Reduction Reactions
Previous Article in Journal
Mixed-Phase (2H and 1T) MoS2 Catalyst for a Highly Efficient and Stable Si Photocathode
Previous Article in Special Issue
Three Component Controls in Pillared Metal-Organic Frameworks for Catalytic Carbon Dioxide Fixation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Photoreduction of Carbon Dioxide to Methanol over Copper Based Zeolitic Imidazolate Framework-8: A New Generation Photocatalyst

1
Department of Fundamental and Applied Sciences, Universiti Teknologi PETRONAS, Seri Iskandar, Perak 32610, Malaysia
2
Chemical Engineering Department, Universiti Teknologi PETRONAS, Seri Iskandar, Perak 32610, Malaysia
3
Biomass Processing Lab., Centre for Biochemical and Biofuel Research, Chemical Engineering Department, Universiti Teknologi PETRONAS, Seri Iskandar, Perak 32610, Malaysia
*
Authors to whom correspondence should be addressed.
Catalysts 2018, 8(12), 581; https://doi.org/10.3390/catal8120581
Submission received: 18 October 2018 / Revised: 17 November 2018 / Accepted: 21 November 2018 / Published: 25 November 2018
(This article belongs to the Special Issue Catalysis by Metal-Organic Frameworks)

Abstract

:
The efficient reduction of CO2 into valuable products such as methanol, over metal-organic frameworks (MOFs) based catalyst, has received much attention. The photocatalytic reduction is considered the most economical method due to the utilization of solar energy. In this study, Copper (II)/Zeolitic Imidazolate Framework-8 (Cu/ZIF-8) catalysts were synthesized via a hydrothermal method for photocatalytic reduction of CO2 to methanol. The synthesized catalysts were characterized by X-ray Photoelectron Spectroscopy (XPS), Field Emission Scanning Electron Microscopy (FESEM) coupled with Energy Dispersive X-ray (EDX), Ultraviolet-visible (UV-vis) spectroscopy, and X-Ray Diffraction (XRD). The host ZIF-8, treated with 2 mmol copper prepared in 2M ammonium hydroxide solution showed the highest photocatalytic activity. The crystal structures of ZIF-8 and 2Cu/ZIF-8N2 catalysts were observed as cubic and orthorhombic, respectively and the XPS analysis confirmed the deposition of Cu (II) ions over ZIF-8 surface among all the prepared catalysts. The orthorhombic structure, nano-sized crystals, morphology and Cu loading of the 2Cu/ZIF-8N2 catalyst were the core factors to influence the photocatalytic activity. The yield of Methanol was found to be 35.82 µmol/L·g after 6 h of irradiations on 2Cu/ZIF-8N2 catalyst in the wavelength range between 530–580 nm. The copper-based ZIF-8 catalyst has proven as an alternative approach for the economical photocatalytic reduction of CO2 to CH3OH.

Graphical Abstract

1. Introduction

The major environmental concern in the 21st century is global warming and carbon dioxide (CO2) is its main contributor [1]. The efficient and economical reduction of CO2 has received great attention in the recent past. To reduce the CO2 concentration, several approaches have been proposed [2]. One of the most favorable and advantageous methods is the photoconversion of CO2 into valuable fuel with the assistance of sustainable energy source such as sunlight [3]. The conversion of CO2 into valuable fuels is one of the ways to address the global warming problem efficiently. For this purpose, researchers have emphasized on the utilization of CO2 as an alternative feedstock to photocatalytically reduce CO2 to methanol. The discovery of photocatalytic reduction of CO2 over various semiconductors, for example, TiO2 [4], nanoparticles or nanotubes [5], Pt-promoted polyaniline-TiO2 nanocomposites [6], graphene oxide-CuO nanocomposites [7] have been used under UV-vis illumination, however, they exhibited low photocatalytic activity and absorbed only ultraviolet range of light [8]. TiO2 is proved to be most suitable photocatalyst of a semiconductor due to its chemical and physical stability, low-cost, non-toxicity and convenience for use. The only drawback of it is that it can only be excited by ultraviolet light and only 4% of sunlight is ultraviolet light [9,10]. Therefore, the development of photocatalyst, particularly visible light response catalyst for the efficient utilization of solar energy is hence of great significance.
The new generation catalyst which has been widely investigated since 1990 known as porous metal-organic frameworks (MOFs) [11]. MOFs are self-assembled by the coordination of metal cations/clusters with organics linkers. The MOFs can be used in several research areas such as nonlinear optics, molecular recognition, sensors, catalysis, gas storage and separation process [12]. Furthermore, application of MOFs extends to dye’s degradation [13], water splitting [14], and photocatalytic CO2 reduction or H2 production [15]. The broad application of MOFs is mainly due to the limited micropore dispersion, high dispersion of metals, prompt mono-dispersion of nanoparticles, and high surface area. All these properties contribute to increase the catalytic activity and selectivity [16]. MOFs demonstrated ultraviolet light photoresponse with absorption peaks around 250 nm. Depending on the type of material used, the absorption bands can expend from 250 nm to 300 nm or even in (>400 nm) visible region without further stretch [17]. Zeolitic imidazolate framework-8 (ZIF-8) is a sub-class of MOF and it exhibits excellent thermochemical and structural stability towards a CO2 reduction in water [18]. ZIF-8 also shows CO2 adsorption capability in liquid medium towards methanol production [19]. Because of uniform pores, channels on the surface and high surface area, ZIF-8 have been used as a catalyst in photocatalytic reaction [20], Knoevenagel reaction [21], cycloadditions and condensation reactions [22]. The bandgap energy of ZIF-8 is very high (Eg = 4.9 eV) which exhibits adsorption only in the UV region [23]. Hence, there is a need to modify pure ZIF-8 to increase the reduction of CO2, such as Zn2GeO4/ZIF-8 nanorods [19]. However, to the best of our knowledge, very limited work reported on new catalyst formulations for methanol production from CO2 under visible light irradiation.
Copper metal is a good applicant for coupling with ZIF-8 due to its amphoteric nature, inexpensive and easy synthesis. Hence the compounds of Cu/ZIF-8 are expected to show better photocatalytic reaction towards methanol production. Incorporation of copper metal over ZIF-8 reduces the effective bandgap energy and shows its incidence in the visible region. It is reported that new nanomaterials can be developed to improve the catalytic properties of Cu2+ and chemical stability of ZIF-8 crystals [22,24]. The Cu (II) ion plays the role of a co-catalyst in the reduction of CO2 by facilitating the transfer of charge and restricts the recombination of electrons and holes [25,26]. It has been reported that to reduce CO2, Cu cations and Cu oxides play an active role as co-catalyst and provide as reductive locations to increase the selectively of CH3OH [27,28,29]. The comparison with the Zn2GeO4 nanorods and Zn2GeO4/ZIF-8 hybrid nanorods consisting of 25 wt.% ZIF-8, Zn2GeO4/ZIF-8 was found to be 3.8 times higher in terms of adsorption capacity for dissolved CO2 compared to Zn2GeO4 nanorods. Consequently, the photoreduction rate of CO2 into CH3OH was found to increase by 62% [19]. According to Zhang et al., the morphology and size of the ZIF-8 crystal can be modified and the CO2 adsorption capacity can be increased if the concentration of ammonia is varied [30]. The ZIF-8 crystals are generally synthesized via N,N-dimethylformamide (DMF)) [31] and methanol [32] organic solvents. These solvents are toxic to the environment and expensive. Therefore, in this study, a nontoxic, environmentally benign and inexpensive solvent such as water is utilized as an alternative solvent to synthesize ZIF-8 material.
To explore the potential of new catalyst formulations for methanol production, Cu/ZIF-8 catalysts were synthesized by hydrothermal method. Present work reports the application of Cu/ZIF-8 catalysts in the photocatalytic reactor for the reduction of CO2 (in continuous flow system) to methanol in the presence of water (in batch mode) under 280 W/m2 irradiations of visible light.

2. Results and Discussion

2.1. Effect of Solvent Concentration on Cu/ZIF-8

The phase crystal structures of ZIF-8 and 2Cu/ZIF-8 with different ammonium hydroxide concentrations (2M, 3M and 4M) were determined by using powder XRD are presented in Figure 1. The XRD patterns revealed that the loading of copper imposed no significant effect on the ZIF-8 structure. As there was no significant peak of copper appeared in XRD spectra, it might be due to the relatively lower amount of Cu. It may be concluded that ammonia concentration can affect the structure, size and morphology of the catalyst, these results are also in line with FESEM images discussed in the next section. The refraction peaks of synthesized ZIF-8 are comparable with the reported literature [33].
ZIF-8 shows cubic structure in the all synthesized composites at Zn-N bond angle with crystal size 32 nm. However, 2Cu/ZIF-8N3 exhibits tetragonal structure with the crystal size of 114 nm and other two catalyst shows an orthorhombic structure where crystal size of 2Cu/ZIF-8N4 and 2Cu/ZIF-8N2 was 200 nm and 50nm, respectively, at the highest intensity peak of Cu-N bond angle in each sample. In 2Cu/ZIF-8N3, the appearance of ZIF-8 is very low due to the presence of some extra peaks of impurity of copper dinitrate(V) sesquihydrate, because it exhibits no valid relationship with crystal structure. Additionally, with the change in ammonia concentration (4M to 2M) the reduction of crystal size (as tabulated in Table A1) was observed due to further deprotonating of the unreacted imidazole ligands in the pores that accelerates the crystallization process as illustrated in Figure 2(Ad). However at excess of ammonia concentration in composite additional ammonium group would reside within the pores lead to increase of particle size [34]. It confirms that the solvent concentration can affect the crystal size of the catalysts without harmless the host material in the composite.
The morphological effects of ZIF-8 and 2Cu/ZIF-8 materials synthesized with different NH4OH concentrations were determined by FESEM as shown in Figure 2(Aa–Ad). The morphology of ZIF-8 is a hexagonal cage. In addition, Figure 2(Ab–Ad) illustrates that the addition of copper over ZIF-8 shaped the rhombic dodecahedron structure, which is a typical ZIF-8 structure as reported [35]. However, the addition of Cu metal did not considerably change the structure of ZIF-8, it could be due to the low Cu loading. However, the aggregate size was significantly reduced as it is highly affected by the solvent concentration. When NH4OH concentrations were decreased from 4M to 2M, the aggregate particle size of the catalysts was significantly decreased from 200 to 50 nm, respectively. The FESEM image of 2Cu/ZIF-8N2 illustrates the finest morphology of the desired crystals for the photocatalytic activity. The reduction of particle size is also in line with XRD pattern as discussed earlier. In Figure 2B, FESEM dot mapping of 2Cu/ZIF-8N2 indicates the dispersion of Cu ions into the composite. Every component present in the sample structure is finely distributed by dots with a specific color. It can be seen from the Figure 2B, that Cu species are homogeneously distributed.
UV-vis optical absorption spectra of the materials recorded within the region from 200 to 800 nm are shown in Figure 3A. 2Cu/ZIF-8 based materials exhibit a systematic increment in the absorption intensity with a relative decrement in ammonium hydroxide concentration. The 2Cu/ZIF-8N2 shows extensive absorption peak in the visible region as compared to the other materials. In Figure 3(Ad), 2Cu/ZIF-8N2 appeared in the ultraviolet range with a sharp peak at 330 nm. The 2Cu/ZIF-8N2 shows extensive absorption peak in the visible region as compared to the other materials and the same observation were characterized by Chary et al. [36]. The origin of the broad absorption peak in the visible region is might be due to the interaction of 2M concentration of ammonium hydroxide solvent with Cu/ZIF-8. The absorption peak in 2Cu/ZIF-8N2 sample at 330 nm shows a strong interaction of N-Cu-N cluster formation, signifying the highly dispersed form of copper. Apart from that, two more absorption shoulders were observed in the current study. An absorption peak centered at 520 nm with energy band gap of 2.38 eV suggested the presence of Zn2+. Similarly, the absorption band around 710 nm with energy band gap of 1.75 eV was assigned to 2Eg → 2T2g transitions of Cu2+, indicating octahedral symmetry of Copper. Moreover, this absorption band was also attributed to the crystalline phase of copper and d-d transitions of copper (II) ion. The bandgap of the developed catalysts is used to explore the effect of ammonium hydroxide concentration using Tauc’s plots as depicted in Figure 3B. In Tauc’s plot, h is Planck’s constant, υ is frequency and α is absorption coefficient [37]. The decrease in NH4OH concentration from 4M to 2M resulted in the decrease of bandgap energy from 4.58 eV for 2Cu/ZIF-8N4, 3.50 eV for 2Cu/ZIF-8N3, 1.75 eV for 2Cu/ZIF-8N2. As shown in Figure 3B, the bandgap value decreased from 4.89 eV to 1.75 eV with the addition of copper into ZIF-8 that shows smaller bandgap rather than stated by Schejn et al. [22]. Hence, it proved that the UV region of host ZIF-8 (4.89 eV) can be affected with the addition of copper and different ratio of NH4OH solvent. It is concluded that 2Cu/ZIF-8N2 exhibits a sharp edge at 1.75 eV activating the visible region of the photocatalyst. The shift in a peak in Figure 3A depicts a strong interaction between the copper metal and ZIF-8 framework. ZIF-8 and 2Cu/ZIF-8N4 present a narrow absorption profile in the visible region compared to other two 2Cu/ZIF-8 materials. The bandgap of ZIF-8 reduced after inclusion of copper on it and shifted to lower bandgap energies with more absorbance and become active in the visible region. The results suggested that 2Cu/ZIF-8N2 can be photoexcited to create more electron−hole pairs under visible light absorbance which could lead to higher photocatalytic efficiency [28,38].
Conclusively, the significant effect of solvent concentration was observed in 2Cu/ZIF-8N2. It is concluded that by increasing the molar concentration of solvent in 2Cu/ZIF-8 catalysts significantly modified the crystal structure from cubic to orthorhombic. The effect of solvent also reduced the crystal size from 200 nm to 50 nm from 4M to 2M, respectively. To investigate the effect of Cu loading, further experiments were performed on Cu/ZIF-8N2 (2M NH4OH concentration) to study the physiochemical properties as well as the activity of the catalysts.

2.2. Effect of Copper Loading on ZIF-8

The phase crystal structures of Cu/ZIF-8N2 (2M NH4OH concentration) with different amount of copper loading (1, 2 and 3 mmol) were measured by using powder XRD as can be seen in Figure 4. The diffraction peaks of all Cu/ZIF-8N2 catalysts with different copper loadings are comparable with the reported literature for ZIF-8 [33]. Hence, the doping of copper did not impose any significant effect on the ZIF-8 structure. However, in the 3Cu/ZIF-8N2 sample, two different peaks appeared due to the increasing copper molar ratios at 35.79° and 38.93° corresponds to Cu2O [39] and CuO [40], respectively. While in other samples copper ion peaks were not observed due to low metal loading. ZIF-8 showed cubic structure as shown in Figure 1a and Figure 2(Aa) whereas, other catalysts (1Cu/ZIF-8N2, 2Cu/ZIF-8N2 and 3Cu/ZIF-8N2) showed the orthorhombic structures as illustrated in Figure 4 and Figure 5a–c. A cubic lattice structure can be converted into orthorhombic lattice upon stretching [41]. Therefore, the orthorhombic structure is visible after the deposition of copper onto the ZIF-8 surface. As no other crystal structures were observed, all Cu/ZIF-8N2 catalysts with different copper loadings showed an orthorhombic structure with ICDD code: 03-065-6209. Peak broadening indicates the formation of nanosized crystals as can be clearly seen in the catalysts XRD pattern (Figure 4).
The one-sided N-Cu-N angle of a crystal lattice was observed when the framework parameters were a = 13.46 Å, b = 3.08 Å, c = 9.08 Å. All samples have a crystal size below 90 nm (Table A1) as predicted from the Scherrer Equation (Equation (11)) for the highest intensity peak of each sample.
The topographic images of Cu/ZIF-8N2 with different amount of copper loading are presented in Figure 5a–c. FESEM images showed that, with the increase in copper loading on host materials, the surface morphology of the catalyst can be affected. However, particles are uniformly distributed and no significant effect on the particle aggregation was observed. FESEM-EDX of the 2Cu/ZIF-8N2 catalyst indicates that Cu, Zn, C and N are the four components observed on the surface of the sample as illustrated in Figure 5d. EDX and FTIR spectra of Cu/ZIF-8N2 are presented in Figures S1 and S2 respectively, which confirm the synthesis of Cu/ZIF-8 composite and presence of imidazole ring attached with metals atoms.
The optical absorbance spectra of all Cu/ZIF-8N2 catalysts with different amount of copper loading were examined by UV-vis spectroscopy as can be seen in Figure 6. An extensive absorption spectrum can be observed for each sample. It can be found that ZIF-8 has a low absorbance in the visible region. After copper was loaded and increased consistently, the absorbance of the all Cu/ZIF-8N2 materials developed more intense in the visible region (λ > 400 nm). It is evident that there is no ending of spectrum in the visible region of all catalysts. Zhang et al. [42] studied the increment in the visible region after deposition of the metal onto the support surface. The UV-vis spectra clearly indicate that the absorption of a metal particle onto the support surface enables it to participate in the photocatalytic reaction. The formation rate and the number of electrons and holes on the surface of the photocatalyst correlate significantly with the absorbance of visible light [43]. The band gap energies of the Cu/ZIF-8N2 catalysts are mentioned in Table A1.
Furthermore, the elemental analysis was performed by XPS to confirm the loading of Cu (II) ion onto the ZIF-8 surface for each Cu/ZIF-8N2 catalyst. The spectrum in Figure 7A shows that the Cu 2p3/2 and Cu 2p1/2 peaks are located at approximately 934 eV and 954 eV, respectively, with a shake-up peak at ~943 eV for all materials. The existence of these peaks confirms the presence of Cu2+ in the catalyst [44]. It is well reported that Cu may exist in the form of CuO [45,46]. Whereas, in the present study, it was observed that Cu is linked with a nitrogen atom of ZIF-8 which is also confirmed with FTIR spectra (Figure S2).
The peak fitted graph is combined with the Cu2p graphs in order to show the presence of Cu2+ and Cu+ ions as shown in Figure 7A. The spectra show that 2Cu/ZIF-8N2 possess the higher concentration of Cu2+ as compared to Cu+. Whereas, other two catalysts exhibit comparatively higher concentration of Cu+ which lead to lower photocatalytic activity. Another resolution spectra of Zn (II) ions is demonstrated in Figure 7B, represents two characteristics peaks at 1022.4 eV and 1045.3 eV correspond to Zn2p3/2 and Zn2p1/2, respectively. These binding energies confirmed the presence of Zn (II) ions attached with nitrogen in the imidazole ring [47].

2.3. Photocatalytic Activity

The photocatalytic reactions of all Cu/ZIF-8 catalysts were assessed by photocatalytic reduction of CO2 in sodium hydroxide (NaOH) and sodium sulfite (Na2SO3) solutions in the presence of visible light for 6 h. The photocatalytic activity performed in this work is in line with the literature [48]. The use of NaOH not only enhanced the CO2 reduction via OH radical formation but also increased the CO2 solubility. However, Na2SO3 increased the proton concentration within the semiconductor particle and hole scavenging for CO2 reduction, which increases electron’s decay time. Na2SO3 which is working as a sacrificial agent in the supporting electrolyte, oxidized to Na2SO4. The experimental results revealed in Figure 8, which exhibit that the ZIF-8 was unable to generate methanol in the absence of copper whereas, other catalysts produced methanol over 6 h of irradiation time. Cu (II)/zeolitic imidazolate framework-8 offered a reductive surface to CO2 for methanol production. 2Cu/ZIF-8N2 catalyst demonstrated the most significant activity with the yield of 35.82 µmol/L·g due to its good crystallinity, morphology and optimal chemical loadings. When the copper loading was increased to 3 mmol, the copper ions were reduced to Cu2O as shown in Figure 4c, which degraded the separation efficiency of electron–hole pairs, consequently, declines the activity of the catalyst. XRD results confirmed that Cu/ZIF-8N2 catalysts were nanosized crystals, which are significant for photocatalytic reduction. The catalyst with high crystallinity presents fewer defects in the reaction system. Furthermore, the defected material played a vital role in the rapid recombination of photogenerated holes and electrons [49], which is responsible to reduce the photocatalytic activity. The catalytic performance of CuO and ZIF-8 was carried out at initial screening separately but no significant desire product was observed in photocatalytic reaction. In addition, we tried pure P25-TiO2 in the reactor system with the same conditions and reaction solvents but there was no substantial liquid product was detected. As predictable, other hydrocarbons were formed during reaction, which were quite low to be detected.
However, in this work, methanol was the main product which is considered as value added chemicals while no other product was formed. To the best of our knowledge, this is the first experiment of methanol production in the presence of Cu/ZIF-8 catalyst.
In this study, the yield of methanol is compared with the literature data as tabulated in Table 1. Due to the different catalyst preparation method, reaction conditions and reactor setup, an exact comparison cannot be made, but a generic statement can be made that the production of methanol can be increased with the addition of metal.
A photocatalytic process requires a semiconductor which absorbs the light and leads to the creation of photoexcited charge carriers. The respective photogenerated charge carriers (electron and hole) must be located at a potential negative enough for the reduction reaction and positive enough for the oxidation half-reaction. The electrons (e) can reduce CO2 to CH3OH while holes can be oxidized by water molecules into the proton. Figure 9 presents the schematic diagram of the photocatalytic reduction of CO2 to CH3OH over the photocatalyst. In our investigation, when the Cu/ZIF-8 catalysts were exposed to visible light, the photogenerated electron–hole pairs were created.
Cu/ZIF-8 + hυ → Cu/ZIF-8*(CB)
Cu/ZIF-8*(CB) → h+ (ZIF-8) + e (Cu2+)*
The formation of proton (H+) from water were carried out by Na2SO3, and at that point it reduces to H*, and this H* may further reduce the CO2.
CO2 (adsorption) + e → CO2*
H2O + Na2SO3 (4h+)→ 2H+ + Na2SO4
CO2* + H+ + e → HCOO
HCOO + H+ → HCOOH (Formic acid)
2H2O (4h+) → O2 + 4H+
Production of formic acid during this reaction is not a single step process and it further reduce to formaldehyde (HCHO) and water (H2O) molecule with the help of electrons and protons. This reaction might be happened in the conduction band of Cu/ZIF-8 over metal atom.
HCOOH + 2H+ + 2e → HCHO + H2O
In the final step of the photocatalytic reduction of CO2, formaldehyde reduces to the most stable liquid product, which is methanol in this study. Later on, methanol desorbs from the catalyst surface and convert to the bulk phase of liquid product.
HCHO + 2H+ + 2e → CH3OH
Therefore, the overall reaction in the CB of Cu/ZIF-8 is
CO2 + 6H+ + 6e → CH3OH + H2O
The electrons in the conduction band (CB) of 2Cu/ZIF-8N2 (ECB = −1.02 V vs. NHE at pH 7) improved the separation of electron–hole pairs and prevents from recombination. The valence band (VB) potential of ZIF-8 is 2.41 V, which is more positive than the VB potential of 2Cu/ZIF-8N2 (0.72 V) thus the oxidation reaction could happen in the VB of ZIF-8. VB and CB values of all synthesized samples during this study tabulated in Table A1 which were calculated with the equation reported by Wang et al. [52].
On the basis of our findings, it could be postulated that the addition of metal to a material can alter the surface properties and overall efficiency of a material. The incorporation of metals as a charge-carrier trap enhance the reaction rate and also increase the lifespan of separated electron-hole pairs [53,54]. During the photocatalytic reaction, the electrons move from lower to higher energy level to make balance and create a Schottky barrier [55]. The Schottky barrier serves as an electron trap which stops the migration of electrons from crossing back to the valence band (VB) and therefore inhibits recombination. It was predicted that the electrons generated in the conduction bands (CB) of the Cu/ZIF-8 catalysts took place after the irradiation of visible light. This could separate the photogenerated electron–hole pairs. When pure ZIF-8 was used, no methanol was produced, the generation of methanol could be due to the reaction of CO2 with protons and electrons (ECO2/CH3OH = −0.38 V) as reported [56]. A semiconductor has two main distinctive characteristics: firstly, a charge-separated state is generated when energy photons are greater than the bandgap of the material, used for the irradiation process and secondly, the mobility of the charge carriers. The charge carriers move away from the location where the separation of the charge occurs [55].

3. Materials and Methods

Materials such as Zinc nitrate hexahydrate (Zn(NO3)2·6H2O), ammonia (25%), copper sulphate pentahydrate (CuSO4·5H2O), sodium sulfite (Na2SO3) and sodium hydroxide (NaOH) were purchased from R&M chemicals (Essex, UK) while 2-methylimidazole (Hmim) was obtained from Sigma Aldrich (St. Louis, MO, USA). Deionized (DI) water was used in the synthesis of ZIF-8 as a solvent. No further purification of all reagents was performed and used directly.

3.1. Preparation of ZIF-8

ZIF-8 crystals were synthesized by the reported method [33] with few modifications. In 10 mL of DI water, 1.17 g of Zn(NO3)2·6H2O was dissolved and 22.70 g of Hmim was added in 80 mL of DI water. The solution of zinc nitrate was added dropwise into Hmim solution with vigorous stirring for 15 min then the solution was kept aside for seven days at room temperature. The solution was separated by removing the upper liquid solution using a syringe and the remaining product was filtered. The obtained product was dried at 70 °C in an oven overnight after washing it for 4–5 times with DI water.

3.2. Synthesis of Catalysts

To synthesize Cu/ZIF-8 catalysts, the molar ratios of ZIF-8 and copper sulphate presented in Table 2 were used. The chemicals were mixed in DI water and then the solution was stirred continuously. Copper sulphate solution was mixed with the ammonium hydroxide (sacrificial agent) solutions before mixing with ZIF-8. The mixture was poured into Teflon-lined stainless-steel autoclave and then hydrothermally treated at 110 °C for two days. The solid product was obtained by filtration and washed with DI water for 3–4 times. The prepared catalyst was finally dried at 85 °C overnight. The method adopted for the synthesis of catalysts was taken from Li et al. [48].

3.3. Characterization

The Ultraviolet-visible (UV-vis) Spectroscopy and Diffuse Reflectance Spectra (DRS) (Agilent Cary 100 UV-Vis Spectrophotometer, Santa Clara, CA, USA) were used to determine the optical absorption and bandgap energy of catalysts, respectively. The UV-vis spectra of the sample were found to be in the range of 200–800 nm. Field Emission Scanning Electron Microscopy (FESEM) and Energy Dispersive X-ray (EDX) (JEOL6340, Akishima, Tokyo, Japan) were used to investigate the morphology and element composition, respectively. The surface compositional charges were studied by X-ray Photoelectron Spectroscopy (XPS, Kratos, Manchester, UK). The analysis was carried out on a K-Alpha instrument applying a monochromatic Al Kα source (hν = 1486.6 eV, 150 W). XRD was used to identify the crystal structures of the catalysts and lattice parameters, (Bruke/D8-Advance, Karlsruhe, German; Cu Kα radiation, λ = 0.154056 nm) at room temperature at a scanning rate of 2°/min ranging from 10° to 80°. The size of crystal was calculated using the Scherrer equation (Equation (11)).
D = Kλ/β Cosθ
where, D = Crystallite size, K is crystallite shape factor (0.9), λ = 1.54 nm, β is the full width at half maximum intensity of the peak (in radian), and Cosθ is the angle (position) of the peak at 2θ.

3.4. Photocatalytic Activity

The 500-Watt optical photoreactor (Shanghai sunny scientific collaboration, Shanghai, China) configuration consists of a three-neck flask with an inner vessel for inlet and outlet water valves shown in Figure 10. The vessel volume was 500 mL. One side of the glass vessel was used for gas bubbling, another side was used for gas evacuation, while the remaining side was employed as sample collecting dock. The reactor vessel was irradiated by a 500 W Xenon lamp placed at the top of the reactor and the wavelength range was observed by intensity meter (Figure A1). An aluminum halt reflector was used to cover the vessel to enhance the visible light illumination and to prevent the intervention of ambient light. The reactor vessel was placed on the magnetic stirrer and the reaction temperature was controlled by the chiller. To confirm the photoreduction of CO2 to methanol formation, blank tests were conducted. The blank tests involved the reactions with the catalyst in the dark and without the catalyst in the visible light.
Initially, 500 mL of NaOH and Na2SO3 solution was poured into the glass vessel and its pH value was measured. Pure CO2 was bubbled into the reactor chamber for 1 h to saturate the solution. 0.5 g of catalyst was added into the reactor vessel and the resultant solution was stirred at a specific speed. The blank sample was collected without irradiation. Later, the light was turned on and eight liquid samples (5 mL each) were collected from the reactor with an interval of 1 h using a syringe. To isolate the catalyst (in particle form) from liquid samples, the catalyst-suspended samples were centrifuged. The collected samples were analyzed using GC-FID (gas chromatography-flame ionization detector, Mundelein, IL, USA) of Agilent 7890 equipped with a blood alcohol column (Agilent J&W DB-ALC column, Santa Clara, CA, USA).

4. Conclusions

In this study, photocatalysts have been synthesized, characterized and tested for the reduction of CO2 to methanol. Cu/ZIF-8 catalysts with different copper loading and ammonium hydroxide concentrations were synthesized using the hydrothermal method. The photocatalysts have been successfully tested for the reduction of CO2 to methanol. The XRD spectra confirmed the synchronization of N-Cu-N angle with orthorhombic structure after the addition of copper metal into ZIF-8 which possessed cubic lattice structure. FESEM images also exhibited the conversion of cubic lattice to the orthorhombic structure. The UV-vis analysis indicated the increase of absorbance in the visible region after the deposition of copper and an optimized amount of solvent onto the support surface. All samples are able to produce methanol due to the presence of Cu2+ ions on the support surface except for ZIF-8 which is confirmed from XPS analysis. Therefore, it is confirmed that copper (II) ions play an essential role in reducing CO2 to CH3OH. The highest yield of methanol was found to be 35.82 µmol/L·g after 6 h of irradiation with 2 mmol Cu loading and 2M ammonium hydroxide concentration. The high yield of methanol is attributed to the synthesis method, solvent effect and Cu loading over ZIF-8. Therefore, it can be concluded that an average Cu loading, proper selection and concentration of solvent significantly affect the photocatalytic reduction of CO2 to methanol.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/8/12/581/s1, Figure S1. EDX images of (a) 1Cu/ZIF-8N2, (b) 2Cu/ZIF-8N2 and (c) 3Cu/ZIF-8N2; Figure S2. FTIR spectra of (a) 1Cu/ZIF-8N2, (b) 2Cu/ZIF-8N2 and (c) 3Cu/ZIF-8N2.

Author Contributions

Writing—original draft, S.G.; Writing—review & editing, M.S.S.; Formal analysis, C.F.K.; Supervision, B.A.; Visualization, M.A.

Funding

This research was funded by Ministry of Higher Education Malaysia (MOHE) for Funding Fundamental Research Grant (FRGS-015-3AA-E98) and “The APC was funded by Centre for Graduate Studies (CGS), UTP, Malaysia and FRGS-015-3AA-E98”.

Acknowledgments

Financial assistance from the Foundation of Universiti Teknologi PETRONAS, under the Research Project: Design and Synthesis of Cu/Zr Zeolite-like metal-organic frameworks Catalyst with Tuneable Pore Size for Enhanced Photoreduction of CO2 to Methanol is gratefully acknowledged.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Figure A1. Reactor setup attached with intensity meter.
Figure A1. Reactor setup attached with intensity meter.
Catalysts 08 00581 g0a1
Table A1. Summarized data of all catalyst synthesized in this study.
Table A1. Summarized data of all catalyst synthesized in this study.
SampleCrystal SizeCrystal StructureBand Gap Energy (eV)VB Potential (V)CB Potential (V)Methanol Yield (µmol/L·g)
ZIF-832 nmCubic4.892.41−2.48No yield
2Cu/ZIF-8N4200 nmOrthorhombic4.582.14−2.440.655
2Cu/ZIF-8N3114 nmTetragonal3.501.6−1.916.19
2Cu/ZIF-8N250 nmOrthorhombic1.750.72−1.0235.82
1Cu/ZIF-8N238 nmOrthorhombic1.800.75−1.052.6
3Cu/ZIF-8N282 nmOrthorhombic1.650.67−0.9727.38

References

  1. Ma, S.; Sadakiyo, M.; Heima, M.; Luo, R.; Haasch, R.T.; Gold, J.I.; Yamauchi, M.; Kenis, P.J. Electroreduction of carbon dioxide to hydrocarbons using bimetallic Cu-Pd catalysts with different mixing patterns. J. Am. Chem. Soc. 2016, 139, 47–50. [Google Scholar] [CrossRef] [PubMed]
  2. Inoue, T.; Fujishima, A.; Konishi, S.; Honda, K. Photoelectrocatalytic reduction of carbon dioxide in aqueous suspensions of semiconductor powders. Nature 1979, 277, 637–638. [Google Scholar] [CrossRef]
  3. Kimfung, L.; Martin, D.; Junwang, T. Conversion of solar energy to fuels by inorganic heterogeneous systems. Chin. J. Catal. 2011, 32, 879–890. [Google Scholar]
  4. Li, K.; An, X.; Park, K.H.; Khraisheh, M.; Tang, J. A critical review of CO2 photoconversion: Catalysts and reactors. Catal. Today 2014, 224, 3–12. [Google Scholar] [CrossRef]
  5. Xie, S.; Wang, Y.; Zhang, Q.; Fan, W.; Deng, W.; Wang, Y. Photocatalytic reduction of CO2 with H2O: Significant enhancement of the activity of Pt–TiO2 in CH4 formation by addition of MgO. Chem. Commun. 2013, 49, 2451–2453. [Google Scholar] [CrossRef] [PubMed]
  6. Liu, G.; Xie, S.; Zhang, Q.; Tian, Z.; Wang, Y. Carbon dioxide-enhanced photosynthesis of methane and hydrogen from carbon dioxide and water over Pt-promoted polyaniline–TiO2 nanocomposites. Chem. Commun. 2015, 51, 13654–13657. [Google Scholar] [CrossRef] [PubMed]
  7. Gusain, R.; Kumar, P.; Sharma, O.P.; Jain, S.L.; Khatri, O.P. Reduced graphene oxide–CuO nanocomposites for photocatalytic conversion of CO2 into methanol under visible light irradiation. Appl. Catal. B 2016, 181, 352–362. [Google Scholar] [CrossRef]
  8. Shi, H.; Chen, G.; Zhang, C.; Zou, Z. Polymeric g-C3N4 coupled with NaNbO3 nanowires toward enhanced photocatalytic reduction of CO2 into renewable fuel. ACS Catal. 2014, 4, 3637–3643. [Google Scholar] [CrossRef]
  9. Li, H.; Li, C.; Han, L.; Li, C.; Zhang, S. Photocatalytic reduction of CO2 with H2O on CuO/TiO2 catalysts. Energy Sources 2016, 38, 420–426. [Google Scholar] [CrossRef]
  10. Sun, Q.; Lv, K.; Zhang, Z.; Li, M.; Li, B. Effect of contact interface between TiO2 and g-C3N4 on the photoreactivity of g-C3N4/TiO2 photocatalyst:(0 0 1) vs. (1 0 1) facets of TiO2. Appl. Catal. B. 2015, 164, 420–427. [Google Scholar]
  11. Eddaoudi, M.; Sava, D.F.; Eubank, J.F.; Adil, K.; Guillerm, V. Zeolite-like metal–organic frameworks (ZMOFs): Design, synthesis, and properties. Chem. Soc. Rev. 2015, 44, 228–249. [Google Scholar] [CrossRef] [PubMed]
  12. McGuire, C.V.; Forgan, R.S. The surface chemistry of metal–organic frameworks. Chem. Commun. 2015, 51, 5199–5217. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Wang, C.-C.; Li, J.-R.; Lv, X.-L.; Zhang, Y.-Q.; Guo, G. Photocatalytic organic pollutants degradation in metal–organic frameworks. Energy Environ. Sci. 2014, 7, 2831–2867. [Google Scholar] [CrossRef]
  14. Meyer, K.; Ranocchiari, M.; van Bokhoven, J.A. Metal organic frameworks for photo-catalytic water splitting. Energy Environ. Sci. 2015, 8, 1923–1937. [Google Scholar] [CrossRef]
  15. Wang, S.; Wang, X. Multifunctional metal–organic frameworks for photocatalysis. Small 2015, 11, 3097–3112. [Google Scholar] [CrossRef] [PubMed]
  16. Zhao, Y.; Zhang, J.; Song, J.; Li, J.; Liu, J.; Wu, T.; Zhang, P.; Han, B. Ru nanoparticles immobilized on metal–organic framework nanorods by supercritical CO2-methanol solution: Highly efficient catalyst. Green Chem. 2011, 13, 2078–2082. [Google Scholar] [CrossRef]
  17. Zhang, T.; Lin, W. Metal–organic frameworks for artificial photosynthesis and photocatalysis. Chem. Soc. Rev. 2014, 43, 5982–5993. [Google Scholar] [CrossRef] [PubMed]
  18. Phan, A.; Doonan, C.J.; Uribe-Romo, F.J.; Knobler, C.B.; O’keeffe, M.; Yaghi, O.M. Synthesis, structure, and carbon dioxide capture properties of zeolitic imidazolate frameworks. Acc. Chem. Res. 2010, 43, 58–67. [Google Scholar] [CrossRef] [PubMed]
  19. Liu, Q.; Low, Z.-X.; Li, L.; Razmjou, A.; Wang, K.; Yao, J.; Wang, H. ZIF-8/Zn2GeO4 nanorods with an enhanced CO2 adsorption property in an aqueous medium for photocatalytic synthesis of liquid fuel. J. Mater. Chem. A 2013, 1, 11563–11569. [Google Scholar] [CrossRef]
  20. Jing, H.-P.; Wang, C.-C.; Zhang, Y.-W.; Wang, P.; Li, R. Photocatalytic degradation of methylene blue in ZIF-8. RSC Adv. 2014, 4, 54454–54462. [Google Scholar] [CrossRef]
  21. Tran, U.P.; Le, K.K.; Phan, N.T. Expanding applications of metal–organic frameworks: Zeolite imidazolate framework ZIF-8 as an efficient heterogeneous catalyst for the knoevenagel reaction. ACS Catal. 2011, 1, 120–127. [Google Scholar] [CrossRef]
  22. Schejn, A.; Aboulaich, A.; Balan, L.; Falk, V.; Lalevée, J.; Medjahdi, G.; Aranda, L.; Mozet, K.; Schneider, R. Cu2+-doped zeolitic imidazolate frameworks (ZIF-8): Efficient and stable catalysts for cycloadditions and condensation reactions. Catal. Sci. Technol. 2015, 5, 1829–1839. [Google Scholar] [CrossRef]
  23. Yang, X.; Wen, Z.; Wu, Z.; Luo, X. Synthesis of ZnO/ZIF-8 hybrid photocatalysts derived from ZIF-8 with enhanced photocatalytic activity. Inorg. Chem. Front. 2018, 5, 687–693. [Google Scholar] [CrossRef]
  24. Dutta, T.; Bagchi, D.; Pal, S.K. Bimetallic zeolitic imidazolate framework as an active excipient of curcumin under physiological condition. Biomed. Phys. Eng. Exp. 2018, 4, 055004. [Google Scholar] [CrossRef]
  25. Wang, S.; Yao, W.; Lin, J.; Ding, Z.; Wang, X. Cobalt Imidazolate Metal–Organic Frameworks Photosplit CO2 under Mild Reaction Conditions. Angew. Chem. 2014, 126, 1052–1056. [Google Scholar] [CrossRef]
  26. Wang, Z.; Teramura, K.; Hosokawa, S.; Tanaka, T. Photocatalytic conversion of CO2 in water over Ag-modified La2Ti2O7. Appl. Catal. B 2015, 163, 241–247. [Google Scholar] [CrossRef]
  27. Slamet, H.W.N.; Purnama, E.; Riyani, K.; Gunlazuardi, J. Effect of copper species in a photocatalytic synthesis of methanol from carbon dioxide over copper-doped titania catalysts. World Appl. Sci. J. 2009, 6, 112–122. [Google Scholar]
  28. Sun, D.; Fu, Y.; Liu, W.; Ye, L.; Wang, D.; Yang, L.; Fu, X.; Li, Z. Studies on Photocatalytic CO2 Reduction over NH2-Uio-66 (Zr) and Its Derivatives: Towards a Better Understanding of Photocatalysis on Metal–Organic Frameworks. Chem. Eur. J. 2013, 19, 14279–14285. [Google Scholar] [CrossRef] [PubMed]
  29. Xiang, T.; Xin, F.; Chen, J.; Wang, Y.; Yin, X.; Shao, X. Selective photocatalytic reduction of CO2 to methanol in CuO-loaded NaTaO3 nanocubes in isopropanol. Beilstein J. Nanotech. 2016, 7, 776. [Google Scholar] [CrossRef] [PubMed]
  30. Zhang, Z.; Xian, S.; Xi, H.; Wang, H.; Li, Z. Improvement of CO2 adsorption on ZIF-8 crystals modified by enhancing basicity of surface. Chem. Eng. Sci. 2011, 66, 4878–4888. [Google Scholar] [CrossRef]
  31. Park, K.S.; Ni, Z.; Côté, A.P.; Choi, J.Y.; Huang, R.; Uribe-Romo, F.J.; Chae, H.K.; O’Keeffe, M.; Yaghi, O.M. Exceptional chemical and thermal stability of zeolitic imidazolate frameworks. Proc. Natl. Acad. Sci. USA 2006, 103, 10186–10191. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Cravillon, J.; Nayuk, R.; Springer, S.; Feldhoff, A.; Huber, K.; Wiebcke, M. Controlling zeolitic imidazolate framework nano-and microcrystal formation: Insight into crystal growth by time-resolved in situ static light scattering. Chem. Mater. 2011, 23, 2130–2141. [Google Scholar] [CrossRef]
  33. Pan, Y.; Liu, Y.; Zeng, G.; Zhao, L.; Lai, Z. Rapid synthesis of zeolitic imidazolate framework-8 (ZIF-8) nanocrystals in an aqueous system. Chem. Commun. 2011, 47, 2071–2073. [Google Scholar] [CrossRef] [PubMed]
  34. Khan, I.U.; Othman, M.H.D.; Jilani, A.; Ismail, A.F.; Hashim, H.; Jaafar, J.; Rahman, M.A.; Rehman, G.U. Economical, environmental friendly synthesis, characterization for the production of zeolitic imidazolate framework-8 (ZIF-8) nanoparticles with enhanced CO2 adsorption. Arab. J. Chem. 2018, 11, 1072–1083. [Google Scholar] [CrossRef]
  35. Wu, H.; Qian, X.; Zhu, H.; Ma, S.; Zhu, G.; Long, Y. Controlled synthesis of highly stable zeolitic imidazolate framework-67 dodecahedra and their use towards the templated formation of a hollow Co3O4 catalyst for CO oxidation. RSC Adv. 2016, 6, 6915–6920. [Google Scholar] [CrossRef]
  36. Chary, K.V.R.; Sagar, G.V.; Naresh, D.; Seela, K.K.; Sridhar, B. Characterization and Reactivity of Copper Oxide Catalysts Supported on TiO2−ZrO2. J. Phy. Chem. B 2005, 109, 9437–9444. [Google Scholar] [CrossRef] [PubMed]
  37. Goyal, S.; Shaharun, M.S.; Kait, C.F. Characterization of copper (ii)-zirconium (iv)-imidazolate framework synthesized by hydrothermal method. IIOAB J. 2016, 7, 5. [Google Scholar]
  38. Jeyalakshmi, V.; Mahalakshmy, R.; Krishnamurthy, K.; Viswanathan, B. Titania Based Catalysts for Photoreduction of Carbon Dioxide: Role of Modifiers; NISCAIR-CSIR: New Delhi, India, 2012. [Google Scholar]
  39. Li, H.; Lei, Y.; Huang, Y.; Fang, Y.; Xu, Y.; Zhu, L.; Li, X. Photocatalytic reduction of carbon dioxide to methanol by Cu2O/SiC nanocrystallite under visible light irradiation. J. Nat. Gas Chem. 2011, 20, 145–150. [Google Scholar] [CrossRef]
  40. Uddin, M.R.; Khan, M.R.; Rahman, M.W.; Yousuf, A.; Cheng, C.K. Photocatalytic reduction of CO2 into methanol over CuFe2O4/TiO2 under visible light irradiation. React. Kinet. Mech. Catal. 2015, 116, 589–604. [Google Scholar] [CrossRef]
  41. Grosu, Y.; Renaudin, G.; Eroshenko, V.; Nedelec, J.-M.; Grolier, J.-P. Synergetic effect of temperature and pressure on energetic and structural characteristics of {ZIF-8+ water} molecular spring. Nanoscale 2015, 7, 8803–8810. [Google Scholar] [CrossRef] [PubMed]
  42. Zhang, Y.; Li, Q.; Liu, C.; Shan, X.; Chen, X.; Dai, W.; Fu, X. The promoted effect of a metal-organic frameworks (ZIF-8) on Au/TiO2 for CO oxidation at room temperature both in dark and under visible light irradiation. Appl. Catal. B 2018, 224, 283–294. [Google Scholar] [CrossRef]
  43. Qin, S.; Xin, F.; Liu, Y.; Yin, X.; Ma, W. Photocatalytic reduction of CO2 in methanol to methyl formate over CuO–TiO2 composite catalysts. J. Colloid Interface Sci. 2011, 356, 257–261. [Google Scholar] [CrossRef] [PubMed]
  44. Jiang, X.; Lin, Q.; Zhang, M.; He, G.; Sun, Z. Microstructure, optical properties, and catalytic performance of Cu2 O-modified ZnO nanorods prepared by electrodeposition. Nanoscale Res. Lett. 2015, 10, 30. [Google Scholar] [CrossRef] [PubMed]
  45. Li, G.; Dimitrijevic, N.M.; Chen, L.; Rajh, T.; Gray, K.A. Role of surface/interfacial Cu2+ sites in the photocatalytic activity of coupled CuO−TiO2 nanocomposites. J. Phy. Chem. C 2008, 112, 19040–19044. [Google Scholar] [CrossRef]
  46. Colon, G.; Maicu, M.; Hidalgo, M.S.; Navio, J. Cu-doped TiO2 systems with improved photocatalytic activity. Appl. Catal. B 2006, 67, 41–51. [Google Scholar] [CrossRef]
  47. Yang, M.; Zhu, L.; Li, Y.; Cao, L.; Guo, Y. Asymmetric interface band alignments of Cu2O/ZnO and ZnO/Cu2O heterojunctions. J. Alloy. Compd. 2013, 578, 143–147. [Google Scholar] [CrossRef]
  48. Li, J.; Luo, D.; Yang, C.; He, S.; Chen, S.; Lin, J.; Zhu, L.; Li, X. Copper (II) imidazolate frameworks as highly efficient photocatalysts for reduction of CO2 into methanol under visible light irradiation. J. Solid State Chem. 2013, 203, 154–159. [Google Scholar] [CrossRef]
  49. Kudo, A.; Miseki, Y. Heterogeneous photocatalyst materials for water splitting. Chem. Soc. Rev. 2009, 38, 253–278. [Google Scholar] [CrossRef] [PubMed]
  50. Liu, Y.; Yang, Y.; Sun, Q.; Wang, Z.; Huang, B.; Dai, Y.; Qin, X.; Zhang, X. Chemical adsorption enhanced CO2 capture and photoreduction over a copper porphyrin based metal organic framework. ACS Appl. Mater. Interfaces 2013, 5, 7654–7658. [Google Scholar] [CrossRef] [PubMed]
  51. Nadeem, S.; Mumtaz, A.; Mumtaz, M.; Mutalib, M.A.; Shaharun, M.S.; Abdullah, B. Visible light driven CO2 reduction to methanol by Cu-porphyrin impregnated mesoporous Ti-MCM-48. J. Mol. Liq. 2018. [Google Scholar] [CrossRef]
  52. Wang, J.; Chen, K.; Shen, Y.; Wang, X.; Guo, Y.; Zhou, X.; Bai, R. Enhanced photocatalytic degradation for organic pollutants by a novel m-Bi2O4/Bi2O2CO3 photocatalyst under visible light. Res. Chem. Intermediat. 2018, 44, 3061–3079. [Google Scholar] [CrossRef]
  53. Ren, M.; Valsaraj, K. Inverse opal titania on optical fiber for the photoreduction of CO2 to CH3OH. Int. J. Chem. React. Eng. 2009, 7. [Google Scholar] [CrossRef]
  54. Usubharatana, P.; McMartin, D.; Veawab, A.; Tontiwachwuthikul, P. Photocatalytic process for CO2 emission reduction from industrial flue gas streams. Ind. Eng. Chem. Res. 2006, 45, 2558–2568. [Google Scholar] [CrossRef]
  55. Linsebigler, A.L.; Lu, G.; Yates, J.T., Jr. Photocatalysis on TiO2 surfaces: Principles, mechanisms, and selected results. Chem. Rev. 1995, 95, 735–758. [Google Scholar] [CrossRef]
  56. Tahir, M.; Tahir, B.; Amin, N.A.S.; Alias, H. Selective photocatalytic reduction of CO2 by H2O/H2 to CH4 and CH3OH over Cu-promoted In2O3/TiO2 nanocatalyst. Appl. Surf. Sci. 2016, 389, 46–55. [Google Scholar] [CrossRef]
Figure 1. XRD images of (a) ZIF-8 (b) 2Cu/ZIF-8N4 (c) 2Cu/ZIF-8N3 (d) 2Cu/ZIF-8N2.
Figure 1. XRD images of (a) ZIF-8 (b) 2Cu/ZIF-8N4 (c) 2Cu/ZIF-8N3 (d) 2Cu/ZIF-8N2.
Catalysts 08 00581 g001
Figure 2. FESEM images of (A) (a) ZIF-8 (b) 2Cu/ZIF-8N4 (c) 2Cu/ZIF-8N3 (d) 2Cu/ZIF-8N2 and (B) dot mapping images of 2Cu/ZIF-8N2.
Figure 2. FESEM images of (A) (a) ZIF-8 (b) 2Cu/ZIF-8N4 (c) 2Cu/ZIF-8N3 (d) 2Cu/ZIF-8N2 and (B) dot mapping images of 2Cu/ZIF-8N2.
Catalysts 08 00581 g002
Figure 3. (A) UV–vis absorbance spectra of (a) ZIF-8 (b) 2Cu/ZIF-8N4 (c) 2Cu/ZIF-8N3 (d) 2Cu/ZIF-8N2 (B) Optical band edges of (a) ZIF-8 (b) 2Cu/ZIF-8N4 (c) 2Cu/ZIF-8N3 (d) 2Cu/ZIF-8N2.
Figure 3. (A) UV–vis absorbance spectra of (a) ZIF-8 (b) 2Cu/ZIF-8N4 (c) 2Cu/ZIF-8N3 (d) 2Cu/ZIF-8N2 (B) Optical band edges of (a) ZIF-8 (b) 2Cu/ZIF-8N4 (c) 2Cu/ZIF-8N3 (d) 2Cu/ZIF-8N2.
Catalysts 08 00581 g003
Figure 4. XRD spectra of (a) 1Cu/ZIF-8N2 (b) 2Cu/ZIF-8N2 (c) 3Cu/ZIF-8N2.
Figure 4. XRD spectra of (a) 1Cu/ZIF-8N2 (b) 2Cu/ZIF-8N2 (c) 3Cu/ZIF-8N2.
Catalysts 08 00581 g004
Figure 5. FESEM images of (a) 1Cu/ZIF-8N2 (b) 2Cu/ZIF-8N2 (c) 3Cu/ZIF-8N2 and (d) EDX image of 2Cu/ZIF-8N2.
Figure 5. FESEM images of (a) 1Cu/ZIF-8N2 (b) 2Cu/ZIF-8N2 (c) 3Cu/ZIF-8N2 and (d) EDX image of 2Cu/ZIF-8N2.
Catalysts 08 00581 g005
Figure 6. Ultra-violet visible absorbance spectra of (a) 1Cu/ZIF-8N2 (b) 2Cu/ZIF-8N2 (c) 3Cu/ZIF-8N2.
Figure 6. Ultra-violet visible absorbance spectra of (a) 1Cu/ZIF-8N2 (b) 2Cu/ZIF-8N2 (c) 3Cu/ZIF-8N2.
Catalysts 08 00581 g006
Figure 7. (A) XPS Cu2p spectra with peak fitted graph of (a) 1Cu/ZIF-8N2 (b) 2Cu/ZIF-8N2 (c) 3Cu/ZIF-8N2 (B) Zn2p spectra of (a) 1Cu/ZIF-8N2 (b) 2Cu/ZIF-8N2 (c) 3Cu/ZIF-8N2.
Figure 7. (A) XPS Cu2p spectra with peak fitted graph of (a) 1Cu/ZIF-8N2 (b) 2Cu/ZIF-8N2 (c) 3Cu/ZIF-8N2 (B) Zn2p spectra of (a) 1Cu/ZIF-8N2 (b) 2Cu/ZIF-8N2 (c) 3Cu/ZIF-8N2.
Catalysts 08 00581 g007aCatalysts 08 00581 g007b
Figure 8. Photocatalytic reduction of CO2 to CH3OH over all Cu/ZIF-8 catalysts at 6 h irradiation.
Figure 8. Photocatalytic reduction of CO2 to CH3OH over all Cu/ZIF-8 catalysts at 6 h irradiation.
Catalysts 08 00581 g008
Figure 9. Schematic reduction activity of CO2 to methanol over Cu/ZIF-8 photocatalyst under visible light irradiation.
Figure 9. Schematic reduction activity of CO2 to methanol over Cu/ZIF-8 photocatalyst under visible light irradiation.
Catalysts 08 00581 g009
Figure 10. Schematic diagram of the reactor setup.
Figure 10. Schematic diagram of the reactor setup.
Catalysts 08 00581 g010
Table 1. Recent Progress in photocatalytically reduction of CO2.
Table 1. Recent Progress in photocatalytically reduction of CO2.
Sr. No.CatalystReactantPreparation MethodReactor ConditionProduct/YieldRef.
1Cu porphyrin-based MOFCO2 and Solution of the triethylamineHydrothermal300 W Xe lamp5.97 µmol/g·h
CH3OH
[50] 2013
2ZIF-8/Zn2GeO4CO2 and the Solution of sodium sulfiteImpregnation500 W Xe arc lamp2.44 µmol/g
CH3OH
[19] 2013
3CuImCO2 and solutions of sodium hydroxide and sodium sulfiteHydrothermal500 W, Xe lamp1712.7 µmol/g
CH3OH
[48] 2013
4Ti-MCM-48(25)CuTPPCO2 and solutions of sodium hydroxide and sodium sulfiteImpregnation500 W Xenon Lamp297.06 μmol/g
CH3OH
[51] 2018
5TiO2P25CO2 and solution of potassium hydrogen carbonateCommercial10 W UV Lamp800 µmol/g·catal
CH3OH
[27] 2009
62Cu/ZIF-8N2CO2 and solutions of sodium hydroxide and sodium sulfiteHydrothermal500 W Xenon Lamp35.82 µmol/L·g
CH3OH
This study
Table 2. Catalysts composition with Cu loading and ammonium hydroxides concentrations.
Table 2. Catalysts composition with Cu loading and ammonium hydroxides concentrations.
S. No.Copper Metal Loading (mmol)ZIF-8 Loading (mmol)Ammonium Hydroxide Concentration (M)Catalyst Notation
1.2442Cu/ZIF-8N4
2.2432Cu/ZIF-8N3
3.2422Cu/ZIF-8N2
4.1421Cu/ZIF-8N2
5.3423Cu/ZIF-8N2

Share and Cite

MDPI and ACS Style

Goyal, S.; Shaharun, M.S.; Kait, C.F.; Abdullah, B.; Ameen, M. Photoreduction of Carbon Dioxide to Methanol over Copper Based Zeolitic Imidazolate Framework-8: A New Generation Photocatalyst. Catalysts 2018, 8, 581. https://doi.org/10.3390/catal8120581

AMA Style

Goyal S, Shaharun MS, Kait CF, Abdullah B, Ameen M. Photoreduction of Carbon Dioxide to Methanol over Copper Based Zeolitic Imidazolate Framework-8: A New Generation Photocatalyst. Catalysts. 2018; 8(12):581. https://doi.org/10.3390/catal8120581

Chicago/Turabian Style

Goyal, Sonam, Maizatul Shima Shaharun, Chong Fai Kait, Bawadi Abdullah, and Mariam Ameen. 2018. "Photoreduction of Carbon Dioxide to Methanol over Copper Based Zeolitic Imidazolate Framework-8: A New Generation Photocatalyst" Catalysts 8, no. 12: 581. https://doi.org/10.3390/catal8120581

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop