Next Article in Journal
Transformation of Enzymatic Hydrolysates of Chlorella–Fungus Mixed Biomass into Poly(hydroxyalkanoates)
Previous Article in Journal
Oxygenated Hydrocarbons from Catalytic Hydrogenation of Carbon Dioxide
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

ZIF-67 Derived Cu-Co Mixed Oxides for Efficient Catalytic Oxidation of Formaldehyde at Low-Temperature

1
Department of Life Sciences, Changzhi University, Changzhi 046011, China
2
College of Environmental Science and Engineering, Taiyuan University of Technology, Jinzhong 030600, China
*
Author to whom correspondence should be addressed.
Catalysts 2023, 13(1), 117; https://doi.org/10.3390/catal13010117
Submission received: 26 November 2022 / Revised: 24 December 2022 / Accepted: 3 January 2023 / Published: 5 January 2023
(This article belongs to the Section Environmental Catalysis)

Abstract

:
It is still an intractable problem to exploit high-efficient Co-based catalysts for low-temperature HCHO oxidation. Herein, we synthesized a series of Cu-doped Co3O4 catalysts (Cu1Co8, Cu1Co4, and Cu1Co2 corresponded to 1/8, 1/4, and 1/2 of Cu/Co molar ratios, respectively) via in situ pyrolysis of bimetal Cu-ZIF-67 precursors and the pure Co3O4 sample was also prepared through directly annealing monometal ZIF-67 for comparison. Performance tests of HCHO oxidation found that Cu doping remarkably enhanced the low-temperature HCHO oxidation performance of Co3O4 sample, and thereinto the Cu1Co4 possessed the optimal HCHO oxidation activity, which achieved 90% HCHO conversion at 108 °C. The characterization results revealed that the stronger interaction between Cu and Co species (Co2+ + Cu2+ ↔ Co3+ + Cu+) of Cu1Co4 not only facilitates the formation of defect sites, Co3+ and surface adsorbed oxygen species but also improves its low-temperature reducibility, and consequently resulting in its superior HCHO oxidation performance. Furthermore, the in-situ DRIFTS results suggested that the formaldehyde oxidation over Cu1Co4 followed HCHO → H2CO2 → HCOO → CO32− → CO2 pathway. The present work provides a novel and facile approach to fabricating highly effective Co-based catalysts for low-temperature HCHO oxidation.

1. Introduction

Indoor formaldehyde (HCHO) is primarily generated from numerous building/upholstery materials and is one of the most harmful indoor airborne pollutants on account of its strong irritation, reproductive toxicity, and potential carcinogenesis [1,2]. In order to reduce the hazard of HCHO to residents, governments and organizations around the world have promulgated a series of recommended HCHO exposure limits. According to the World Health Organization (WHO, 2000), the HCHO guideline value in indoor air is set at 0.1 mg/m3 [3]. In Singapore, the HCHO recommendation value should not exceed 0.12 mg/m3 [4]. Recently, China has revised and issued the new “Standards for indoor air quality” (GB/T 18883-2022), in which the HCHO concentration threshold is below 0.08 mg/m3 [5]. Hence, it is imperative to explore high-efficient and low-cost HCHO elimination technologies for improving the living environment of people and meeting more and more rigorous policies. In the past decades, various removal techniques, including adsorption [6], plasma decomposition [7], photocatalysis [8], and thermal catalytic oxidation [9,10], have been studied and employed for HCHO purity. Amongst them, thermocatalysis is deemed a particularly suitable and promising approach owing to its high HCHO degradation and mineralization efficiency, excellent reusability, and no secondary pollution [11,12,13,14].
Up to now, a number of catalysts for HCHO oxidation have been designed and synthesized [15,16,17,18,19,20]. Although precious-metal-based catalysts (Pt/TiO2, Pd/CeO2, Au/FeOx, Ir/Al2O3, etc.) manifested exceptional HCHO oxidation performance [21,22,23,24,25,26,27], the scarce terrestrial abundance and exorbitant prices of precious metals seriously restrict their large-scale application [28,29,30]. Therefore, abundant and low-cost transition-metal oxides have aroused increasing interest. Particularly, spinel-type Co3O4-based catalysts have been extensively investigated and utilized for HCHO catalytic oxidation profiting from their excellent reducibility and stability [31,32,33]. For instance, Lv et al. [34] reported that the cellular-like Co3O4 catalyst synthesized through a seeded method could totally oxidize HCHO at 165 °C. Bai et al. [35] compared the HCHO oxidation activity of nano-Co3O4, 2D-Co3O4, and 3D-Co3O4 catalysts and discovered that the 3D-Co3O4 catalyst possessed the optimal catalytic performance, mainly deriving from its unique 3D pore channel structure. However, the low-temperature O2 activation ability of pristine Co3O4 catalysts is relatively poor, which in turn results in the accumulation of HCHO oxidation intermediates on active sites and deactivation of catalysts at low-temperature [31,36].
Generally, ion doping is an effective strategy to improve the O2 activation ability of Co3O4 catalysts due to the strong interaction between cobalt and doping ions [37,38,39,40]. Thereinto, copper (Cu) element is considered an excellent dopant that benefits from its low valency and large ion radius, which could weaken the strength of the Co-O bond, as well as promote the generation of defects [41,42]. For example, Fan et al. [43] prepared a series of CuaCo1−aOx (a = 0.1, 0.2, 0.4, 0.6) catalysts through a solvothermal method. With an increase in the Cu doping amount, the toluene oxidation activity of the resultant catalysts was first increased and then decreased. Among them, the Cu0.4Co0.6Ox catalyst was the most active sample, which was ascribed to its adequate Co3+ and surface-adsorbed oxygen species. Moreover, Zhang et al. [44] also found that the incorporation of appropriate amounts of Cu into Co3O4 could facilitate O2 activation and thereby promote the oxidation of toluene and propane. As is well known, O2 activation is regarded as a crucial step during low-temperature HCHO oxidation, and adequate surface-adsorbed oxygen species could boost the oxidation of HCHO and intermediates [45,46]. Inspired by these, we speculate that Cu-doping might enhance the HCHO degradation performance of Co3O4 at low-temperature. Nevertheless, to our knowledge, the study of the Cu dopant effect on the HCHO oxidation activity of Co3O4 is still scarce and needs to be further investigated.
As a novel class of porous crystalline materials, zeolitic imidazolate framework-67 (ZIF-67) consisting of Co2+ and imidazole linkers has drawn great research interest owing to its high surface area, tunable architectures, and compositions in recent years [47]. More importantly, the ZIF-67 has been extensively applied as the precursor to synthesize various Co-based porous materials (e.g., Co3O4, NiCo2O4, CoS) by means of its unique thermal behavior and chemical reactivity [48,49,50,51]. In this work, we prepared a series of Cu-doped Co3O4 catalysts via in situ pyrolysis of bimetal Cu-ZIF-67 precursors. Compared with the pure Co3O4 catalyst, these Cu-doped Co3O4 samples displayed remarkably improved low-temperature HCHO oxidation performance, especially Cu1Co4, which exhibited the highest catalytic activity, achieving 90% of HCHO conversion at 108 °C. Comprehensive characterizations were carried out to elucidate the effect of Cu doping on the physicochemical properties of Co3O4, and the structure-activity relationship was discussed. Moreover, the plausible HCHO oxidation pathway over the Cu1Co4 catalyst was also explored by in-situ DRIFTS.

2. Results and Discussion

2.1. HCHO Oxidation Performance Tests

The HCHO oxidation ignition curves of Co3O4, Cu1Co8, Cu1Co4, and Cu1Co2 are shown in Figure 1. The HCHO oxidation activity of catalysts exhibited a volcano-shaped relationship that initially increased and subsequently decreased with the increase in the Cu/Co ratio. Thereinto, the Cu1Co4 displayed the optimal HCHO oxidation activity. The ignition temperatures of 50% and 90% HCHO conversion (T50 and T90) of catalysts were summarized in Table 1. The T50 and T90 of Co3O4 were 147 and 157 °C, respectively. By contrast, for the Cu1Co8, Cu1Co4, and Cu1Co2, their T50, respectively, decreased to 103, 98, and 101 °C. Meanwhile, their T90 were as low as 122, 108, and 112 °C. Therefore, the HCHO degradation performance of catalysts was ranked as follows: Cu1Co4 > Cu1Co2 > Cu1Co8 > Co3O4. The above results demonstrated that moderate Cu doping could effectively improve the HCHO oxidation ability of Co3O4.
In view of the ineluctable existence of moisture in the actual indoor air, the effect of relative humidity on the HCHO oxidation activity of Cu1Co4 was assessed through a transient response experiment at 110 °C (Figure 2a). In the first 100 min, the HCHO conversion of Cu1Co4 was ~95% in the presence of H2O. Next, when H2O was removed from the feed gas, the HCHO conversion of Cu1Co4 gradually decreased and remained steady at ~83%. Whereas the HCHO oxidation activity of Cu1Co4 was completely recovered after the reintroduction of H2O, demonstrating that H2O had a promotional effect on Cu1Co4 for HCHO oxidation. It could be because H2O accelerated the activation of O2, which was regarded as a key step for HCHO oxidation [52]. Moreover, the stability of Cu1Co4 in the HCHO oxidation was evaluated by a 48-h continuous test at 110 °C. As depicted in Figure 2b, the Cu1Co4 maintained ~95% HCHO conversion, and no evident performance deterioration could be discerned during the whole test process, indicating the good durability of Cu1Co4.

2.2. Physicochemical Property Analyses

In order to determine the crystal structure and crystalline size of Co3O4, Cu1Co8, Cu1Co4, and Cu1Co2, XRD measurements were carried out (Figure 3a). The Co3O4 sample exhibited nine well-defined diffraction peaks at 19.1°, 31.3°, 36.9°, 38.5°, 44.9°, 55.7°, 59.4°, 65.3°, 77.4°, corresponding to (111), (220), (311), (222), (400), (422), (511), (440), (533) lattice planes of spinel-type Co3O4 (JCPDS 42-1467), respectively [53]. As for the Cu1Co8, Cu1Co4, and Cu1Co2, their XRD patterns were similar to the Co3O4 sample, and no metallic copper or copper oxides-related diffraction peaks were detected, possibly resulting from the high dispersion of copper species or the incorporation of Cu ions into the crystalline lattice of Co3O4. Of note, the (311) diffraction peaks of these Cu-Co mixed oxides shifted towards lower angles compared with pure Co3O4 (Figure 3b). It might be owing to the substitution of smaller Co2+ (0.65 Å) and Co3+ (0.55 Å) cations by larger Cu2+ (0.73 Å) cations, which would enlarge the lattice constant of Co3O4 [40,54]. The crystalline size of all catalysts was determined via the Scherrer formula based on the strongest (311) planes and summarized in Table 2. It could be found that there was no significant crystalline size difference between the pristine Co3O4 and these Cu-Co mixed oxides.
The Raman spectra of samples are displayed in Figure 4a. The typical Raman-active peaks of crystalline Co3O4 were observed in all samples. In detail, the peaks at around 184, 465, 506, 600, and 654 cm−1 were attributed to the F 2 g ( 1 ) , Eg, F 2 g ( 2 ) , F 2 g ( 3 ) , and A1g vibration modes of Co3O4, respectively [44]. In addition, no evident characteristic bands of CuOx were detected in the Cu1Co8, Cu1Co4, and Cu1Co2. As the Cu/Co ratio increased, the Raman peak positions of samples gradually shifted to higher wavenumbers (i.e., blue shift), which could be derived from the photon-confinement effect induced by the surface defects of Co3O4 [42,55]. Therefore, the above result demonstrated that Cu doping could effectively increase the surface defect concentration of Co3O4. Usually, adequate surface defect sites are advantageous to O2 activation and generation of surface-adsorbed xygen species, which are vital for low-temperature HCHO oxidation [45,46].
The FT-IR spectra of samples are illustrated in Figure 4b. For all catalysts, the broad peak at 3700–3000 cm−1 and weak peak at about 1630 cm−1 were assigned to the υ(OH) and δ(OH) of surface adsorbed water [43]. Meanwhile, two strong stretching vibration peaks located at 663 and 566 cm−1 were observed, which corresponded to the characteristic absorption peaks of four-coordinated Co2+-O and six-coordinated Co3+-O, respectively [42]. Likewise, no typical Cu-O characteristic absorption bands were detected over Cu1Co8, Cu1Co4, and Cu1Co2 catalysts.
Figure 5a–d shows the SEM images of Co3O4 and Cu-doped Co3O4 samples. For all samples, their SEM images exhibited well-defined cubic morphology, and their average size was approximately 170 nm, which suggested that Cu-doping did not obviously change the shape of the Co3O4 sample. In order to study the element distribution of Cu1Co4, EDS mapping was carried out. As depicted in Figure 5e, the Co, Cu, and O elements were evenly dispersed throughout the entire selected region. Furthermore, more detailed structure characteristics of Cu1Co4 were further determined by TEM. As displayed in Figure 6a, the cube-like Cu1Co4 consisted of numerous aggregated nanoparticles with a size of 10–15 nm, in good conformity with the estimated result of XRD (~14 nm). Simultaneously, the piled pores generated by the stacking of nanoparticles could also be discerned. The HRTEM images (Figure 6b,c) clearly showed the interplanar distance of 0.46, 0.28, and 0.24 nm, corresponding to the (111), (220), and (311) crystal planes of Co3O4, respectively [29,42]. In addition, no lattice fringes associated with metallic copper or copper oxides were observed. Combined with the XRD and EDS mapping results (Figure 3b and Figure 5e), it could be speculated that Cu ions were incorporated into the crystalline lattice of Co3O4.
To reveal the specific surface area and porous structure of Co3O4, Cu1Co8, Cu1Co4, and Cu1Co2, the N2 adsorption-desorption isotherm tests were performed (Figure 7). It could be seen that all samples exhibited the characteristics of IV isotherms with type H3 hysteresis loops, indicating the presence of piled mesopores [27], which was in line with the TEM results (Figure 6a). Meanwhile, the BJH pore size distribution results also attested to the existence of mesopores. The textural parameters of samples are summarized in Table 2. Interestingly, compared with Cu-doped Co3O4 samples, the pure Co3O4 catalyst possessed relatively larger SBET and Vp but displayed the worst HCHO degradation performance, demonstrating that the SBET and Vp of catalysts might not be the major factors that determine their HCHO oxidation performance.
XPS tests were carried out to elucidate the surface element compositions and chemical states of catalysts. As illustrated in Figure 8a, the Co 2p spectra of samples displayed a spin-orbit doublet at a binding energy of 779.5–780.0 and 794.6–795.1 eV, which was identified as Co 2p3/2 and Co 2p1/2, respectively [53]. The ΔECo 2p (energy differences between Co 2p3/2 peak and Co 2p1/2 peak) of all samples were ca. 15.1 eV, suggesting the coexistence of Co2+ and Co3+ species [29]. It is interesting that the Co 2p profiles of Cu-Co samples moved towards lower binding energy (~0.5 eV) in comparison to the pure Co3O4 sample, revealing that there was a mutual interaction between Cu and Co species [43]. Furthermore, the surface Co3+/Co2+ molar ratios of samples were summarized in Table 3 and obeyed the sequence as follows: Cu1Co4 (0.78) > Cu1Co2 (0.69) > Cu1Co8 (0.59) > Co3O4 (0.53). Apparently, the Cu doping significantly increased the Co3+/Co2+ ratios of samples, which could be owing to the mutual interaction between Cu and Co species (Cu2+ + Co2+ ↔ Cu+ + Co3+) [42]. It is well-known that the Co3+ species possess strong oxidation ability and abundant Co3+ is beneficial to enhance the HCHO oxidation performance [31,40,41]. Hence, the Cu1Co4 with the highest Co3+/Co2+ ratio exhibited the best HCHO decomposition ability amongst these catalysts.
As shown in Figure 8b, the asymmetrical Cu 2p3/2 peaks of Cu-Co samples could be fitted into two independent peaks. Specifically, the peak at 931.9 ± 0.1 eV was attributed to Cu+ species, while the other peak at 934.3 ± 0.1 eV belonged to Cu2+ species. The Cu+/Cu2+ molar ratio of Cu1Co4 was much higher than those of Cu1Co8 and Cu1Co2 (0.29 vs. 0.16 and 0.21), which revealed that the mutual interaction between Cu and Co species over Cu1Co4 was stronger compared with Cu1Co8 and Cu1Co2 (Cu2+ + Co2+ ↔ Cu+ + Co3+) [42].
The O 1s spectra of samples (Figure 8c) could be deconvoluted into two components. The peaks located at 529.7–530.1 and 531.5 ± 0.1 eV corresponded to the lattice oxygen (Oβ) and surface adsorbed oxygen (Oα) species, respectively [29]. Noticeably, the binding energy of Oβ centered at 530.1 eV for the pure Co3O4 sample shifted to 529.7 eV for Cu-doped Co3O4 samples, demonstrating the mutual interaction between Cu and Co species changed the coordination environment of lattice oxygen [41]. It is reported that the more electrophilic Cu2+ will draw electrons away from Co in the Cu-doped Co3O4 samples [56]. It means that the interaction between Cu2+–O2−–Co2+ entity makes the charge transfer between Cu and Co cations dexterously as below: Cu2+ + Co2+ ↔ Cu+ + Co3+. Therefore, the electron density of lattice oxygen in the Cu2+–O2−–Co2+ entity would increase because of the electronic delocalization effect, and the binding energy of lattice oxygen shifted towards a lower value due to the shielding effect [57,58,59]. As shown in Table 3, the Oα/Oβ molar ratios of samples decreased in the order of Cu1Co4 (0.55) > Cu1Co2 (0.53) > Cu1Co8 (0.51) > Co3O4 (0.33). The above results indicated that Cu doping effectively enhanced the O2 activation ability of Co3O4. It could be because the mutual interaction between Cu and Co species promotes the generation of surface defects, which is in line with the Raman results (Figure 4a). It has been established that the Oα species is the reactive oxygen species (ROS) in the deep oxidation of HCHO [29,45]. In this view, the Cu1Co4 with the highest Oα/Oβ ratio should show the best HCHO degradation performance.
Normally, the low-temperature reducibility of catalysts is closely associated with their HCHO oxidation activity. H2-TPR experiments were implemented to investigate the effect of the Cu/Co molar ratio on the redox properties of samples, and the corresponding curves are illustrated in Figure 9. For the pristine Co3O4, there existed two well-defined H2 consumption peaks at 355 and 460 °C, which belonged to the sequential reduction of Co3+ → Co2+ → Co0 [60]. It is worth noting that the incorporation of Cu species dramatically decreased the initial reduction temperature of samples, possibly deriving from the mutual interaction between Cu and Co species [44]. For the Cu1Co8, it exhibited a discernible double-peak with maxima at 266 and 317 °C, in which the peak at low-temperature could be attributed to the simultaneous reduction of Cu2+/Cu+ → Cu0 and Co3+ → Co2+ and the latter could be assigned to the reduction of Co2+ → Co0 [29,43,61]. The TPR profiles of the Cu1Co4 were analogous to that of the Cu1Co8, but the reduction temperature of Cu1Co4 further moved to a lower temperature (248 and 303 °C). On the contrary, for the Cu1Co2, its reduction peaks slightly shifted towards a higher temperature (262 and 308 °C), suggesting that excess Cu doping would inhibit the reduction of Cu-Co oxides. Therefore, it was concluded that the incorporation of proper Cu could significantly enhance the low-temperature reducibility of Co3O4, consequently leading to the improvement of HCHO oxidation activity.

2.3. Reaction Pathway

In order to elucidate the mechanism of HCHO oxidation over the Cu1Co4 catalyst, an in-situ DRIFTS measurement was conducted (Figure 10). Several absorption vibration bands were detected during the early stage of exposure to HCHO feed gas. The peaks centered at 1420 and 1436 cm−1 were assigned to the υ(OCO) in dioxymethylene (DOM) species that was generated via the attack of nucleophilic surface adsorbed oxygen species [62]. The peak located at ~1340 cm−1 was ascribed to υs(COO) of formate species; the bands at 1550, 1567, and 1583 cm−1 could be attributed to the υas(COO) of formate species; the peaks at 2884, 2936 and 2966 cm−1 belonged to the υ(CH) of formate species [63,64,65]. As the reaction continued, the characteristic absorption peak of carbonate species (CO32−) at ~1320 cm−1 could also be observed [64,66]. It demonstrated that DOM and formate species were the initial intermediates, which would be further oxidized into carbonate species. In addition, the broad band at 3700–3000 cm–1 and the weak band at ~1635 cm–1 were respectively assigned to υ(OH) and δ(OH) of H2O molecules adsorbed on Cu1Co4 surface, which could be originated from the reactant gas or products of HCHO degradation [29].
Based on the aforementioned results, a probable HCHO degradation route over the Cu1Co4 catalyst is proposed. Firstly, the HCHO molecules adsorbed on the Cu1Co4 catalyst are promptly transformed into DOM species through the attack of nucleophilic surface adsorbed oxygen species, which were formed by the activation of O2 over the defect sites of the Cu1Co4 catalyst. Next, DOM species are converted into formate species via dehydrogenation, and then, formate species further react with surface-adsorbed oxygen species to produce carbonate species. Lastly, carbonate species decompose into CO2 and H2O. The stronger interaction between Cu and Co species (Co2+ + Cu2+ ↔ Co3+ + Cu+) of the Cu1Co4 catalyst endowed it with richer defect sites, higher Co3+/Co2+ ratio, superior O2 activation ability, and low-temperature reducibility, consequently leading to its better HCHO degradation ability.
O2 → 2[O]s
HCHO + [O]s → [H2CO2]s
[H2CO2]s → [HCOO]s + H+
[HCOO]s+ [O]s → [CO32−]s + H+
[CO32−]s + 2H+ → CO2 + H2O

3. Experimental

3.1. Materials

Co(NO3)2·6H2O (99%), Cu(NO3)2·3H2O (99%), and CTAB (99%) and 2-methylimidazole (98%) were purchased from Macklin Ltd. (Shanghai, China). Ethanol (99.7%) was purchased from Fuchen Ltd. (Tianjin, China). Deionized water (DI water) was obtained from UPR-Ⅳ-20L (Ulupure Ltd., Chengdu, China) and used throughout.

3.2. Preparation of Co3O4 and Cu-Co Oxides

Firstly, 9.08 g of 2-methylimidazole was dissolved in 80 mL DI water (marked as solution A). Then, a proper amount of Co(NO3)2·6H2O and Cu(NO3)2·3H2O (Co2+ + Cu2+ = 2 mmol, Cu2+/Co2+ molar ratios were 0, 1/8, 1/4 and 1/2, respectively) and 10 mg CTAB was dissolved in 20 mL DI water (marked as solution B). After that, solution B was quickly injected into solution A, and the obtained mixed solution was vigorously stirred at 25 °C for 1.5 h. These as-synthesized ZIF-67 precursors were harvested by centrifugation, washed 3 times with ethanol, and dried at 70 °C overnight. Finally, the ZIF-67 with different Cu2+/Co2+ molar ratios were annealed at 350 °C for 2 h (1 °C/min) under air atmosphere, and these acquired samples were labeled as Co3O4, Cu1Co8, Cu1Co4, and Cu1Co2, respectively.

3.3. Catalyst Characterization

PXRD measurements were conducted on a Rigaku Smartlab9 X-ray diffractometer (Rigaku, Tokyo, Japan) with Cu Kα radiation (λ = 1.5418 Å) operated at 40 kV and 40 mA. The Raman spectra were recorded on a HORIBA LabRAM HR Evolution Raman spectrometer (Horiba, Paris, France) using a 532 nm He-Ne laser. The FT-IR spectra were collected on a Nicolet iS10 FT-IR spectrometer (Thermo Scientific, Waltham, MA, USA). The morphologies and microstructures of catalysts were observed with FESEM (FEI, Waltham, MA, USA, QUANTA FEG 450) and TEM (Thermo Scientific, Hillsboro, OR, USA, Talos F200X), respectively. N2 adsorption-desorption isotherms of catalysts were determined using an Autosorb-iQ-2 instrument (Quantachrome, Boynton Beach, FL, USA) at −196 °C. XPS spectra of catalysts were recorded on an ESCALAB 250Xi photoelectron spectroscopy (Thermo Scientific, Waltham, MA, USA) with an Al Kα X-ray source (1486.6 eV). The binding energy (B.E.) of spectra was calibrated by the C 1s peak (284.8 eV). H2-TPR was tested in a ChemStar TPx instrument (Quantachrome, Boynton Beach, FL, USA) equipped with a TCD detector. Specifically, 30 mg of catalysts were first pretreated in Ar at 300 °C for 30 min, then cooled to 50 °C in Ar. Afterward, catalysts were reduced in a flow of 5%H2-Ar mixture (30 mL/min) from 50 °C to 700 °C with a ramping rate of 10 °C/min.
The in-situ DRIFTS was performed on a Bruker Tensor 27 FTIR spectrometer (Bruker, Karlsruhe, Germany) equipped with CaF2 windows. The Cu1Co4 catalyst was firstly pretreated in N2 at 150 °C for 30 min to desorb impurities and then cooled to 120 °C to record the background spectrum. Afterward, feed gas containing HCHO (~80 ppm) was introduced into in situ chamber, and the spectrum was collected continuously by background subtraction.

3.4. Catalytic Evaluation

Typically, 100 mg of catalyst (0.30–0.45 mm) was packed in a quartz reactor (i.d. = 4 mm), which was placed in a furnace, and the reaction temperature was controlled by a K-type thermal couple. All gas flows were controlled by mass flow controllers. The composition of reactant gas was 80 ppm HCHO, 60% relative humidity (RH), 20% O2, and N2 balance. The total flow rate was 100 mL min−1, corresponding to a weight hourly space velocity (WHSV) of 60,000 mL gcat−1 h−1. The initial test temperature was 50 °C. The interval of adjacent temperature points is 10 °C, and the heating rate was 2 °C/min. The residence time for each temperature point was 1 h. The inlet and outlet concentrations of HCHO and CO2 were analyzed simultaneously using an FTIR gas analyzer (Gasmet Instruments DX-4000, Vantaa, Finland). The HCHO conversion was calculated as follows:
HCHO   Conversion   ( % ) = [ CO 2 ] out [ CO 2 ] in [ HCHO ] in × 100 %
where [CO2]in and [CO2]out are the CO2 concentration (ppm) in the inlet and outlet gas, respectively. [HCHO]in represents the HCHO concentration (ppm) in the inlet gas. No other carbon compounds other than CO2 were detected during HCHO catalytic oxidation.

4. Conclusions

In conclusion, a series of Cu-doped Co3O4 catalysts with different Cu/Co molar ratios were constructed by in situ pyrolysis of bimetal Cu-ZIF-67 precursors and utilized for low-temperature HCHO catalytic oxidation. All Cu-doped Co3O4 catalysts showed remarkably improved HCHO oxidation activity in comparison to the pure Co3O4, amongst which the Cu1Co4 stood out (T90 = 108 °C). The excellent HCHO oxidation performance of Cu1Co4 could be related to its adequate defect sites, surface adsorbed oxygen species, higher Co3+/Co2+ ratio, and superb low-temperature reducibility, which derived from its stronger interaction between Cu and Co species (Co2+ + Cu2+ ↔ Co3+ + Cu+). Moreover, in-situ DRIFTS results demonstrated that the dioxymethylene, formate, and carbonate species were primary reaction intermediates of HCHO oxidation over the Cu1Co4 catalyst. We anticipate that the outcomes in the present work might offer an instructive guideline for the rational design of high-performance Co-based catalysts for low-temperature HCHO oxidation.

Author Contributions

Methodology, Q.Z.; validation, Q.Z. and N.X.; writing-original draft preparation, Q.Z.; writing-review and editing, N.X., S.W., H.H. and Q.L.; supervision, N.X.; project administration, N.X. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Basic Research Project of Shanxi Province (20210302124346), Initial Scientific Research Fund for PhD of Changzhi University (XN0359), and Scientific and Technological Innovation Programs of Higher Education Institutions in Shanxi (2021L527 and 2021L523).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Salthammer, T.; Mentese, S.; Marutzky, R. Formaldehyde in the indoor environment. Chem. Rev. 2010, 110, 2536–2572. [Google Scholar] [CrossRef] [PubMed]
  2. Tang, X.; Bai, Y.; Duong, A.; Smith, M.T.; Li, L.; Zhang, L. Formaldehyde in China: Production, consumption, exposure levels, and health effects. Environ. Int. 2009, 35, 1210–1224. [Google Scholar] [CrossRef] [PubMed]
  3. WHO. Air Quality Guidelines for Europe; WHO Regional Office for Europe: Copenhagen, Denmark, 1987. [Google Scholar]
  4. Singapore Ministry of Environment. Guidelines for Indoor Air Quality in Office Premises; Institute of Environmental Epidemiology: Singapore, 1996.
  5. GB/T 18883-2022; Indoor Air Quality Standard. China’s State Administration for Market Regulation: Beijing, China, 2022.
  6. Suresh, S.; Bandosz, T.J. Removal of formaldehyde on carbon-based materials: A review of the recent approaches and findings. Carbon 2018, 137, 207–221. [Google Scholar] [CrossRef]
  7. Chang, M.B.; Lee, C.C. Destruction of formaldehyde with dielectric denier discharge plasmas. Environ. Sci. Technol. 1995, 29, 181–186. [Google Scholar] [CrossRef] [PubMed]
  8. Chen, M.; Wang, H.; Chen, X.; Wang, F.; Qin, X.; Zhang, C.; He, H. High-performance of Cu-TiO2 for photocatalytic oxidation of formaldehyde under visible light and the mechanism study. Chem. Eng. J. 2020, 390, 124481. [Google Scholar] [CrossRef]
  9. Chen, X.; Wang, H.; Chen, M.; Qin, X.; He, H.; Zhang, C. Co-function mechanism of multiple active sites over Ag/TiO2 for formaldehyde oxidation. Appl. Catal. B 2021, 282, 119543. [Google Scholar] [CrossRef]
  10. Wang, C.; Li, Y.; Zhang, C.; Chen, X.; Liu, C.; Weng, W.; Shan, W.; He, H. A simple strategy to improve Pd dispersion and enhance Pd/TiO2 catalytic activity for formaldehyde oxidation: The roles of surface defects. Appl. Catal. B 2021, 282, 119540. [Google Scholar] [CrossRef]
  11. Ye, J.; Yu, Y.; Fan, J.; Cheng, B.; Yu, J.; Ho, W. Room-temperature formaldehyde catalytic decomposition. Environ. Sci. Nano 2020, 7, 3655–3709. [Google Scholar] [CrossRef]
  12. Zhu, D.; Chen, M.; Huang, Y.; Li, R.; Huang, T.; Cao, J.; Shen, Z.; Lee, S. FeCo alloy encased in nitrogen-doped carbon for efficient formaldehyde removal: Preparation, electronic structure, and d-band center tailoring. J. Hazard. Mater. 2022, 424, 127593. [Google Scholar] [CrossRef]
  13. Zhang, C.; Liu, F.; Zhai, Y.; Ariga, H.; Yi, N.; Liu, Y.; Asakura, K.; Flytzani-Stephanopoulos, M.; He, H. Alkali-metal-promoted Pt/TiO2 opens a more efficient pathway to formaldehyde oxidation at ambient temperatures. Angew. Chem. Int. Ed. 2012, 51, 9628–9632. [Google Scholar] [CrossRef]
  14. Xu, Y.; Dhainaut, J.; Dacquin, J.; Mamede, A.; Marinova, M.; Lamonier, J.; Vezin, H.; Zhang, H.; Royer, S. La1−x(Sr, Na, K)xMnO3 perovskites for HCHO oxidation: The role of oxygen species on the catalytic mechanism. Appl. Catal. B 2021, 287, 119955. [Google Scholar] [CrossRef]
  15. Bai, B.; Qiao, Q.; Li, J.; Hao, J. Progress in research on catalysts for catalytic oxidation of formaldehyde. Chin. J. Catal. 2016, 37, 102–122. [Google Scholar] [CrossRef]
  16. Nie, L.; Yu, J.; Jaroniec, M.; Tao, F. Room-temperature catalytic oxidation of formaldehyde on catalysts. Catal. Sci. Technol. 2016, 6, 3649–3669. [Google Scholar] [CrossRef]
  17. Guo, J.; Lin, C.; Jiang, C.; Zhang, P. Review on noble metal-based catalysts for formaldehyde oxidation at room temperature. Appl. Surf. Sci. 2019, 475, 237–255. [Google Scholar] [CrossRef]
  18. Miao, L.; Wang, J.; Zhang, P. Review on manganese dioxide for catalytic oxidation of airborne formaldehyde. Appl. Surf. Sci. 2019, 466, 441–453. [Google Scholar] [CrossRef]
  19. Chen, R.; Sun, Z.; Hardacre, C.; Tang, X.; Liu, Z. The current status of research on the catalytic oxidation of formaldehyde. Catal. Rev. 2022. [Google Scholar] [CrossRef]
  20. Torres, J.; Royer, S.; Bellat, J.; Giraudon, J.; Lamonier, J. Formaldehyde: Catalytic oxidation as a promising soft way of elimination. ChemSusChem 2013, 6, 578–592. [Google Scholar] [CrossRef]
  21. Shi, K.; Wang, L.; Li, L.; Zhao, X.; Chen, Y.; Hua, Z.; Li, X.; Gu, X.; Li, L. Mild preoxidation treatment of Pt/TiO2 catalyst and its enhanced low temperature formaldehyde decomposition. Catalysts 2019, 9, 694. [Google Scholar] [CrossRef] [Green Version]
  22. Li, K.; Ji, J.; He, M.; Huang, H. Complete oxidation of formaldehyde over a Pd/CeO2 catalyst at room temperature: Tunable active oxygen species content by non-thermal plasma activation. Catal. Sci. Technol. 2020, 10, 6257–6265. [Google Scholar] [CrossRef]
  23. Chen, B.; Zhu, X.; Crocker, M.; Wang, Y.; Shi, C. FeOx-supported gold catalysts for catalytic removal of formaldehyde at room temperature. Appl. Catal. B 2014, 154–155, 73–81. [Google Scholar] [CrossRef]
  24. Sun, X.; Lin, J.; Wang, Y.; Li, L.; Pan, X.; Su, Y.; Wang, X. Catalytically active Ir0 species supported on Al2O3 for complete oxidation of formaldehyde at ambient temperature. Appl. Catal. B 2020, 268, 11741. [Google Scholar] [CrossRef]
  25. Ye, J.; Zhu, B.; Cheng, B.; Jiang, C.; Wageh, S.; Al-Ghamdi, A.; Yu, J. Synergy between platinum and gold nanoparticles in oxygen activation for enhanced room-temperature formaldehyde oxidation. Adv. Funct. Mater. 2022, 32, 2110423. [Google Scholar] [CrossRef]
  26. Xiang, N.; Hou, Y.; Han, X.; Li, Y.; Guo, Y.; Liu, Y.; Huang, Z. Promoting effect and mechanism of alkali Na on Pd/SBA-15 for room temperature formaldehyde catalytic oxidation. ChemCatChem 2019, 11, 5098–5107. [Google Scholar] [CrossRef]
  27. Xiang, N.; Han, X.; Bai, Y.; Li, Q.; Zheng, J.; Li, Y.; Hou, Y.; Huang, Z. Size effect of γ-Al2O3 supports on the catalytic performance of Pd/γ-Al2O3 catalysts for HCHO oxidation. Mol. Catal. 2020, 494, 111112. [Google Scholar] [CrossRef]
  28. Xiang, N.; Han, X.; Zheng, J.; Li, Q.; Zhao, Q.; Hou, Y.; Huang, Z. Effect of manganese modification on the low-temperature formaldehyde oxidation performance of ZIF-67 derived Co3O4. J. Fuel Chem. Technol. 2022, 50, 859–867. [Google Scholar]
  29. Xiang, N.; Bai, Y.; Li, Q.; Han, X.; Zheng, J.; Zhao, Q.; Hou, Y.; Huang, Z. ZIF-67-derived hierarchical hollow Co3O4@CoMn2O4 nanocages for efficient catalytic oxidation of formaldehyde at low temperature. Mol. Catal. 2022, 528, 112519. [Google Scholar] [CrossRef]
  30. Zhu, D.; Huang, Y.; Cao, J.; Lee, S.; Chen, M.; Shen, Z. Cobalt nanoparticles encapsulated in porous nitrogen-doped carbon: Oxygen activation and efficient catalytic removal of formaldehyde at room temperature. Appl. Catal. B 2019, 258, 117981. [Google Scholar] [CrossRef]
  31. Li, R.; Huang, Y.; Zhu, D.; Ho, W.; Lee, S.; Cao, J. A review of Co3O4-based catalysts for formaldehyde oxidation at low temperature: Effect parameters and reaction mechanism. Aerosol Sci. Eng. 2020, 4, 147–168. [Google Scholar] [CrossRef]
  32. Fan, Z.; Fang, W.; Zhang, Z.; Chen, M.; Shangguan, W. Highly active rod-like Co3O4 catalyst for the formaldehyde oxidation reaction. Catal. Commun. 2018, 103, 10–14. [Google Scholar] [CrossRef]
  33. Fan, Z.; Zhang, Z.; Fang, W.; Yao, X.; Zou, G.; Shangguan, W. Low-temperature catalytic oxidation of formaldehyde over Co3O4 catalysts prepared using various precipitants. Chin. J. Catal. 2016, 37, 947–954. [Google Scholar] [CrossRef]
  34. Lv, L.; Su, Y.; Liu, X.; Zheng, H.; Wang, X. Synthesis of cellular-like Co3O4 nanocrystals with controlled structural, electronic and catalytic properties. J. Alloys Compd. 2013, 553, 163–166. [Google Scholar] [CrossRef]
  35. Bai, B.; Arandiyan, H.; Li, J. Comparison of the performance for oxidation of formaldehyde on nano-Co3O4, 2D-Co3O4, and 3D-Co3O4 catalysts. Appl. Catal. B 2013, 142–143, 677–683. [Google Scholar] [CrossRef]
  36. Huang, M.; Chen, J.; Tang, H.; Jiao, Y.; Zhang, J.; Wang, G.; Wang, R. Improved oxygen activation over metal-organic-frameworks derived and zinc-modulated Co@NC catalyst for boosting indoor gaseous formaldehyde oxidation at room temperature. J. Colloid Interf. Sci. 2021, 601, 833–842. [Google Scholar] [CrossRef] [PubMed]
  37. Lu, S.; Wang, F.; Chen, C.; Huang, F.; Li, K. Catalytic oxidation of formaldehyde over CeO2-Co3O4 catalysts. J. Rare Earths 2017, 35, 867–874. [Google Scholar] [CrossRef]
  38. Zhou, M.; Cai, L.; Bajdich, M.; Garcia-Melchor, M.; Li, H.; He, J.; Wilcox, J.; Wu, W.; Vojvodic, A.; Zheng, X. Enhancing catalytic CO oxidation over Co3O4 nanowires by substituting Co2+ with Cu2+. ACS Catal. 2015, 5, 4485–4491. [Google Scholar] [CrossRef]
  39. Lei, J.; Wang, S.; Li, J.; Xu, Y.; Li, S. Different effect of Y (Y = Cu, Mn, Fe, Ni) doping on Co3O4 derived from Co-MOF for toluene catalytic destruction. Chem. Eng. Sci. 2022, 251, 117436. [Google Scholar] [CrossRef]
  40. Tu, S.; Chen, Y.; Zhang, X.; Yao, J.; Wu, Y.; Wu, H.; Zhang, J.; Wang, J.; Mu, B.; Li, Z.; et al. Complete catalytic oxidation of formaldehyde at room temperature on MnxCo3−xO4 catalysts derived from metal-organic frameworks. Appl. Catal. A 2021, 611, 117975. [Google Scholar] [CrossRef]
  41. Xu, W.; Chen, X.; Chen, J.; Jia, H. Bimetal oxide CuO/Co3O4 derived from Cu ions partly-substituted framework of ZIF-67 for toluene catalytic oxidation. J. Hazard. Mater. 2021, 403, 123869. [Google Scholar] [CrossRef]
  42. Mu, G.; Zeng, Y.; Zheng, Y.; Cao, Y.; Liu, F.; Liang, S.; Zhan, Y.; Jiang, L. Oxygen vacancy defects engineering on Cu-doped Co3O4 for promoting effective COS hydrolysis. Green Energy Environ. 2021; in press. [Google Scholar] [CrossRef]
  43. Fan, L.; Li, M.; Zhang, C.; Ismail, A.; Hu, B.; Zhu, Y. Effect of Cu/Co ratio in CuaCo1−aOx (a = 0.1, 0.2, 0.4, 0.6) flower structure on its surface properties and catalytic performance for toluene oxidation. J. Colloid Interf. Sci. 2021, 599, 404–415. [Google Scholar] [CrossRef]
  44. Zhang, W.; Descorme, C.; Valverde, J.; Giroir-Fendler, A. Cu-Co mixed oxide catalysts for the total oxidation of toluene and propane. Catal. Today 2022, 384–386, 238–245. [Google Scholar] [CrossRef]
  45. Li, R.; Huang, Y.; Zhu, D.; Ho, W.; Cao, J.; Lee, S. Improved oxygen activation over a carbon/Co3O4 nanocomposite for efficient catalytic oxidation of formaldehyde at room temperature. Environ. Sci. Technol. 2021, 55, 4054–4063. [Google Scholar] [CrossRef]
  46. Huang, Y.; Liu, Y.; Wang, W.; Chen, M.; Li, H.; Lee, S.; Ho, W.; Huang, T.; Cao, J. Oxygen vacancy-engineered δ-MnOx/activated carbon for room temperature catalytic oxidation of formaldehyde. Appl. Catal. B 2020, 278, 119294. [Google Scholar] [CrossRef]
  47. Duan, C.; Yu, Y.; Hu, H. Recent progress on synthesis of ZIF-67-based materials and their application to heterogeneous catalysis. Green Energy Environ. 2022, 7, 3–15. [Google Scholar] [CrossRef]
  48. Wu, R.; Qian, X.; Rui, X.; Liu, H.; Yadian, B.; Zhou, K.; Wei, J.; Yan, Q.; Feng, X.; Long, Y.; et al. Zeolitic imidazolate framework 67-derived high symmetric porous Co3O4 hollow dodecahedra with highly enhanced lithium storage capability. Small 2014, 10, 1932–1938. [Google Scholar] [CrossRef]
  49. Hu, H.; Guan, B.; Xia, B.; Lou, X.W. Designed formation of Co3O4/NiCo2O4 double-shelled nanocages with enhanced pseudocapacitive and electrocatalytic properties. J. Am. Chem. Soc. 2015, 137, 5590–5595. [Google Scholar] [CrossRef]
  50. Hu, H.; Guan, B.; Lou, X.W. Construction of complex CoS hollow structures with enhanced electrochemical properties for hybrid supercapacitors. Chem 2016, 1, 102–113. [Google Scholar] [CrossRef] [Green Version]
  51. Hu, Z.; Chen, J.; Yan, D.; Li, Y.; Jia, H.; Lu, C. Enhanced catalytic activities of MnOx/Co3O4 nanocomposites prepared via MOFs-templated approach for chlorobenzene oxidation. Appl. Surf. Sci. 2021, 551, 149453. [Google Scholar] [CrossRef]
  52. Daté, M.; Okumura, M.; Tsubota, S.; Haruta, M. Vital role of moisture in the catalytic activity of supported gold nanoparticles. Angew. Chem. 2004, 116, 2181–2184. [Google Scholar] [CrossRef]
  53. Ma, Y.; Wang, L.; Ma, J.; Wang, H.; Zhang, C.; Deng, H.; He, H. Investigation into the enhanced catalytic oxidation of o-xylene over MOF-derived Co3O4 with different shapes: The role of surface twofold-coordinate lattice oxygen (O2f). ACS Catal. 2021, 11, 6614–6625. [Google Scholar] [CrossRef]
  54. Aadil, M.; Nazik, G.; Zulfiqar, S.; Shakir, I.; Aboud, M.; Agboola, P.; Haider, S.; Warsi, M. Fabrication of nickel foam supported Cu-doped Co3O4 nanostructures for electrochemical energy storage applications. Ceram. Int. 2021, 47, 9225–9233. [Google Scholar] [CrossRef]
  55. Wang, Z.; Wang, W.; Zhang, L.; Jiang, D. Surface oxygen vacancies on Co3O4 mediated catalytic formaldehyde oxidation at room temperature. Catal. Sci. Technol. 2016, 6, 3845–3853. [Google Scholar] [CrossRef]
  56. Zhang, Y.; Zhou, X.; Zhang, F.; Tian, T.; Ding, Y.; Gao, H. Design and synthesis of Cu modified cobalt oxides with hollow polyhedral nanocages as efficient electrocatalytic and photocatalytic water oxidation catalysts. J. Catal. 2017, 352, 246–255. [Google Scholar] [CrossRef]
  57. Zhang, Y.; Zeng, Z.; Li, Y.; Hou, Y.; Hu, J.; Huang, Z. Effect of the A-site cation over spinel AMn2O4 (A = Cu2+, Ni2+, Zn2+) for toluene combustion: Enhancement of the synergy and the oxygen activation ability. Fuel 2021, 288, 119700. [Google Scholar] [CrossRef]
  58. Hernández, W.; Centeno, M.; Ivanova, S.; Eloy, P.; Gaigneaux, E.; Odriozola, J. Cu-modified cryptomelane oxide as active catalyst for CO oxidation reactions. Appl. Catal. B 2012, 123–124, 27–35. [Google Scholar] [CrossRef]
  59. Wang, J.; Zhang, G.; Zhang, P. Layered birnessite-type MnO2 with surface pits for enhanced catalytic formaldehyde oxidation activity. J. Mater. Chem. A 2017, 5, 5719–5725. [Google Scholar] [CrossRef]
  60. Zhao, J.; Tang, Z.; Dong, F.; Zhang, J. Controlled porous hollow Co3O4 polyhedral nanocages derived from metal-organic frameworks (MOFs) for toluene catalytic oxidation. Mol. Catal. 2019, 463, 77–86. [Google Scholar] [CrossRef]
  61. Fierro, G.; Jacono, M.; Inversi, M.; Dragone, R.; Porta, P. TPR and XPS study of cobalt-copper mixed oxide catalysts: Evidence of a strong Co-Cu interaction. Top. Catal. 2000, 10, 39–48. [Google Scholar] [CrossRef]
  62. Chen, D.; Qu, Z.; Sun, Y.; Gao, K.; Wang, Y. Identification of reaction intermediates and mechanism responsible for highly active HCHO oxidation on Ag/MCM-41 catalysts. Appl. Catal. B 2013, 142–143, 838–848. [Google Scholar] [CrossRef]
  63. Zhang, C.; He, H.; Tanaka, K.-I. Catalytic performance and mechanism of a Pt/TiO2 catalyst for the oxidation of formaldehyde at room temperature. Appl. Catal. B 2006, 65, 37–43. [Google Scholar] [CrossRef]
  64. Sun, X.; Lin, J.; Guan, H.; Li, L.; Sun, L.; Wang, Y.; Miao, S.; Su, Y.; Wang, X. Complete oxidation of formaldehyde over TiO2 supported subnanometer Rh catalyst at ambient temperature. Appl. Catal. B 2018, 226, 575–584. [Google Scholar] [CrossRef]
  65. Wang, J.; Li, J.; Jiang, C.; Zhou, P.; Zhang, P.; Yu, J. The effect of manganese vacancy in birnessite-type MnO2 on room-temperature oxidation of formaldehyde in air. Appl. Catal. B 2017, 204, 147–155. [Google Scholar] [CrossRef]
  66. Yang, T.; Huo, Y.; Liu, Y.; Rui, Z.; Ji, H. Efficient formaldehyde oxidation over nickel hydroxide promoted Pt/γ-Al2O3 with a low Pt content. Appl. Catal. B 2017, 200, 5. [Google Scholar] [CrossRef]
Figure 1. Ignition curves of Co3O4, Cu1Co8, Cu1Co4 and Cu1Co2 for HCHO oxidation (WHSV = 60,000 mL gcat−1 h−1, 80 ppm HCHO, 60% relative humidity, 20 vol% O2, and N2 balance).
Figure 1. Ignition curves of Co3O4, Cu1Co8, Cu1Co4 and Cu1Co2 for HCHO oxidation (WHSV = 60,000 mL gcat−1 h−1, 80 ppm HCHO, 60% relative humidity, 20 vol% O2, and N2 balance).
Catalysts 13 00117 g001
Figure 2. (a) Effect of relative humidity on the HCHO oxidation activity of Cu1Co4; (b) Stability test of Cu1Co4 for HCHO oxidation.
Figure 2. (a) Effect of relative humidity on the HCHO oxidation activity of Cu1Co4; (b) Stability test of Cu1Co4 for HCHO oxidation.
Catalysts 13 00117 g002
Figure 3. (a,b) XRD patterns of Co3O4, Cu1Co8, Cu1Co4, and Cu1Co2.
Figure 3. (a,b) XRD patterns of Co3O4, Cu1Co8, Cu1Co4, and Cu1Co2.
Catalysts 13 00117 g003
Figure 4. (a) Raman and (b) FT-IR spectra of Co3O4, Cu1Co8, Cu1Co4, and Cu1Co2.
Figure 4. (a) Raman and (b) FT-IR spectra of Co3O4, Cu1Co8, Cu1Co4, and Cu1Co2.
Catalysts 13 00117 g004
Figure 5. SEM images of (a) Co3O4, (b) Cu1Co8, (c) Cu1Co4 and (d) Cu1Co2. (e) EDS-mapping of Co, Cu, and O in Cu1Co4.
Figure 5. SEM images of (a) Co3O4, (b) Cu1Co8, (c) Cu1Co4 and (d) Cu1Co2. (e) EDS-mapping of Co, Cu, and O in Cu1Co4.
Catalysts 13 00117 g005
Figure 6. (a) TEM and (b,c) HRTEM images of Cu1Co4.
Figure 6. (a) TEM and (b,c) HRTEM images of Cu1Co4.
Catalysts 13 00117 g006
Figure 7. (a) N2 adsorption-desorption isotherms and (b) corresponding pore size distribution curves of Co3O4, Cu1Co8, Cu1Co4, and Cu1Co2.
Figure 7. (a) N2 adsorption-desorption isotherms and (b) corresponding pore size distribution curves of Co3O4, Cu1Co8, Cu1Co4, and Cu1Co2.
Catalysts 13 00117 g007
Figure 8. XPS spectra of Co3O4, Cu1Co8, Cu1Co4 and Cu1Co2. (a) Co 2p, (b) Cu 2p3/2, and (c) O 1s.
Figure 8. XPS spectra of Co3O4, Cu1Co8, Cu1Co4 and Cu1Co2. (a) Co 2p, (b) Cu 2p3/2, and (c) O 1s.
Catalysts 13 00117 g008
Figure 9. H2-TPR profiles of Co3O4, Cu1Co8, Cu1Co4 and Cu1Co2.
Figure 9. H2-TPR profiles of Co3O4, Cu1Co8, Cu1Co4 and Cu1Co2.
Catalysts 13 00117 g009
Figure 10. In-situ DRIFTS of HCHO oxidation on the Cu1Co4 catalyst.
Figure 10. In-situ DRIFTS of HCHO oxidation on the Cu1Co4 catalyst.
Catalysts 13 00117 g010
Table 1. Ignition temperatures of Co3O4, Cu1Co8, Cu1Co4, and Cu1Co2 for HCHO oxidation.
Table 1. Ignition temperatures of Co3O4, Cu1Co8, Cu1Co4, and Cu1Co2 for HCHO oxidation.
CatalystT50 (°C)T90 (°C)
Co3O4147157
Cu1Co8103122
Cu1Co498108
Cu1Co2101112
Table 2. Crystallite size and textural properties of Co3O4, Cu1Co8, Cu1Co4, and Cu1Co2.
Table 2. Crystallite size and textural properties of Co3O4, Cu1Co8, Cu1Co4, and Cu1Co2.
CatalystCrystallite Size (nm)SBET (m2/g)Vp (cm3/g)Average Pore Size (nm)
Co3O418430.2220
Cu1Co815370.1820
Cu1Co414440.2018
Cu1Co217320.1519
Table 3. Surface element properties of Co3O4, Cu1Co8, Cu1Co4, and Cu1Co2.
Table 3. Surface element properties of Co3O4, Cu1Co8, Cu1Co4, and Cu1Co2.
CatalystCo3+/Co2+Cu+/Cu2+Oα/Oβ
Co3O40.53-0.33
Cu1Co80.590.160.51
Cu1Co40.780.290.55
Cu1Co20.690.210.53
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhao, Q.; Xiang, N.; Wen, S.; Huo, H.; Li, Q. ZIF-67 Derived Cu-Co Mixed Oxides for Efficient Catalytic Oxidation of Formaldehyde at Low-Temperature. Catalysts 2023, 13, 117. https://doi.org/10.3390/catal13010117

AMA Style

Zhao Q, Xiang N, Wen S, Huo H, Li Q. ZIF-67 Derived Cu-Co Mixed Oxides for Efficient Catalytic Oxidation of Formaldehyde at Low-Temperature. Catalysts. 2023; 13(1):117. https://doi.org/10.3390/catal13010117

Chicago/Turabian Style

Zhao, Qingsong, Ning Xiang, Shiting Wen, Haibo Huo, and Qiaoyan Li. 2023. "ZIF-67 Derived Cu-Co Mixed Oxides for Efficient Catalytic Oxidation of Formaldehyde at Low-Temperature" Catalysts 13, no. 1: 117. https://doi.org/10.3390/catal13010117

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