Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Art 7

Download as pdf or txt
Download as pdf or txt
You are on page 1of 23

Chemical Engineering Journal 429 (2022) 131579

Contents lists available at ScienceDirect

Chemical Engineering Journal


journal homepage: www.elsevier.com/locate/cej

Activity, selectivity, and stability of earth-abundant CuO/Cu2O/Cu0-based


photocatalysts toward CO2 reduction
Shahzad Ali a, b, Abdul Razzaq b, Hwapyong Kim a, Su-Il In a, *
a
Department of Energy Science & Engineering, DGIST, 333 Techno Jungangdaero, Hyeonpung-eup, Dalseong-gun, Daegu 42988, Republic of Korea
b
Department of Chemical Engineering, COMSATS University Islamabad, Lahore Campus, 1.5 KM Defense Road, Off Raiwind Road, Lahore 54000, Pakistan

A R T I C L E I N F O A B S T R A C T

Keywords: Cu-based photocatalysts are widely used for the photocatalytic reduction of CO2 owing to their non-toxicity,
CO2 photoreduction earth abundance, extended light absorption, suppressed charge recombination, and good catalytic perfor­
Cu-based photocatalyst mance. In addition to low cost, abundant availability, and the ease of synthesis, they exhibit unique features,
Photocatalytic activity
such as broad optical absorption, reaction intermediates stabilization, and C–C coupling ability, which lead to the
Selectivity
Stability
formation of C2 + products. The photocatalytic activity of Cu-based photocatalysts is essentially linked to the
Plasmonic photocatalysts optical absorption and interfacial charge transfer at the junction of Cu and the semiconductor substrate. How­
Heterostructured photocatalyst ever, the poor resistance of Cu to oxidation seriously perturbs the effective utilization of its unique features in
Interfacial charge transfer practical applications. To date, various approaches, such as the use of metal/non-metal co-catalysts, Z-scheme
C2+ products heterostructures, and hole scavengers, have been proposed to improve its photocatalytic performance and
maintain its stability in prolonged reactions. In addition to these approaches, as single metal atom catalysts,
atomically dispersed Cu photocatalysts have gained immense attention because they can regain the desired
oxidation state, and hence exhibit good stability. The designation of suitable oxidation states of Cu for various
CO2 reduction reaction steps is challenging because of the rapid change in its oxidation states under ambient/
irradiated environments. However, in situ spectroscopic analyses have nominated Cu2+ for CO2 adsorption, Cu1+
for the photoreduction reaction, and Cu0 for effective charge separation. In this review, the recent advancements
in the photocatalytic activity, selectivity, and stability of Cu-based photocatalysts are discussed systematically.
Certain concepts and mechanisms related to the photocatalytic performance of Cu-based catalysts have also been
discussed. Finally, the future research directions are discussed based on the available relevant literature.

1. Introduction approaches that are more sustainable and less catastrophic to nature
[3–4]. In the first scenario, a globalized and mega-plantation drive is
Excessive burning of fossil fuels has led to the escalation of CO2 required, but it may take several years for these plants to mature.
concentration in the environment, reflecting the destructive effects of However, the second scenario seems promising as a synthetic process,
climate change, such as temperature and sea level rise [1–2]. If these once developed, can be upscaled in a short time span [5–6]. In this
emissions are left unabated, the environment-related issues will become context, photocatalytic CO2 reduction is a replica of natural photosyn­
more serious. As a remedial measure, we should either increase the thesis and can convert CO2 to solar fuels, such as CH4 and other value-
magnitude of natural photosynthesis to out-pace the concentration of added chemicals under sunlight at ambient conditions [7]. Although it
perilous CO2 or opt for alternative strategies, such as the use of synthetic produces fuels that are compatible with the present utilities,

Abbreviations: RT, reduced titania; Vo, oxygen vacancies; IFCT, interfacial charge transfer; STF, solar to fuel; NP, nanoparticles; UV, ultraviolet; IR, infrared; NIR,
near-infrared; LSPR, localized surface plasmon resonance; CB, conduction band; VB, valence band; PMET, plasmon-mediated electron transfer; PICTT, plasmon-
induced interfacial charge transfer transition; XAS, X-ray absorption spectroscopy; XPS, X-ray photoelectron spectroscopy; PL, photoluminescence; FDTD, finite
difference time domain; DFT, density functional theory; NS, nanosphere; SAC, single atom catalyst; QD, quantum dot; EPR, electron paramagnetic resonance; DRIFT,
diffuse reflectance infrared Fourier transform; TPR, temperature programmed reduction; XANES, X-ray absorption near edge structure; EXAFS, extended X-ray-
absorption fine-structure; TPD, temperature programmed desorption; TGA, thermal gravimetric analysis; TNTAs, titania-nanotube arrays; FT-IR, Fourier transform
infrared; IB, intermediate band; NHE, normal hydrogen electrode.
* Corresponding author.
E-mail address: insuil@dgist.ac.kr (S.-I. In).

https://doi.org/10.1016/j.cej.2021.131579

Available online 10 August 2021


1385-8947/© 2021 Elsevier B.V. All rights reserved.
S. Ali et al. Chemical Engineering Journal 429 (2022) 131579

commercially promising photocatalysts yet to be synthesized [8–10]. [42–43]. Therefore, various efforts are being made to achieve atomic-
Photocatalytic CO2 reduction has been extensively investigated level dispersion for reversible redox reactions [16].
during the past decade and has become an interesting research domain Various comprehensive reviews are available on the reduction of
[11–12]. To overcome the inherent challenges of the photocatalytic CO2 CO2 via photocatalysis/photothermal catalysis using H2 or H2O as a
reduction, such as limited light absorption, rapid charge-recombination, reducing agent [44]. In particular, certain reviews have been published
low affinity/adsorption for reactants, and photocorrosion, various for Cu-abundant photocatalysts. For example, Varma et al. wrote a
modification techniques such as bandgap engineering and dye sensiti­ comprehensive review on the synthesis of Cu-based materials and their
zation have been proposed [13]. Additionally, novel materials with catalytic applications [25]. Li et al. reviewed the progress in the syn­
unique features have been developed, such as MXenes and single metal thesis and applications of one-dimensional Cu-based photocatalysts for
atom catalysts (SACs) [14–16]. Despite these efforts, the development of CO2 reduction [45]. Xie et al. and Fornasiero et al. discussed the
an economically viable photocatalytic CO2 reduction system has not response of Cu, Cu alloys, and CuS/Cu2S for the photocatalytic reduction
been realized because of the adverse impacts of the aforementioned is­ of CO2 [30,46]. All these reviews idealize Cu-based photocatalysts with
sues on the activity, selectivity, and stability of conventional photo­ a prime focus on their low cost. In this review, we have summarized the
catalysts [17]. Hence, for the practical-scale implementation of this effect of the Cu oxidation state on the activity, stability, selectivity, and
technology, cost-effective materials with scalable activity and stability C2 + product forming ability of Cu-based photocatalysts, thus demon­
are highly desirable. strating their bifold significance. Cu passivation by photocorrosion has
Earth-abundant transition metals/metal oxides have been exten­ also been discussed in detail along with various Cu stabilization tech­
sively explored in the quest to obtain higher yields of solar fuels. In this niques. Finally, considering the efficacious role of Cu in the photo­
regard, considerable progress has been made, as some Cu-based photo­ catalytic reduction of CO2, the future perspectives of Cu-based
catalysts yield solar products at reasonable production rates (up to photocatalysts are discussed.
certain millimoles per hour) [18–20]. However, such photocatalysts
show a solar to fuel (STF) efficiency of less than 1.0% for extended re­ 2. Photocatalytic activity of Cu-based photocatalysts
action times, which is not sufficient for commercial-scale applications.
Therefore, to improve the compatibility of these photocatalysts with The activity of Cu-based photocatalysts depends on various param­
solar photovoltaics, either the STF efficiency should be increased or their eters, such as the composition of the photocatalyst (Cu is used alone or in
selectivity should be diverted toward C2 + products, particularly those tandem with other metal oxide photocatalysts), size of the Cu particles,
with higher market value (i.e., ethane and propane), for economic gains optoelectronic properties, form of Cu incorporation on/in other photo­
[21–22]. Cu-based photocatalysts steers the selectivity toward catalysts, that is, doped, deposited, or finely dispersed, oxidation state of
C2 + products through C–C coupling, and Cu controls this reaction [23]. Cu, and type of the reaction system [20,47–49].
In this regard, a suitable oxidation state of Cu is essential to stabilize the
reaction intermediates and couple them to form the final C2 + products 2.1. Extended optical response triggered activity
[24]. Therefore, Cu-based photocatalysts with suitable Cu oxidation
states are highly desirable owing to their fascinating properties. It is well-established that the terrestrial solar spectrum is roughly
Cu is a 3d transition metal exhibiting variable oxidation states, and composed of 4% ultraviolet (UV) light, 43% visible light, and 53%
hence, it can interact with CO2 and H2O [25–27]. It can interact with infrared (IR) light [50]. Therefore, with the development of visible-light-
these reactants through either one- or two-electron pathways depending active photocatalysts, the interest of researchers has also shifted to
upon its oxidation state to produce adsorbates with different configu­ capturing the IR light (which accounts for a major portion of the solar
rations and adsorption strengths [19]. These configurations affect the spectrum) [51]. The basic strategies to achieve a visible-near infrared
selectivity of the adsorbates to various adsorbents. This is because under (NIR) light-active photocatalyst include: (i) the development of plas­
illumination, these adsorbates are transformed into various products via monic photocatalysts, hence harvesting the benefits of matching the
multiple reaction mechanisms [28–29]. Because of these unique char­ frequency of incident light with the oscillation of plasmonic material
acteristics, Cu-based nanocatalysts are used in a variety of catalytic re­ electrons, and (ii) the development of visible-NIR active photocatalysts
actions [25,30–31]. However, under illumination, controlling the mainly by: (a) the creation of an intermediate band via defect generation
oxidation states of Cu2O/CuO is challenging [32]. This is because the between the CB and VB of semiconductor photocatalysts, and (b)
irreversible redox disproportionation reaction of Cu2O designing narrow bandgap semiconductors [52].
(Cu2O → Cu + CuO) reduces its stability [33]. Therefore, although Cu- Plasmonic photocatalysts employ noble metals, such as Au, Ag, and
based photocatalysts exhibit high photoactivity, earth abundance, and Pt [53]. However, Cu or Cu-based materials (sulfides and selenides) also
non-toxicity, their stability should be improved in addition to selectivity offer excellent plasmonic behavior, making earth metals a cheap and
to realize their commercial applications [34–35]. abundant substitute [54–56]. Plasmonic response, termed as local sur­
The stability of Cu can be improved using various methods, such as face plasmon resonance (LSPR), arises from the frequency matching of
passivating the surface of Cu nanoparticles (NPs) with stable oxides like light irradiation and oscillation of free surface electrons, resulting in the
Al2O3 and fabricating core–shell structures with oxidation-resistant generation of free electrons (hot electrons) and holes. These excited hot
metals like Pt [36–37]. However, these approaches reduce the effec­ electrons are then consumed following one of the two pathways, that is,
tive utilization of the surface area of Cu for harvesting photons and plasmon-mediated electron transfer (PMET) or plasmon-induced inter­
executing the reactions [38]. Further Z-scheme charge transfer path­ facial charge transfer transition (PICTT) [57–58]. The former pathway is
ways have also been used to maintain the stability by constructing a a two-step process, which involves the generation of hot electrons by
heterostructure of Cu2O with a suitable metal oxide. In such hetero­ LSPR decay, followed by the transfer of these hot electrons to the CB of
structures, the weak oxidizing holes in Cu2O are quenched by the pho­ the coupled photocatalyst (semiconductor). The PICTT pathway in­
togenerated electrons from the metal oxides [39–40]. Although this volves single-step direct excitation and the injection of hot electrons
pathway avoids Cu oxidation but it reduces Cu to elemental form and from the plasmonic metal to the energy state of the coupled semi­
restricts its spontaneous reversion to the oxidized state [41]. However, conductor. PICTT requires strong and firm interaction/mixing between
small Cu NPs easily oxidize and reduce in ambient conditions under the the orbitals of the plasmonic metal and semiconductor, and is more
action of air and low-pressure hydrogen, respectively [41]. Recently, it efficient than PMET, an indirect pathway.
has been demonstrated that the interaction of the photocatalyst support Recently, Mahmoud et al. reported a facile synthesis procedure for
with atomically dispersed Cu atoms leads to the spontaneous reduction the in situ growth of Cu NPs on ordered octahedrons of Cu2O [59]. The
and oxidation of these atoms under light and air exposure, respectively Cu NPs/Cu2O octahedrons synthesized by the electroless reduction

2
S. Ali et al. Chemical Engineering Journal 429 (2022) 131579

approach exhibited extended optical absorption up to the NIR region, under ambient conditions, which is conventionally irreversible, whereas
which is mainly attributed to the LSPR effect of the Cu NPs acting as an DFT calculations investigate the electronic states at the interface of Cu
optical antenna. However, the characteristic peak of Cu LSPR was and TiO2. The binary Cu/TiO2 composite transformed into the ternary
damped with the emergence of broad absorbance, thus indicating the Cu/Cu2O/TiO2 composite because of the oxidation under ambient
mixing of the Cu and Cu2O electronic levels. This conceptual view was conditions. The oxidation of Cu occurs on a surface that is in direct
established based on the photoluminescence (PL) spectroscopy results. contact with the oxygen environment. Hence, this ternary composite
The prepared Cu NPs/Cu2O octahedron photocatalyst employed for the (Cu/Cu2O/TiO2) showed three different configurations of the Cu core
photoreduction of CO2 yielded large amounts of CO and CH4. The with a Cu2O covering and TiO2: (i) loaded, (ii) embedded, and (iii) in
enhanced CO2 photoreduction performance of the composite was core–shell configuration (Cu/Cu2O as the core and TiO2 as the shell).
attributed to the LSPR effect and hot electron transfer by the PICTT The improvement in the photocatalytic performance of the catalysts
pathway because of the strong interface junction between the Cu NPs depended on the LSPR effects of the Cu core, which generated additional
and Cu2O octahedrons induced by their facile electroless reduction weak absorption peaks, an interfacial electric field, and hot electrons (at
approach. the interface). These factors depended on the thickness of the Cu2O
In the development of Vis-NIR active photocatalysts by creating in­ covering layer, TiO2 nanosphere (NS) size, and Cu/Cu2O/TiO2 NS size.
termediate band defects such as reduced/black TiO2, the optical ab­ As shown in (Fig. 1a), the LSPR response of all the configurations
sorption peak shows significant broadening and red-shift toward the NIR appeared at approximately 648 nm; however, the absorption depended
region [2,60]. However, this approach suffers from low charge carrier on the configuration type and TiO2 NS size. The maximum red shift in
density and slow mobility, which affect the reaction kinetics. Another the LSPR peak was observed for the core–shell Cu/Cu2O/TiO2 configu­
strategy under the respective domain is to employ a near-zero bandgap ration, whereas the maximum intensity was observed for the embedded
semiconductor with captivating aspects of extraordinarily high charge configuration. Moreover, an increase in the size of the TiO2 NS led to an
carrier density and electrical conductivity. All these approaches for the increase in the LSPR red shift and the absorption intensity. Furthermore,
development of plasmonic photocatalysts and visible-NIR active pho­ an interfacial electric field was induced by the LSPR effect within the
tocatalysts either by incorporating intermediate band defects or using Cu/Cu2O/TiO2 ternary composite. In the case of the core–shell config­
zero bandgap materials must fulfill the important requisite of their band uration, the increase in the strength of the electric field decreased with
edge alignment with the redox potentials of the targeted reactant/ an increase in the TiO2 NS radius. However, for the other two configu­
product [53]. rations (loading and embedding), the strength of the interfacial electric
The LSPR in Cu-based photocatalysts is an interesting research topic. field increased with an increase in the TiO2 NS size. For the core–shell
Specifically, the mechanism underlying the LSPR-induced transition and configuration, the thickness of the Cu2O covering and the TiO2 NS
its impact on the photocatalytic performance of Cu-based photocatalysts affected the LSPR of the core Cu NS, thus limiting the strength
remains an interesting domain for investigation [61]. Yao et al. pro­ enhancement of the interfacial electric field. In contrast, in the other two
posed a credible mechanistic view of Cu/TiO2 photocatalysts based on configurations, a single covering of Cu2O on the core Cu NS did not
finite difference time domain (FDTD) and density function theory (DFT) affect the LSPR effect and interfacial electric field. Furthermore, the hot
calculations. The FDTD method is linked to the oxidation of core Cu electrons generated by the LSPR of the Cu core transferred to the TiO2

Fig. 1. (a) Extinction cross-section of the three proposed configurations, that is, loading, embedding, and core–shell for the Cu/Cu2O/TiO2 ternary composite and,
(b–d) proposed mechanisms for photoexcited charge transfer within the ternary Cu/Cu2O/TiO2 composite, Reproduced from [62] with permission from the Royal
Society of Chemistry.

3
S. Ali et al. Chemical Engineering Journal 429 (2022) 131579

NS via “hot spots” generated at the interface of Cu and Cu2O. It is well plasmon states. The hot electrons could overcome the junction barrier at
known that hot electrons are usually generated by the intraband or the Cu/Cu2O interface and were injected into the Cu2O CB, from where
interband transitions within the Cu core. Based on the theoretical they could easily flow into the TiO2 CB. For the Cu/TiO2/Cu2O config­
models, calculations, and experimental results reported in the literature, uration (Fig. 1c), the Schottky barrier of Cu/TiO2 was lower than that of
the authors proposed different mechanisms that might occur during the Cu/Cu2O, as indicated by the DFT calculations. Thus, under Vis-NIR
photocatalytic reactions, as shown in Fig. 1(b–d). The mechanism light irradiation, the hot electrons generated within the Cu core were
employing the Cu/Cu2O/TiO2 configuration with the energy band favorably transferred to the TiO2 CB. However, a few electrons possibly
structure of Cu/Cu2O is denoted as Cu LSPR-1 (Fig. 1b). In this config­ overcame the Schottky barrier of Cu/Cu2O and moved toward the Cu2O
uration, hot electrons were generated under Vis-NIR light irradiation CB. In this configuration, TiO2 and Cu2O alone could also generate
(the green line represents the intraband transition and the blue line exciton pairs by absorbing the UV and visible light, respectively. Nor­
represents the interband transition). These hot electrons occupied the mally (specifically for the embedded configuration) Cu is in contact with

Fig. 2. (a) UV–Vis DRS of atomic layer and bulk CuS, (b) production of CO and O2 by the photocatalytic reduction of CO2 and, (c) proposed charge transition process
within the conductor photocatalysts with band edges alignment according to the CO2 redox potential [64] Copyright 2019, American Chemical Society, (d) proposed
charge transfer mechanism for the Cu2-xS/g-C3N4 photocatalyst [61] Copyright 2020, John Wiley and Sons.

4
S. Ali et al. Chemical Engineering Journal 429 (2022) 131579

both Cu2O and the TiO2 NS, and the band structure is more similar to the portion of the visible light region. In contrast, the CSCN (0.4) composite
Z-scheme (termed the Cu Z-scheme), as shown in (Fig. 1d). Under sample absorbed the complete region of the visible and IR light, but with
UV–Vis light irradiation, exciton pairs were generated in both the Cu2O a slightly lower absorbance intensity. It is well-known that Cu2-xS forms
covering and TiO2 NS, and the electrons from the TiO2 CB easily an intermediate band (IB) within the bandgap, which is primarily
transferred to the CB of the Cu core as the TiO2 CB was more negative. In induced by Cu defects. Hence, under NIR light irradiation, electrons can
contrast, the holes accumulated in the VB of the Cu2O cover jumped to be excited from the VB to the IB and then from the IB to the CB.
the Cu core VB, where it recombined with the electrons coming from the Moreover, the formation of the S-C bond at the interface between Cu2-xS
TiO2 NS CB, thus completing the Z-scheme mechanism. The holes in the and g-C3N4, as confirmed by the high-resolution X-ray photoelectron
TiO2 VB and the electrons in the Cu2O CB remained alive and were spectroscopy (HR-XPS) and PL spectroscopy results, clearly indicates the
available for the respective photoreactions. On the basis of the mecha­ occurrence of efficient charge separation from g-C3N4 to Cu2-xS. The
nistic concepts and proposals, it can be stated that the main factors proposed mechanism indicating the light absorption regions with
affecting the photocatalytic performance of Cu and TiO2 composites are interfacial charge transfer and the photoreduction of CO2 to the
as follows: (i) the composite synthesis process leading to the formation respective products is shown in Fig. 2d.
of any one of the loading, embedding, or core–shell configurations, (ii)
the LSPR effect of the Cu core promoting the hot electrons and interfacial 2.2. Cu as a co-catalyst
electric field, (iii) the thickness of the Cu2O covering and size of the TiO2
NS, and (iv) the interface engineering leading to the occurrence of Cu When used as a co-catalyst with metal oxides, Cu either provides
LSPR-1, Cu LSPR-2, or Cu Z-scheme mechanisms [62]. photogenerated charge-trapping sites or facilitates the construction of
Copper sulfide (CuS) is a well-known narrow bandgap Cu-based type-II/Z-scheme heterostructures [66–67]. Owing to its lower Fermi
photocatalyst, which exhibits fascinating properties and intriguing IR level than those of widely reported photocatalysts, Cu traps the photo­
active photocatalytic response [63]. However, such behavior is observed generated charges, making them available for the photocatalytic
at specific layer thicknesses and/or with nanoarchitecture designs. The reduction of CO2. The early studies on Cu-based photocatalysts pri­
CuS photocatalysts so-designed show significant potential as conductive marily covered the domain of Cu-doped photocatalysts, such as Cu-
photocatalysts with semi-metallic/metallic nature. doped TiO2 (with optimized amounts of Cu as the dopant) [68–70].
Recently, Li et al. synthesized ultrathin layers of CuS and investi­ However, excessive metal loading always results in the agglomeration of
gated their IR light-active response for the photocatalytic reduction of particles, consequently reducing the active surface area of the photo­
CO2, mainly to CO. It was observed that both the bulk CuS and atomic catalysts. Such an observation was also reported by Koci et al.; they
layer CuS exhibited intense light absorption starting from the visible found that the surface area of the photocatalyst decreased by 1.6 times
light to the IR region, but when utilized for the photocatalytic reduction when the loading of CuO was doubled [71]. In addition to blocking the
of CO2, only the CuS atomic layers generated CO and O2 (from H2O pores, these agglomerates also act as charge recombination centers.
oxidation). This suggests that bulk CuS is inactive because of its inability Further, the optimum size is also essential for achieving enhanced yield,
to excite interband electrons under Vis-IR light absorption. Under Vis-IR which strongly depends on the dispersion of the metal over the photo­
light irradiation, the CuS atomic layers exhibit both intraband and catalyst domain/support [72–73]. Gray et al. found that instead of
interband charge transitions, thus photoreducing CO2 to CO along with adding bulk CuO onto TiO2, the addition of a dispersion of a small
the oxidation of H2O to O2. The optical absorption spectra and CO2 amount of Cu (0.1 wt%) significantly enhances its photocatalytic ac­
reduction performance of the bulk CuS and atomic layer CuS are shown tivity [74]. This indicates that the small size and uniform distribution of
in Fig. 2a and 2b, respectively. In general, such IR light-triggering of the the metal on the substrate is crucial for improving the activity of a
photocatalytic reactions induced by electron-hole pair generation is photocatalyst. Furthermore, the loaded metal/metal oxide over the
believed to be due to one of the following transition processes: (i) in­ surface of the photocatalysts should allow the light to reach the surface
dividual interband transition process or (ii) successive intraband and of the photocatalyst, which is only possible when an optimum amount of
interband transitions. A schematic of the transition processes is shown in the fine metal/metal oxide is uniformly distributed. Ultra-small clusters
Fig. 2c. The terms B-1 and B1 represent the fully occupied band and the or atoms develop atomic-level linkages between Cu and the support (Ti-
lowest unoccupied band, respectively, for the conductor photocatalysts. O-Cu), which enhance the charge transfer among them [72,74]. This
During the individual transition process, the IR light triggers the elec­ section describes only the criterion for loading co-catalysts over metal
trons to B1 or near the Fermi level just above the CB of the conductor- oxide supports for preparing photocatalysts. The effect of these co-
type photocatalyst (scheme I and II). However, for the successive tran­ catalysts on the activity and selectivity of the resulting photocatalyst
sition processes, the IR light induces the excitation of electrons from B-1 is discussed in detail in the following section.
to approximately at the Fermi level in the CB (Scheme III), following the
transition from the CB to B1 (Scheme IV). Once the electrons are excited 2.3. Copper oxide itself as a photocatalyst
to B1 via interband transition, more electrons from B-1 are triggered by
the IR light near the Fermi level. For both the transition processes, an Copper oxides are narrow-bandgap p-type semiconductors (CuO,
important and critical criterion is the good matching of the band edge Eg = 1.35 to 1.7 eV; Cu2O, Eg = 1.9 to 2.2 eV) possessing excellent
positions with respect to the redox potentials of the desired photo­ optoelectronic properties imparted by Cu [67], and hence show signif­
catalytic conversion reactants/products. Hence, copper-based materials, icant potential for application in the photoreduction of CO2. However,
specifically sulfides, in addition to cheap and non-toxic rare earth metal- the VB of Cu2O (0.81 eV vs. NHE) is not favorable for steering the
based photocatalysts, offer an excellent opportunity for designing IR oxidation of H2O (1.23 eV vs. NHE) to H+ and O2 [75–76]. For this
light-active photocatalysts [64]. reason, it is not commonly used alone as a photocatalyst, but is generally
In a recent study, Jiang et al. reported Cu2-xS coupled with graphitic modified and/or coupled with other metal oxides with a strong VB
carbon nitride (g-C3N4), demonstrating the synergetic effect of NIR light oxidation potential [77]. Recently Yu et al. found that Cl- -modified
absorption on the Cu2-xS photocatalyst with efficient charge separation Cu2O can oxidize water without necessitating the formation of a heter­
at the interface between Cu2-xS and g-C3N4 [65]. When employed for the ojunction [78]. It was found that the Cl- doping lowered the band
photocatalytic conversion of CO2, the CSCN (0.4) composite sample bending, providing Cu2O with a suitable bandgap; however, this was
(representing 48.9% of Cu2-xS in g-C3N4) yielded 2.6 and 6.6 times more achieved at the expense of the decreased reduction potential of the CB
CO and CH4, respectively, as compared to pure Cu2-xS. The light ab­ electrons, which was still negative enough to reduce CO2. Additionally,
sorption spectroscopy results revealed that Cu2-xS absorbed the entire the Cl- modification/doping was found to be beneficial for photoexcited
region of the visible and IR light, whereas g-C3N4 was sensitive to only a charge separation, CO2 adsorption, better conduction of charges, and

5
S. Ali et al. Chemical Engineering Journal 429 (2022) 131579

the stabilization of the intermediates. transferred to Cu, as shown in Fig. 3b. In such an arrangement, the
In another study, Rajh et al. investigated the facet-dependent activity oxidation and reduction potentials of electrons and holes, respectively,
of Cu2O photocatalysts, and their findings revealed that the (1 1 0) facets decrease. Such a loss in the redox ability of the photogenerated charges
are active for CO2 reduction, while the (1 0 0) facets are inactive. They hampers the use of their true potential for photocatalysis. As is evident
found that the oxidation state of Cu2O changed reversibly from +2 to +1 from Fig. 3b, in type-II heterostructures, the separation of the photo­
after the CO2 adsorption and under illumination, respectively. The co- generated charges is achieved on account of the curtailment in their
adsorption of CO2 with water changed Cu1+ to Cu2+, quickly facili­ redox ability as well as activity. Therefore, this charge-transfer mecha­
tating its activation. Usually, Cu is an active site for CO2 reduction, but it nism is considered to be ineffective for CO2 reduction. In contrast, the Z-
also oxidizes water in addition to reducing CO2. Another interesting scheme charge transfer mechanism leads to the survival of the electrons
finding was the expansion of the crystal lattice by CO2 adsorption, which and holes with strong reduction and oxidation potentials, respectively,
reverted after the reduction of the adsorbed CO2 [19]. For the reduction at the expense of the weak electrons from the oxidation photocatalyst
of the liquid-phase CO2, the synthesized Cu2O exhibited an internal and the holes from Cu2O [83]. Hence, Z-scheme heterostructures are
quantum yield of 72%, which is comparable to that for the electro­ more advantageous for efficient photocatalytic CO2 reduction than type-
reduction of CO2. Further, Tang et al. observed that the selectivity of II heterostructures.
Cu2O for CO2 reduction can be increased as compared to that for H2 The charge transfer mechanism at the interface of the photocatalysts
production by controlling the facets. They found that the (1 0 0) facets defines the direction of charge transfer within the Z-scheme hetero­
were more active than the (1 1 0) ones. However, their study was less structure, which is largely governed by the band alignment. Under light
supported by spectroscopic evidence as compared to the study carried irradiation, the photogenerated electrons tend to accumulate on the
out by Rajh et al. They attributed the (1 0 0) facets to active proton semiconductor with a less negative Fermi level instead of being com­
reduction sites and found that the facets with lower Miller indices were bined with the counter-generated holes. Similar to the case of the Ti3C2
more photocatalytically active, which contradicts the previous reports MXene QDs@Cu2O NW heterostructure, the electrons from Cu2O shuttle
[79]. toward MXenes owing to its less negative Fermi level than that of the
Cu2O CB electrons [84]. This process continues until the equilibrium is
achieved. However, if these electrons are consumed upon reaching the
2.4. Heterostructures of Cu MXenes, the charges continue to flow. Once the equilibrium is achieved,
the band positions are altered. For example, in the case of Cu2O-TiO2,
There is no doubt that Cu as a co-catalyst with semiconductors is the band position rises up by 0.11 eV relative to the pre-contact condi­
efficacious in annulling charge recombination, thereby improving the tion, as shown in Fig. 4a and b [82,85]. Thus, the reduction potential of
yield of photocatalytic reactions [80–81]. In addition to avoiding charge these electrons determines the thermodynamic feasibility of the desired
recombination, these photogenerated charges should possess ample reaction; for instance, the MXene QDs@Cu2O NWs enable the photo­
redox potential to effectively convert CO2 into useful chemicals. The CB catalytic conversion of CO2 to methanol because the required CO2 to CO
of Cu2O is favorable for CO2 reduction, but its VB cannot oxidize water reduction potential (− 0.53 V vs. NHE) is more negative [84].
efficiently, as discussed in the previous section [82]. Hence, to align well The merits of the Z-scheme heterostructure can be further improved
with the high redox potential, wide bandgap photocatalysts are a better by increasing the light-harvesting capacity of wide bandgap oxidation
choice, but with compromised optical absorption. Thus, attaining higher photocatalysts that can supply a large number of protons by water
redox potentials and extended absorption on a single catalyst no longer oxidation. In this regard, Ali et al. reported Z-scheme photocatalysts
seems viable because of the trade-off between these two features [38]. composed of Cu2O and reduced titania (RT). In these photocatalysts, the
To address this concern, Cu2O/CuO can be stacked with an oxidation narrow bandgap of RT supplies a large number of e- and h+ to quench
photocatalyst to form a heterostructure while simultaneously main­ the Cu2O holes and oxidize water, respectively. Although the upward
taining its beneficial features [40]. The heterostructure formed by the band bending of RT weakens the oxidation potential of its holes, it still
selective coupling of semiconductors, one with strong reduction poten­ lies in the range suitable for water oxidation [40]. In addition, metal NPs
tial and the other with strong oxidation potential, results in well-aligned are deposited across the Z-scheme photocatalyst to further improve its
redox potentials for the CO2 photoreduction reaction (Fig. 3a). In activity. Xu et al. found that the deposition of Ag NPs over the ZnO-Cu2O
addition, by doing so, small bandgap materials (i.e., Cu2O), which are Z-scheme heterostructure further suppresses the charge recombination
considered to be dormant to CO2 reduction because of their unsuitable owing to the extra charge trapping sites provided by the Ag NPs.
reduction or oxidation potentials, can be effectively utilized for the Therefore, the resultant heterostructure (Ag-Cu2O-ZnO) shows signifi­
photoreduction of CO2 [38]. cantly enhanced CO production (~1.5 times) as compared to the bare
Two types of charge transfer can occur in CuO/Cu2O hetero­ Cu2O-ZnO [76].
structures based on their band alignment. In type-II heterojunction,
electrons are transferred to the oxidation photocatalyst, while holes are

Fig. 3. (a) Band potentials of various photocatalysts and (b) mechanism of charge transfer across different heterostructures [38] Copyright 2018, Elsevier.

6
S. Ali et al. Chemical Engineering Journal 429 (2022) 131579

Fig. 4. Energy band diagrams for (a) Cu2O and TiO2 before contact and (b) Cu2O/TiO2 composite [82] Copyright 2017, Elsevier.

2.5. Atomically dispersed Cu-based photocatalysts atomic-level utilization of single metal atoms considerably enhances the
efficiency of photocatalysts [43,86–88]. When combined with other
Recently, atomically dispersed reaction spots that are created by atomic-level reaction sites, such as Vo, these single metal atoms become
doping the photocatalyst support with Cu or some other heterogeneous more reactive. Li et al. carried out CO2 activation on sites doped with Cu
single atoms have become popular for developing high-performance Cu- atoms in the vicinity of defects (Ti3+, Vo). This arrangement spontane­
based photocatalysts. This is because the chemical and physical prop­ ously transformed CO2 to CO, even under dark conditions. The results
erties of Cu-based photocatalysts are altered upon the introduction of showed that Cu atoms with defective sites of TiO2 were more reactive,
these sites. Therefore, these sites have an impressive effect on the CO2 demonstrating the synergistic effect of Cu and TiO2 [86]. However, the
adsorption and its activation pathway, which ultimately controls the Cu atoms underwent oxidation during the course of the photocatalytic
reactivity and selectivity of the photocatalyst [16]. Moreover, the reaction. Recently, advanced spectroscopic techniques have confirmed

Fig. 5. Evolution of the solar products from Cu0/Cu1+@TiO2 relative to the other oxidation states of Cu [92] Copyright 2019, Elsevier, (b) proposed mechanism for
CO2 reduction over Cu-mTiO2 and, (c) the relative composition of the solar products obtained from Cu-mTiO2 [87] Copyright 2019, American Chemical Society.

7
S. Ali et al. Chemical Engineering Journal 429 (2022) 131579

the reduction of Cu under irradiation [42,87,89]. In addition, doping Cu++e-→Cu0 (2)


with a single metal atom helps to stabilize the photogenerated charges,
but brings minimal changes to the main absorption edge of the photo­ Cu0+CO2+2H+→Cu2++CO+H2O (3)
catalyst matrix. However, the optimally deposited Cu shows absorption
starting from the wavelength of 700 nm owing to the LSPR effect [89].
2+ 0
Cu +Cu →2Cu +
(4)
The addition of such plasmonic electrons increases the electron density,
which further promotes the CO2 photoreduction [90–91].
The Cu oxidation state significantly affects the activity of SACs. In a
2.5.1. Mechanism of reversible redox of Cu SACs
recent study, Li et al. investigated the synergistic effect of in situ-doped
The mechanism underlying the reversible redox reaction of Cu SACs
Cu0 and Cu1+ on the reduction of TiO2 (Cu0/Cu1+/Ti3+). They found
has been widely studied by monitoring the changes in the oxidation
that Cu1+-O acted as a bridge between the TiO2 support and Cu0,
states of Cu using sophisticated techniques such as electron para­
facilitating the transfer of the photogenerated charges from TiO2 to Cu0.
magnetic resonance (EPR) and X-ray absorption studies. In a previous
The shuttled electrons at Cu0 were efficiently utilized for reducing CO2
study Cu/TiO2 was synthesized by simple mixing and then annealing in
to various solar products (H2, CO, CH4), as shown in Fig. 5a. They noted
air at 400 ◦ C. The catalysts (state A) had the ability to expel O2 to form
that mixed oxidation states are essential for better performance, as the
Cu/TiO2-x (state B), as shown in Fig. 6a. This unique feature can be
yield of Cu1+-TiO2 was not satisfactory. They also prepared Cu-TiO2 by
explained on the basis of attachment of Cu species (Cu(OH)2) with -OTi
mechanical mixing and evaluated its CO2 reduction ability. The as-
(OH) groups. As can be seen in the working scheme, Cu2+ in its octa­
prepared Cu-TiO2 yielded the same products as those yielded by the in
hedral form attaches with two -OTi(OH) and OH groups, and there it
situ-prepared Cu-TiO2. However, the yield was not comparable, which
exists in higher coordinated form, contrary to CuO/Cu2O where it exists
can be attributed to the weak interaction of Cu and the TiO2 support
in low symmetry. This highly coordinated form of Cu(OH)2 attached
[92]. These findings highlight the importance of the oxidation states of
with oxygen frameworks on the surface of TiO2 inserts to four coordi­
Cu and their well-designed interactions with metal oxide supports for
nation sites in the TiO2 host, consequently Cu2+ has distorted geometry.
efficient charge transfer in photocatalysis. However, such interactions
As a result, the TiO6 insertion octahedron became more distorted and
alter the active oxidation states of Cu during the photoreaction [42].
facilitated the expulsion of O2 from the TiO2 lattice [89].
Therefore, the stabilization of these oxidation states is essential for a
In another study by Nam et al., the Cu doping of Ti vacancies
smooth photoreaction.
endowed the photocatalyst (under irradiation) with features same as
The metal-support interaction of SACs causes reversible changes to
those exhibited by SACs. Some Cu dx2-y2 and dz2 orbitals were aligned
the oxidation state of Cu, which is contrary to the common perception
as the mid-gap states along the TiO2 CB and VB. Upon illumination, the
that Cu undergoes an irreversible redox disproportion reaction. There­
photogenerated electrons became localized in the Cu dz2 state; conse­
fore, the Cu oxidation state change mechanism of such Cu SACs under
quently, the oxidation state of Cu changed from the Cu2+ to Cu1+,
illumination is different from that of Cu [93]. Owing to this behavior,
resulting in lattice distortion, which eventually converted Ti4+ to Ti3+.
SACs can be easily regenerated by exposure to the ambient environment.
This transformation of the Ti oxidation state could be gauged from the
Similar behavior was observed by Ozin et al. for Cu-TiO2 photocatalysts
change in the color of the photocatalyst from white to black. Cu acted as
during the photocatalytic reduction of CO2. In their experiment, the
a redox metal cofactor for TiO2 and tuned its activity periodically under
color of the catalyst was changed from white to red under illumination,
the irradiation and dark conditions. The Cu/TiO2-x photocatalyst in the
which is attributed to the reduction of Cu2+ to Cu0 by the transfer of the
aforementioned studies remained in an active state until all the Cu2+
photogenerated electrons from TiO2 to Cu2+. Furthermore, the activity
was converted to the Cu1+ state, which upon further reduction con­
of the photocatalyst reversed when the reactor was resealed after the
verted to Cu0 and became dormant. This change in the oxidation state of
exposure to the ambient conditions. This was probably due to the re-
Cu could be reversed by exposing Cu/TiO2-x to air until its dark color
oxidation of Cu [43]. Nam et al. and Yoon et al. also reported the
changed to white, as shown in Fig. 6b [42]. Such color transformation
reduction of Cu under illumination and re-oxidation upon exposure to
under irradiation in an oxygen environment has also been reported by
air and water, respectively [42,89]. Similar results have been reported
Biswas et al; they attributed this phenomenon to the change in the
by Yuan et al. for Cu atomically dispersed over mesoporous TiO2 (m-
oxidation state of Cu1+ to Cu0 [41].
TiO2), in which the oxidation state of Cu changed, as shown in Fig. 5b.
It is well documented that the reduction of TiO2 for the generation of
However, they did not describe the reversible behavior of Cu. They
Vo and Ti3+ is a tedious process. However, by utilizing the beneficial
noted that the Cu atomically dispersed over m-TiO2 adsorbed and acti­
features of SACs, the desired changes can be introduced simply by
vated CO2 more strongly than m-TiO2, which was reflected by a higher
illumination at room temperature and under moderate conditions.
yield, as shown in Fig. 5c [87].
Therefore, the single Cu atoms embedded into such photocatalysts help
In addition to TiO2, Cu SACs can be synthesized with other semi­
in stabilizing the defects (Vo, Ti3+). A similar role of Cu atoms in sta­
conductors, for example, Cao et al. synthesized Cu-dispersed carbon
bilizing Vo in CeO2-x was reported by Shi et al. [95].
nitride (Cu-CN) SACs using a thermal polymerization method. When Cu-
CN was employed for the photocatalytic CO2 reduction, the Cu in it
showed a reversible oxidation behavior, similar to that shown by the 2.6. Monitoring Cu oxidation states
above mentioned SACs. This behavior can be explained by the following
equations (Eqs. (1) and (2)). Cu+ accepted the photoelectrons and The identification of optimum Cu oxidation states can provide useful
converted into Cu2+ after the completion of the photocatalytic CO2 re­ information for the modification of photocatalysts, which may further
action cycle (Eq. (3)). Furthermore, Cu2+ was reduced by Cu0 to finally be conducive to improving the activity and stability of the photo­
revoke it to the desirable oxidation state of Cu+ (Eq. (4)) [94]. A slightly catalysts. However, this is an arduous task because Cu valence changes
different mechanism of the isolation of Cu atoms in a Pd lattice was rapidly under illumination which makes ex situ techniques Eective.
reported by Long et al. for Cu-Pd/TiO2 photocatalysts. The synchrotron- Interestingly, the in situ monitoring of oxidation states has made it
radiation-based spectroscopy results revealed that the isolation of the Cu possible, for instance, advanced in situ studies with X-ray absorption
atoms led to the adsorption of a large amount of CO2, thereby sup­ spectroscopy (XAS) have established that the Cu2+ state is favorable for
pressing the H2 evolution and activating CO2 by escalating the d-band the CO2 adsorption, while the Cu1+ state promotes the CO2 reduction
center of Cu [88]. [19]. In addition, the oxidation states of Cu can be determined using: (i)
XPS, (ii) EPR, (iii) in situ diffuse reflectance infrared Fourier transform
Cu-CN+sunlight→e-+h+ (1)
spectroscopy (DRIFT), (iv) ambient pressure XPS, and (v) H2

8
S. Ali et al. Chemical Engineering Journal 429 (2022) 131579

Fig. 6. Working models explaining the reversible oxidation-state change of Cu-TiO2 SACs for (a) photothermal, Reproduced from [89] with permission from the
Royal Society of Chemistry and (b) photocatalytic water splitting [42] Copyright, Springer Nature.

temperature-programmed reduction (TPR). Chang et al. monitored the


pre- and post-reaction oxidation states of Cu, and their XPS analysis
illustrated no appreciable change in the oxidation state [96]. In contrast,
Wu et al. noticed the transformation of Cu1+ to Cu0 during the photo­
catalytic reaction using XPS. They reduced Cu2O/TiO2 to Cu/TiO2 using
a thermal hydrogen treatment to further investigate the effect of this
transformation on the activity of the photocatalyst. The photocatalytic
activity results confirmed that Cu1+ is more suitable for CH4 and CO
production than Cu0 [97]. When XPS is unable to detect the signatory
satellite peak for Cu2+, a combination of XPS and EPR is employed [40].
This inability for the detection of Cu2+ can be attributed to its reduction
under the high vacuum operation of the X-ray photoelectron spectro­
scope. Excitement by X-rays and the presence of Cu2+ in low concen­
trations are the other reasons for this inability. However, Rajh et al.
reported that X-rays do not alter the oxidation state of Cu [19,98].
Likewise, the strong EPR peak at 2.1 g-value is attributed to Cu2+;
conversely, if the spectrum is a straight line, it is designated as Cu1+/
Cu0. Similarly, once we are able to differentiate between Cu2+ and Cu1+,
it becomes difficult to distinguish between Cu1+/Cu0. To resolve these Fig. 7. (a) DRIFT spectra for Cu-TiO2 with various oxidation states [99]
Copyright 2019, Elsevier and (b) in situ EXAFS spectra for Cu-mTiO2 [87]
issues, Auger peaks are analyzed in XPS and DRIFT studies with CO as a
Copyright 2019, American Chemical Society.
probe molecule. Similarly, the XANES spectrum of Cu (in situ or ex situ)
can be easily resolved to distinguish the various oxidation states of Cu.
Li et al. reported a similar study in which in situ DRIFT was employed that the oxidation of Cu1+/ Cu0 to Cu2+ by holes after the photocatalytic
to distinguish among the oxidation states of Cu for the photocatalytic reaction may be a possible cause of deactivation [99]. Koci et al.
CO2 reduction. The activity of the photocatalyst for CO generation employed XPS and XAS to identify the suitable oxidation states of Cu in
increased when pristine Cu2+-TiO2 was reduced to Cu1+-TiO2-x under an Cu0/TiO2 and Cu2+/TiO2 for the photocatalytic hydrogenation of CO2.
Ar atmosphere. This photocatalyst destabilized the inert CO2 molecules Their study revealed that in both the cases, the Cu1+ state appeared
even in the dark owing to the unique features of TiO2-x, that is, the under light illumination, which could be attributed to the oxidation of
presence of the Ti3+ state, and the affinity of Cu1+ for CO. Their post- Cu0 (Cu0 + h+ → Cu1+) or the reduction of Cu2+ (Cu2+ + e- → Cu1+). In
reaction DRIFT studies corroborated the transformation of the desired the case of Cu0/TiO2, the activity for methane production was the
Cu1+ to the passive Cu2+ state. However, this active state could also be maximum with the appearance of Cu1+ (Cu-Cu2O/TiO2) [71]. This
obtained by thermal treatment [86]. Inspired from the efficacy of the possible combination showed the best performance owing to the high
reduced Cu1+ state, they further extended the degree of reduction by reduction potential of Cu1+ for CH4 reduction, fast charge transport, and
carrying out a thermal treatment under hydrogen and compared the suppressed charge recombination by Cu0 [100]. Xiong et al. moved one
results with those obtained under the above mentioned conditions. The step further, and monitored the changes in the oxidation state of Cu
DRIFT analysis results (Fig. 7a) showed that Cu1+ was more conducive under light irradiation. The in situ XAS study reported by them showed
for CO2 reduction than Cu2+ because the optimum reduction potential of that Cu2+ first reduced to Cu1+ and then finally to Cu0, as shown in
Cu1+/Cu0 (0.52 eV) as compared to that of Cu2+/Cu0 (0.34 eV) allowed Fig. 7b [87]. Wu et al. also reported Cu1+ to be the suitable oxidation
Cu1+ to capture the photogenerated holes more efficiently. Furthermore, state for the photoreduction of CO2, although mixed oxidation states
for highly reduced samples, a combination of Cu1+ and Cu0, which could were present in the synthesized photocatalyst (Cu/TiO2). In another
be achieved through the spatial resolution of charges, proved to be more study, XPS and Fourier-transformed extended X-ray absorption fine
conducive because of the capturing of electrons and holes by Cu1+ and structure (FT-EXAFS) analyses confirmed that the activity of a photo­
Cu0, respectively. In addition, the post-reaction DRIFT study revealed catalyst for methanol production decreases when the oxidation state of
Cu changes from Cu1+ to Cu0 [101–102]. With the exception of a few

9
S. Ali et al. Chemical Engineering Journal 429 (2022) 131579

(11)

2H++2e-→H2 E redox=− 0.41V
studies (which reported Cu2+ to be the active state), most of the sur­
veyed literature (Table 1) designates Cu1+ as the active oxidation state There is a broad consensus among researchers that after the
for the photocatalytic reduction of CO2 [103]. adsorption, the CO2 photoreduction follows one of the following
mechanisms: (i) formaldehyde, (ii) carbene, and (iii) glyoxal pathways.
3. Selectivity of Cu-based photocatalysts However, the glyoxal pathway is not as widely reported as the other two.
Therefore, here we have discussed the carbene and formaldehyde
The selectivity of a photocatalyst is governed by many factors, which pathways, as shown in Fig. 8a. The formaldehyde pathway proceeds
can be classified as thermodynamic and dynamic. The thermodynamic through the formation of two intermediate products, H2CO and CH3OH.
factors consider the photon energy and the CB and VB edges of the The final product formation is strongly interlinked with the ability of the
photocatalyst necessary for the CO2 reduction. No photocatalyst pos­ photocatalyst to bind these intermediates if they are strongly adsorbed,
sesses the required reduction potential for transferring the 1st e- to CO2 and their further proton coupling yields CH4; conversely, they desorb as
(Eq. (5)). Therefore, it is achieved by designing surfaces that can 4- and 6-electron products. On the other hand, the carbene pathway
destabilize CO2 by bending it, and under the bent configuration, electron involves the formation of CO as a major intermediate product, which
transfer becomes feasible. The design of such sites falls under the then undergoes complete reduction to “C.” Thereafter, “C” is further
following prime dynamic factors: (i) light intensity, (ii) availability of reduced to form CH∙2 and CH∙3 radicals if the photocatalyst surface is
photogenerated charges, (iii) catalyst active sites, (iv) adsorption and incapable of stabilizing them, and then they are desorbed as methane;
stabilization of the reactants and intermediates, and (v) suitable CB and otherwise, they couple to form C2 + products, like, C3H8 [104]. There­
VB edges [104–105]. Thus, catalytic active sites with tuned adsorption/ fore, Cu role will be interesting in deciding the selectivity by adopting
stabilization characteristics in the presence of the required photo­ the formaldehyde or carbene pathways.
generated charge density can assure the desired selectivity for CO2 The adsorption of CO2 is the first step in its photoreduction reaction,
reduction [21]. The selectivity of a photocatalyst for CO2 reduction is and the strength of adsorption is a measure of the selectivity of the
primarily governed by the preferential adsorption of the reactants [106]. photocatalyst; therefore, the role of Cu in this context should be dis­
Therefore, a tradeoff should exist between the CO2 and H2O adsorption cussed [99,109].
in addition to the aforementioned conditions along with the presence of
H+ ions (Eq. (6)–(10)), which is imperative to guarantee the photo­
catalytic reduction of CO2 [107]. As is evident from Eq. (11), the dif­ 3.1. CO2 adsorption and activation by Cu
ference in the reduction potentials for the H+ and CO2 reduction
reactions is very small, and in some cases, the reduction of H+ ions is The surface chemistry of a photocatalyst can be altered by modifying
thermodynamically more feasible than the reduction of CO2. Thus, it is its surface by various techniques such as the introduction of Vo (Lewis
very challenging to suppress the production of H2 to sustain the supply acid), and its proximal surface hydroxyls (Lewis base) define the acidity
of H+ ions [108]. In such cases, two electron reduction products of CO2 and basicity of the surface [110]. The strength of CO2 adsorption is
are produced because the photocatalysts are incapable of stabilizing the defined by the relative concentration of the acidic and basic surface
intermediate products. However, no concrete mechanism has been sites. The addition of Cu to metal oxide photocatalysts enhances their
established regarding the origin of the variety of products. CO2 capturing ability by introducing various types of basic adsorption
sites ranging from weak to strong. The interaction of CO2 and its in­
(5)

CO2+e-→CO2▪-E redox=− 1.90V termediates with these sites plays a pivotal role in determining the
selectivity of the photocatalyst [109]. Usually, Cu induces moderate
(6)

CO2+2H++2e-→HCOOH E redox=− 0.61V
binding sites, which are considered to be instrumental for CO2 reduction
(7) [111]. Cu enhances the CO2 adsorption and activation ability of the

CO2+2H++2e-→CO+H2O E redox=− 0.53V
photocatalyst as compared to other metal oxide sites, such as TiO2, as
(8)

+ -
CO2+2H +2e →HCHO+H2O E redox=− 0.48V demonstrated by TPD, DRIFT, and XAS results [19,87–88,95].
(9)

+ -
CO2+4H +4e →CH3OH+H2O E redox=− 0.38V

(10)

CO2+8H++8e-→CH4+2H2O E redox=− 0.24V

Table 1
Factors affecting the activity and selectivity of bimetallic Cu-based photo­
catalysts for C1 selectivity.
Modification of Solar Selectivity/activity Factors changing
the products variation selectivity
photocatalyst

Pt-TiO2 → Cu- H2, CO, hydrocarbon Increased electron


Pt-TiO2 CH4 products = 23.0% → trapping and CO2
62.3% binding by Cu [118]
CO to CH4 = 83.3% →
80.0%
Pt-TiO2 → Pt- H2, CO, hydrocarbon Enhanced CO2
Cu2O-TiO2 CH4 products = 66.2% → adsorption and
100.0% abundant supply of H+
CO to CH4 = 71.1% → because of H2
96.6% suppression by Cu [97]
Pt-Cu- CH4, CH4 enhanced by 1.6 Switching the order
TiO2 → Cu-Pt- C2H6 times helped Cu atop to
TiO2 adsorb more CO2 [18] Fig. 8. (a) Possible photocatalytic CO2 reduction mechanism [95], (b) FT-IR
Ag-Cu- CH4 23% higher activity Switching the order spectra of CeO2− x and Cu/CeO2− x [104] Copyright 2019, American Chemical
TiO2 → Cu- helped Cu atop to boost
Society and (c) TPD CO2 desorption profiles for TiO2, Cu-TiO2, and Co-Cu-TiO2
AgTiO2 activity [119]
[113] Copyright 2019, American Chemical Society.

10
S. Ali et al. Chemical Engineering Journal 429 (2022) 131579

Furthermore, Yarmo et al. compared the relative abilities of CuO and approximately five times more CO with suppressed hydrogen genera­
Cu2O to adsorb CO2. Thermodynamic calculations were performed to tion. Thus, a combination of Cu, which shows the capability for CO2
determine the possible interaction of CO2 with CuO/Cu2O. This calcu­ activation, and Pt for water dissociation can effectively transform CO2
lation showed that reaction (12) is not possible because of thermody­ into hydrocarbon fuels [60,118]. Wu et al. reported a similar imple­
namic barriers; however, reaction (13) can proceed smoothly. mentation of this concept. They found that the activity and selectivity of
Furthermore, by different characterization studies, they proved that Cu2O/TiO2 for CH4 production improved after the deposition of Pt. Prior
reaction (13) is the dominant path for enhanced CO2 adsorption. The to the deposition, the Cu2O/TiO2 photocatalyst exhibited one third of its
thermal gravimetric analysis (TGA) showed that the CuO samples selectivity for CO production, which decreased to 3.4% after the addi­
exhibited 20 times higher weight loss than the Cu2O samples, as shown tion of Pt, and the amount of CH4 generated increased from 0.99 to
in reaction (14). Similar results were obtained from the XPS and 1.42 µmol g-1h− 1 after the deposition of Pt. In addition, the hydrogen
temperature-programmed desorption (TPD) measurements. A strong production of Pt-Cu2O/TiO2 was completely suppressed as compared to
CO2 desorption peak was observed at 505 K, indicating that the CO2 that of Pt/TiO2. This enhancement in the activity and selectivity was
adsorption ability of CuO was higher than that of Cu2O [112]. attributed to the preferential CO2 adsorption by Cu2O, improved charge
separation, and water dissociation by Pt, which ensured the abundant
Cu2O+CO2→CuO+CO ΔG=+147.9kJmol− 1
(12) availability of e- and H+ for the multi-electron reduction of CO2 to
CuO+CO2→CuCO3 ΔG=− 4.9kJmol− 1
(13) produce CH4 rather than CO [97]. In another study, Wang et al. reported
the same role of Cu and Pt in enhancing the CO2 reduction selectivity of
CuCO3→Cu2O+CO2+1/2O2 ΔG=− 4.2kJmol − 1
(14) a Cu-deposited Pt/TiO2 photocatalyst. The CO2 reduction selectivity of
TiO2 increased from 41% to 85% with 2.64 times suppressed H2 gen­
As a result of the activation process, different intermediates were
eration when Pt was deposited on it and Cu was used as the co-catalyst.
formed, such as bicarbonate (HCO3- ), monodentate carbonate (m-CO-32 ),
This study also entailed an interesting finding to control the selectivity
and bidentate carbonate (b-CO3-2 ), as shown in the DRIFT and TPD of bimetallic catalysts by changing the order of metal deposition. For
spectra (Fig. 8b and 8c) [95,113]. Usually, these intermediates are example, when Pt was deposited on Cu/TiO2, its selectivity for CO2
transformed into C1 products, but it has been reported that Cu stabilizes reduction suddenly decreased to 47 µmol g-1h− 1 while doubling the H2
these intermediates and facilitates their C–C coupling to form production rate to 51 µmol g-1h− 1 [118]. The findings of these studies
C2 + products [18,72,113]. In addition, in their study on Cu-TiO2, were further supported by In et al., who highlighted the role of the hi­
Hemminger et al. reported that Cu also offers dissociative adsorption of erarchical deposition of Cu and Pt over the surface of reduced TiO2,
water, ensuring an abundant supply of H+ for multi-proton-coupled known as blue titania (BT). They noticed that the CH4 production
electron transport for generating C2 + products. In addition, the selectivity of the photocatalyst increased by 1.61 times when Pt was
neighboring environment of Cu also affects its selectivity, as observed in deposited on Cu-BT as compared to the case when Cu was deposited on
the case of bimetallic photocatalysts, metal oxides with defects, and non- Pt-BT. The improvement in the activity of the photocatalyst was
metals, such as graphene [18,72]. Apart from the Cu surface properties, attributed to the efficient capture of the photogenerated electrons by Pt
the ambient environment to which a Cu-based photocatalyst is exposed as compared to that by Cu and the enhanced CO2 adsorption by Cu [18].
also influences its selectivity. Cu is a well-known adsorbent for CO2 and Granchak et al. also emphasized the importance of hierarchal nano­
other organics present in the atmosphere during the synthesis process structures, and found that the activity of Cu-deposited Ag/TiO2 was 23%
[114]. When catalysts with pre-adsorbed species are used for the pho­ higher than that of Ag-deposited Cu/TiO2. Moreover, Cu-deposited Ag/
tocatalytic reduction of CO2, the yield emanates from the adsorbed TiO2 produced two and five times more CH4 than Cu/TiO2 and Ag/TiO2,
species and CO2 reduction as well [115]. Therefore, it is challenging to respectively. They related this activity enhancement of Cu@Ag/TiO2 to
determine the individual contribution of these adsorbed species to the the better electron transfer ability of Ag from titanium to copper [119].
overall selectivity/yield of the photocatalyst as the product composition Similarly, Xu et al. reported that in Ag-Cu2O/ZnO, Cu enhances the CO2
is usually determined by gas chromatography-flame ionization detec­ adsorption, while Ag extracts the photogenerated electrons to ensure
tion, which cannot distinguish the origin of the carbon. Therefore, the their abundant availability for CO2 reduction [76]. Another study by
over-estimation of the selectivity/yield is an obvious outcome, and it can Brisard et al. revealed the bifunctionality of a partial monolayer of Cu
be avoided by employing gas chromatography mass spectrometry, in over the Pd surface. The Cu monolayer enhanced the CO2 adsorption on
which isotopic 13CO2 is reduced to produce 13C-containing solar prod­ the Pd surface, while Pd promoted the conversion of CO2 to CO [120].
ucts [116]. The relative concentration of these products in the chro­ Therefore, it can be concluded that bimetallicity extends the utilization
matogram can easily designate the participation of the aforementioned of different catalytic features over a single surface. The efficacy of
reactions. The role of Cu with other co-catalysts in tuning the selectivity bimetallicity in improving the activity and selectivity of Cu-based pho­
is discussed in the following section. tocatalysts is briefly summarized in Table 1, and the overall performance
of C1 products is summarized in Table 2.
3.2. C1 selectivity
3.3. C2 + selectivity
The role of co-catalysts in curbing the e-/h+ recombination, and
hence in improving the activity of photocatalysts, is well-documented in Earlier studies on photocatalytic CO2 reduction were focused on
the literature. Moreover, bimetals also tune the adsorption behavior of enhancing the optical response, suppressing the charge recombination,
gases over the catalyst surface, which leads to a selectivity control over and converting CO2 to C1 products [121]. With significant advancement
the products [27]. Noble metals such as Ag, Au, Pt, and Pd have been in these avenues, great interest has been aroused in converting CO2 to
extensively studied for application as co-catalysts for effectively con­ C2 + products owing to its economic benefits [21]. In this regard,
verting CO2 into solar fuels. These co-catalysts can impart special fea­ considerable efficiencies with higher C2 selectivity have been achieved
tures to photocatalysts by providing active catalyst sites possessing by the electrocatalytic reduction of CO2 [122–123]. However, the use of
higher electron density, which is endowed by their ability to capture the electricity, which is a refined form of energy generated from solar,
photogenerated charges. These electron-rich catalytic sites may tune the hydel, or fossil fuels involves generation and transmission losses.
activity and selectivity of the photocatalyst depending on the semi­ Therefore, the direct use of sunlight to reduce CO2 to store energy in the
conductor to which they are attached [117]. Wang et al. reported that Pt form of chemical bonds is simpler in terms of operation and for
attached with nitrogen-doped TiO2 nanotubes (Pt/N-TiO2) is more achieving high STF efficiencies [124]. For C2 + selectivity, a photo­
prone to hydrogen formation than Cu/N-TiO2, which produces catalyst should exhibit high charge density, multi-electron transfer [19],

11
S. Ali et al. Chemical Engineering Journal 429 (2022) 131579

Table 2
Summary of activity, selectivity, and stability of Cu-based photocatalysts with suitable oxidation states for C1 products.
Composition of Photocatalyst Active oxidation state Light Source, Reaction Conditions Main Products
1+ 1+ − 2
Cu0.50-TiO2 supported on SiO2 [41] Cu conversion of Cu Xe-arc lamp (intensity = 2.4 mW cm ) CO 60 µmol g-1h− 1
to Cu0 λ = 250–400 nm CH4 10 µmol g-1h− 1
Continuous gas flow system
Pd7Cu/TiO2 Cu0 Xe-arc lamp (Power = 300 W, 2.0 mW cm− 2) H2 1.45 µmol g-1h− 1
“7” is molar ratio of Pd to Cu [88] λ less than 400 nm CO 1.90 µmol g-1h− 1
Batch reactor CH4 19.60 µmol g-1h− 1
Photocatalyst exhibits stable performance for 28 h
Pt-Cu2O/TiO2 [97] Cu1+ Xe-arc lamp (300 W, 20.5 mW cm− 2) CO 0.05 µmol g-1h− 1
λ = 250–400 nm CH4 1.42 µmol g-1h− 1
Stable for 3 cycles (total 12 h)
Cu-Pt/TiO2 [118] Cu1+ Xe-arc lamp (200 W) H2 25 µmol g-1h− 1
λ = 320–780 nm CO 8.3 µmol g-1h− 1
Continuous gas flow system CH4 33 µmol g-1h− 1
Cu-AgTiO2 [119] Cu0 DRSh-1000 lamp (5x10–6 einsteins/min) CH4 3.0 µmol g-1h− 1
Batch reactor
Cu-Pt/TiO2 [117] Newport Solar simulator (AM 1.5) CH4 574 nmol cm-2h− 1
Continuous gas flow system CH4 610 nmol cm-2h− 1
For 99.9% and 0.99% CO2 respectively
CQDs/Cu2O Xe-arc lamp (300 W) CH3OH 55.7 µmol g-1h− 1
Carbon quantum dots (CQDs) [133] λ = full spectrum
Cu2O/WO3 [139] Xe-arc lamp (300 W) CO 11.7 µmol
λ > 400 nm H2 5.7 µmol
Batch reactor 4-cycle stable
Cu-TiO2 [130] Cu1+ Two 250 W IR lamp for heating up to 150 ◦ C Maximum rate of CO formation at 1.5 h was
Mercury vapor lamp (10 mW cm− 2) 7.5 µmol g-1h− 1
λ > 390 nm
Cu2-xS/g-C3N4 (CSCN) [65] Xe-arc lamp (300 W) Full spectrum
CO 319.4 µmol g-1h− 1
CH4 23.7 µmol g-1h− 1
> 390 nm
CO101.1 µmol g-1h− 1
CH4 21.0 µmol g-1h− 1
CuS atomic layer [64] Xe-arc lamp with AM 1.5G filter (300 W) CO14.5 µmol g-1h− 1
λ > 800 nm 8-cycle stability
100% selectivity
Ag-Cu2O-ZnO [76] Cu1+ Xe-arc lamp (300 W) CO 13.45 µmol g− 1
Z-scheme heterostructure
Ti3C2(QDs)@Cu2O(NWs) [84] Xe-arc lamp (300 W) CH3OH 78.5 µmol g-1h− 1
Cu@Nb3O8 Cu2+ Hg-Xe lamp CO ≈ 1.42 µmol for ≈ 20 h
(Cu clusters) [142] λ > 240–300 nm
Cu2O/graphene/TNTA (Heterostructure) Cu1+ Xe-arc lamp (300 W, 100.0 mW cm− 2) CH3OH 45 μmol cm-2h− 1
[131] λ > 400 nm (60 h stability with 6 cycles and only decrease of 18%
Double chamber reactor after that)
Cu2O-SiC [141] Cu1+ Xe-arc lamp (500 W) CH3OH 191.0 µmol g− 1
Liquid phase reaction
Cu2O-TiO2-x Cu1+ Xe-arc lamp with AM 1.5G filter (100 W, 100.0 CH4 462 nmol g− 1
Z-scheme heterojunction [40] mW cm− 2) 7-cycle stability
1
RuOx@ cuboid-Cu2O [141] Xe-arc lamp (150 W) CO ~ 200.0 ppm g−
Liquid phase reaction
Z–Scheme Cu1+ Xe-arc lamp (300 W) CO 3.3 µmol g− 1

α-Fe2O3/Cu2O heterostructures [85] λ > 400 nm


Au3Cu@STO/TiO2 [73] Xe-arc lamp (300 W) CH4 421.2 µmol g-1h− 1
Liquid phase reaction CO 3770.0 µmol g-1h− 1
Stable up to 6 cycles and retained 87.6% stability
Cu2O [19] Cu2+/ Cu1+ Xe-arc lamp (300 W) CH3OH 1.2 mol g− 1h− 1
Liquid phase reaction
Cu2O [79] Cu1+ Xe-arc lamp (150 W) CO ~ 60 ppm g− 1

λ ≥ 350 nm
Liquid phase reaction
Cu-TiO2 [43] Cu1+/ Cu0 Xe-arc lamp (300 W) CO ~ 2.0 µmol g-1h− 1
Liquid phase reaction Stable for 6 cycles
ZnO-Cu2O Cu1+ Xe-arc lamp (300 W) CH4 1080 µmol g-1h− 1
Z-scheme hybrid nanostructure [67] Liquid phase reaction 3 cycles
Cl-doped Cu2O [78] Cu1+ Xe-arc lamp (350 W) with a 420 nm cutoff filter CO 1.74 μmol cm− 2
CH4 to 0.39 μmol cm− 2
4-cycle stability
CuO-TiO2− xNx(heterojunction) [143] Cu2+ Xe-arc lamp (300 W, 100.0 mW cm− 2) CH4 41.3 ppm g− 1 h− 1
Cu2O/TiO2 nanojunction (type-II) [66] Cu1+ Xe-arc lamp (300 W) CH4
28.4 ppm g− 1 h− 1
Cu-CN Cu1+/ Cu0 Xe-arc lamp with AM 1.5 filter (300 W, 100.0 CO 11.21 µmol g− 1
Thermal polymerization method [94] mW cm− 2) CH3OH 1.75 µmol g− 1
CH3OH 0.61 µmol g− 1
Cu-CeOx-2 Cu0 Xe-arc lamp (300 W) CO 8.25 µmol g− 1
SACs [95]
(continued on next page)

12
S. Ali et al. Chemical Engineering Journal 429 (2022) 131579

Table 2 (continued )
Composition of Photocatalyst Active oxidation state Light Source, Reaction Conditions Main Products

Cu-TiO2 (Cu0/Cu1+@Ti3+) [92] Cu1+/ Cu0 Xe-arc lamp (300 W) CO ~ 1.70 µmol g− 1
CH4 7.75 µmol g− 1
10 cycles
[96] Cu-TiO2 Cu1+ Xe-arc lamp (300 W, 7.3 mW cm− 2) CO 10.9 µmol g-1h− 1
λ = 300–400 nm CH4 4.2 µmol g-1h− 1
Continuous gas flow system 3 cycles
BiVO4-C-Cu2O [135] Cu1+ Xe-arc lamp (300 W, 100.0 mW cm− 2) CH4 ~ 3.01 µmol g-1h− 1
λ > 420 nm 5-cycle stability (Total 20 h)

high CO2 adsorption/reduction, suppressed water oxidation, and strong products is shown in Fig. 9d. In another study, Liu et al. found that the
binding of the reaction intermediates [125]. Cu and its compounds show surface of Co-Cu/TiO2 showed CH∙3 radical enrichment as compared to
efficient multi-electron transfer because they can utilize weakly bonded Cu/TiO2. This CH∙3 radical enrichment enhanced the selectivity of Co-
d-band electrons. In addition, their short bandgap ensures the genera­ Cu/TiO2 for C2 + products (ethane and propane). Before the addition of
tion of sufficient charges by harvesting sunlight in the visible region Co, Cu/TiO2 was more prone to the formation of C1 products, as shown
[126]. in Fig. 10a and b. They noted the strong adsorption and activation of
However, this seems to be a challenging task, which cannot be CO2 in the form of different intermediates, including HCO-3 , m-CO-32 , and
executed until a perfectly architectured catalyst with symmetry capable b-CO-2 , as confirmed by the TPD and DRIFT results. They ascribed the
3
of conferring C–C coupling is available. Converting CO2 to C2 + products C2 + selectivity of Cu/TiO2 to the efficient separation of electrons and
is a thermodynamically arduous task, and Cu-based photocatalysts have holes in Cu and Co, respectively. As a result, Co produced more protons,
recently emerged as promising candidates to efficiently generate C2 which eventually generated more CH∙3 radicals. Moreover, the excess
products. However, the generation of C2 + products is still challenging protons promoted the formation of CH∙3 radicals from the already
because of the complexity of the intermediate reactions. The
formed CH4, which probably ensures a sufficient supply of these radicals
C2 + product formation ability of Cu-based photocatalysts stems from
to produce thermodynamically challenging C3H8. The overall mecha­
their CO2 adsorption and intermediate stabilization features. Recently,
nism of these reactions is shown in Fig. 10c.
In et al. reported the unprecedented CH4-to-C2H6 selectivity transition of
In addition to the above-mentioned constraints for C2 + products,
Cu-deposited BT-Pt catalysts. This phenomenon can be attributed to the
Shankar et al. found that the inter-adsorbate repulsive forces between
quick channeling of the high electronic flux from BT to Cu by Pt, and the
the intermediates formed during the photocatalytic CO2 reduction re­
high density of the stabilized intermediates (methyl radicals, CH∙3 )
action restrict the C–C coupling [125]. These repulsive forces can be
triggered the formation of multi-electron product, as shown in Fig. 9a avoided by designing adjacent reaction sites with opposite charges.
and b. Liu et al. also reported the electron enrichment of CuO(QDs)-MIL- Following this scheme, they prepared plasmonic AgCu-deposited TiO2
125(Ti) by constructing a heterojunction with g-C3N4, which shifted the nanotube arrays (AgCu-TNTAs) with the dominant multipolar plasmon
selectivity to C2 + products (Fig. 9c). Thus, the photogenerated electrons resonance mode, which enabled the reaction sites to have opposite
from MIL-125(Ti) and g-C3N4 collectively enhanced the electron density charges. Therefore, the repulsion between CO∙2- and CO* (excited form
over the CuO QDs. As a result, the initial product composition changed
of CO) reduced considerably. Furthermore, they noticed that both Ag
from C1 (CH3OH, CO) to mainly C2 + products (CH3CH2OH, CH3CHO)
and Cu could stabilize CH∙3 radicals because the C–C coupling was
for g-C3N4-CuO(QDs)@MIL-125(Ti), amounting to 77.7% of the total
favored on these surfaces. They explained the stabilization of
formed products [128]. The overall mechanism of the formation of these
CH∙3 radicals on the basis of charge transfer, where the plasmonic elec­
trons were injected into the TNTAs at the interface leaving behind holes,
which imparted a positive charge to the metals. These positively charged
metals can extend the lifetime of the CH∙3 radicals. Moreover, they
highlighted the importance of hot spots, where a strong local electric
field can facilitate the polarization of the CO2 molecule, which also fa­
vors the formation of various C2 + products, that is, ethane, as shown in
Fig. 11a and b [125]. Chen et al. also reported the photocatalytic
reduction of CO2 to CH3OH over Cu NP (4–5 nm)-decorated graphene
oxide (GO). The possible formation of C2 + products, as reported by
Shankar et al., sounds conceivable because they noted that upon the
introduction of Cu NPs, the Fermi level of the Cu-GO composite shifted
to more negative levels as Cu accepted the photogenerated electrons and
the holes accumulated on the GO surface. Therefore, in addition to
efficient charge separation at the Cu-GO interface, oppositely charged
reaction sites were created, which probably led to the formation of
C2 + products [72].
The selectivity of bimetallic photocatalysts (involving Cu) also de­
pends on the oxidation state of Cu. In this regard, Qin et al. reported
highly dispersed Cu-TiO2, which exhibited selectivity toward the pro­
duction of olefins (ethylene and propylene). They noted that Cu1+
facilitated C–C coupling, which led to the formation of C2 + products,
and the selectivity reached the maximum value (60.4%) when the Cu1+
(Cu1+/Cu1+ + Cu0) ratio reached 75.7%, as shown in Fig. 12a and b. In
Fig. 9. C2 + selectivity for (a-b) xPt-BT to bimetallic CuxPty-BT[127] Copyright addition, other reaction parameters such as the temperature and feed
2017, Elsevier, Reproduced from [18] with permission from the Royal Society gas composition also contributed to the formation of olefins. At ambient
of Chemistry, (c) CuO(QDs)@MIL-125(Ti) to g-C3N4-CuO(QDs)@MIL-125(Ti) temperature, methane was the main product; however, at higher
and (d) reaction mechanism [128] Copyright 2020, Elsevier.

13
S. Ali et al. Chemical Engineering Journal 429 (2022) 131579

Fig. 10. Change in the selectivity of bimetallic photocatalysts from solar products to C2 + products (a) Cu-TiO2, (b) Co-Cu-TiO2, and (c) reaction scheme for the
photocatalytic reduction of CO2 over Co-Cu/TiO2 [113] Copyright 2019, American Chemical Society.

Fig. 11. (a–b) Effect of bimetallicity on the selectivity of AgCu-TNTAs for C2 + products, particularly ethane [125] Copyright 2021, American Chemical Society.

14
S. Ali et al. Chemical Engineering Journal 429 (2022) 131579

Fig. 12. Improvement in the C2 + product selectivity of (a-b) Cu-TiO2-RO [129] Copyright 2019, Elsevier and (c-d) CuOx@ZnO with a change in the oxidation state
of Cu [24] Copyright 2021, American Chemical Society.

reaction temperatures, the selectivity switched to olefins, and a similar can also occur [131]. A change in the oxidation state of Cu significantly
trend was observed with an increase in the light intensity [129]. Another affects its photocatalytic yield [19,41]. Hence, maintaining the stability
study by Wang et al. demonstrated the role of the oxidation state in of Cu is imperative to garner its inherent features for photocatalytic
maneuvering the selectivity of CuOx/ZnO. For the synthesized CuOx/ applications. To do this, various remedial strategies have been explored,
ZnO, the highly dispersed Cu oxidation state changed to Cu1+ during the including the preparation of bimetallic composites, wrapping Cu with
irradiation. Cu1+ facilitated the adsorption of the in situ generated CO, non-metals such as graphene, and formation of heterojunctions with
which further underwent C–C coupling through the formation of an other photocatalysts [40,64,76,118–119,131].
intermediate, *OC-COH. Owing to this transition in the oxidation state
of Cu, a C2H4 selectivity of 32.9% was achieved, as illustrated in Fig. 12c
4.1. Metals and non-metals
and d [24]. Table 3 summarizes the activity and selectivity of Cu-based
photocatalysts for C2 + products.
The combination of Cu with other metals, particularly Pt, for
On the basis of this discussion, it can be stated that the stabilization
inhibiting the oxidation of Cu has been well-reported in the literature.
of the reaction intermediates and the abundance of the photogenerated
However, to date, this inhibition has not been backed by any concrete
charges help to achieve C2 + selectivity in bimetallic Cu-based photo­
mechanism, except considering water oxidation to proceed with metals
catalysts. Fig. 13 shows the other factors that contribute to the
combined with Cu. Recently In et. al. reported a bimetallic catalyst, Cu
C2 + selectivity of Cu-based photocatalysts.
deposited over Pt@reduced-TiO2 (RT), which showed excellent stability
for a total of 60 h (Fig. 14a). They attributed this excellent durability to
4. Stability of Cu-based photocatalysts
the effective electron extraction ability of Pt, which led to the accu­
mulation of electrons on Cu to protect it from corrosion [18]. Similar
To date, numerous studies have been conducted to identify the
results have been reported by Wu et al. for a Pt-deposited Cu2O/TiO2
causes for the deactivation of photocatalysts, particularly Cu-based
photocatalyst, which exhibited excellent stability for three cycles (12 h)
photocatalysts [130]. The possible causes of deactivation are (i) filling
(Fig. 14b) [97]. In addition to these beneficial characteristics conferred
of Vo, (ii) change in the oxidation state of Cu by the photogenerated
by Pt, the core–shell configuration of the finally formed Pt-Cu2O/CuO
holes, (iii) consumption of surface hydroxyl groups, that is, Ti3+/
also played an important role in maintaining the stability of the pho­
Ti4+–OH, and (iv) piling up of the reaction intermediates [130]. Despite
tocatalyst. In this regard, Shankar et al. highlighted that the core of
its merits, the use of Cu in photocatalysis remains in its infancy owing to
noble metals efficiently captures the photogenerated electrons from the
its poor stability due to the changes in its oxidation states by the pho­
metal oxide semiconductor for a sustained supply to Cu [97,117–118].
togenerated e- and h+. As the oxidation potential of holes matches well
These results clearly explain the role of Pt in maintaining the stability of
with that of Cu, the corrosion of Cu is an obvious outcome. Similarly, the
Cu-based photocatalysts. In another study, MXene quantum dots (QDs)
reduction potential of the Cu is 0.05 eV (vs. NHE); it is highly compet­
were grown over Cu2O, promoting the stability of the stabilized Cu2O, as
itive with that for CO2 reduction and is less negative than the potential
evident from the photocatalytic yield, which remained ~ 89% in the 6th
required for producing solar products. Therefore, the reduction of Cu
cycle of testing [84]. Liu et al. reported the stability of g-C3N4-CuO(QDs)

15
S. Ali et al. Chemical Engineering Journal 429 (2022) 131579

Table 3 enrichment of CuO because of the heterostructure charge transfer


Summary of activity, selectivity, and stability of Cu-based photocatalysts with mechanism. These features prevented the oxidation of active Cu species,
suitable oxidation states for C2 products. as confirmed by the XPS results before and after the irradiation
Composition of Active Light Source, Reaction Main Products (Fig. 14d) [128].
Photocatalyst oxidation Conditions Apart from noble metals, several non-metals, such as graphene, have
state also proven to be advantageous for maintaining the stability of Cu [12].
Cux%–Pty%–BT Xe-arc lamp with an CH4 3.0 mmol g- Tang et al. reported a significant enhancement in the stability of Cu2O
1 − 1
(Reduced AM 1.5 filter (100 W, h wrapped with graphene. They found that the CO yield for graphene-
titania) [18] 100 mW cm− 2) C2H6 0.15 mmol g-
1 − 1 wrapped Cu2O increased continuously over a prolonged irradiation
h
Stable for 10 cycles time of 20 h as compared to that for pristine Cu2O and intentionally
(total 60 h) oxidized Cu2O [132]. On the other hand, Cao et al. proposed a different
Cu-Co/MIL-125 Cu1+ / Cu2+ Xe-arc lamp (300 W) CO 150.58 µmol g- approach for their Cu2O/graphene/titania-nanotube array (TNTA) het­
1 − 1
[113] h erostructure. The oxidation and reduction reactions were carried out in
CH4 169.79 µmol g-
1 − 1
h
two different compartments of the reactor, as shown in (Fig. 15a). They
C2H6 prevented the self-oxidation and reduction of Cu2O on the illuminated
267.60 µmol g-1h− 1 side by the transfer of electrons to the TNTA heterostructure for the CO2
CH3H8 reduction and holes for the water oxidation. The XPS and XRD analyses
10.07 µmol g-1h− 1
showed that only a small amount of Cu2O was reduced in the early stage
GO-Cu2O [80] Cu0 Xe-arc lamp (300 W, CH3CHO
100.0 mW cm− 2) 3.88 µmol g-1h− 1 of the reaction, after which it remained unchanged. This can be attrib­
CH3OH 2.94 µmol g- uted to the transfer of electrons to the graphene layer, leading to the
1 − 1
h reduction of CO2. The leftover holes on the exposed site triggered the
g-C3N4/ Cu1+ Xe-arc lamp (300 W) CO 60.0 µmol g- oxidation of water along with the partial oxidation of Cu2O to CuO on
1 − 1
CuO@MIL-125 λ = 420 nm h
the illuminated side. Despite this, the photocatalysts maintained their
(Ti) [128] Batch reactor CH3OH
332.4 µmol g-1h− 1 stability for 10 cycles (Fig. 15b) [131].
C2H5OH In addition to graphene-based materials, carbon QDs have also been
501.9 µmol g-1h− 1 reported to maintain the stability of Cu photocatalysts. MacFarlane et al.
CH3CHO
investigated the photocatalytic action of carbon-based materials, carbon
177.2 µmol g-1h− 1
Cu/TiO2-RO Cu1+ / Cu0 Xe-arc lamp (136.0 Ethylene and quantum dots (CQDs) with Cu2O. They attributed the enhanced stability
[129] mWcm− 2) propylene of the CQDs to their improved hole-accepting ability [133]. Yu et al.
1.52 µmol g-1h− 1 carried out the XPS analysis of carbon-coated Cu2O (C-Cu2O) and found
Selectivity of 60.4% that the carbon coating helped in stabilizing the oxidation state of Cu2O.
over methane
They noted that after six cycles of irradiation, the pristine Cu2O un­
CuOX@p-ZnO Cu1+ Xe-arc lamp (300 W) CO 27.3 µmol g− 1
[24] CH4 17.9 µmol g− 1 derwent an oxidation state change because a distinct CuO signatory peak
C2H6 22.3 µmol g− 1 was observed at the binding energy of 935.6 eV; however, no significant
H2O as hole scavenger CO 3.3 µmol g− 1 difference was observed for C-Cu2O. Furthermore, surface X-ray
CH4 2.2 µmol g− 1
diffraction (S-XRD) also confirmed that the crystalline structure of C-
C2H6 2.7 µmol g− 1
Ag-Cu-TNTAs Solar simulator CH4 9.38 µmol g-
Cu2O could be retained even after the irradiation. The SEM images also
[125] 100 W cm− 2 1 − 1
h corroborated the S-XRD results with minor changes in the morphology
C2H6 14.5 µmol g- of the tested C-Cu2O. On the other hand, Cu2O showed particle
1 − 1
h agglomeration. Owing to such remarkable passivation of the active
C-Cu2O [134] Cu1+ Xe-arc lamp (350 W, CH4 and C2H4
oxidation states (Cu1+), C-Cu2O showed only a 7.0% loss in the activity
1.16 mW cm− 2) with a 0.385 µmol
420 nm cutoff filter 6-cycle stability as compared to that of Cu2O (46.0%). Even under prolonged testing, C-
Liquid-phase reaction (Total 18 h) Cu2O continued to yield solar products (CH4, C2H4) until the 28th hour
[134]. Jung et al. also reported similar results for carbon-coated Cu2O
nanowire arrays. They found that C-Cu2O maintained 91% of its activity
@MIL-125(Ti) heterostructures for up to four cycles, as shown in until the 5th cycle of testing as compared to the pristine Cu2O, and the
Fig. 14c, even for C2 + products. They attributed it to the encapsulation stability was further enhanced to 98% when a Z-scheme heterostructure
of the CuO QDs in the pores of MIL-125(Ti) and the enhanced e-

Fig. 13. The summarized overview of factors governing the C2 + selectivity of Cu-based photocatalysts.

16
S. Ali et al. Chemical Engineering Journal 429 (2022) 131579

Fig. 14. Stability of various photocatalysts (a) bimetallic CuxPty-BT, Reproduced from [18] with permission from the Royal Society of Chemistry, (b) Pt-Cu2O TiO2
[97] Copyright 2016, Elsevier and (c-d) stability and XPS results for g-C3N4-CuO (QDs)@MIL-125(Ti) [128] Copyright 2020, Elsevier.

Fig. 15. (a) Double chamber reactor for separate H2O oxidation and CO2 reduction and (b) the stability exhibited by the Cu2O/graphene/TNTAs photocatalyst in this
reactor [131] Copyright 2016, Elsevier.

17
S. Ali et al. Chemical Engineering Journal 429 (2022) 131579

with BiVO4 was constructed (BVO-C-Cu2O) [135]. efficacious route to boost the stability of Cu-based photocatalysts. In
From this discussion, it can be concluded that the stability of Cu can addition to the quenching of Cu oxidizing holes, the Z-scheme charge
be extended by combining it with noble metals, such as Pt, or non- transfer mechanism involves the shifting away of the oxidation sites
metals, such as graphene. However, the commercial applications of from the surface of Cu [38]. This shifting prevents the oxidation of Cu by
such combinations are limited. Therefore, a cost-effective approach •
OH radicals because the VB of Cu is not positive enough to oxidize these
should be developed to maintain the stability of Cu. In this regard, radicals (2.3 eV vs. NHE) [136]. Furthermore, in the Z-scheme charge
constructing Z-schemes of CuO/Cu2O with other earth-abundant metal transfer, the continuous injection of electrons helps stabilize the Cu
oxides having a stronger oxidation potential than Cu could be a possible oxidation states.
avenue. Owing to their charge transfer mechanism, Z-scheme hetero­
structures have emerged as potential alternatives to the aforementioned
Cu passivation techniques. Their charge transfer mechanism arrests the
4.2. Z-scheme heterostructures superfluous Cu oxidizing holes while leaving behind the desirable
charges [38,138]. Various reports have corroborated the indispensable
Metals and non-metals are undoubtedly helpful for ensuring the role of Z-scheme heterostructures in improving the durability of the
stability of Cu-based photocatalysts. However, several approaches have photocatalytic reduction of CO2. For example, the Cu2O/g-C3N4 Z-
been proposed to improve the stability of Cu-based photocatalysts scheme heterostructure constructed by Ivaska et al. showed excellent
without using any noble metals. One such promising approach is the stability over 10 cycles, as shown in Fig. 16a. Similarly, Grela et al.
suitable charge transfer by heterostructures. Constructing hetero­ (Fig. 16b) and Wang et al. reported Cu2O-TiO2 and Cu2O-WO3, respec­
structures is a promising alternative to using expensive metals/non- tively, which showed excellent stability over 4 cycles [38,137,139].
metals for improving the stability of Cu-based photocatalysts. Among They noted that the morphology and oxidation states of the photo­
the heterostructures of copper oxides, type-II heterostructures are not catalysts did not change during the photocatalytic reduction of CO2,
favorable for this task, because electrons from Cu are transferred to which can be attributed to the beneficial features of the Z-scheme charge
other semiconductors, while holes accumulate on the surface of Cu [9]. transfer mechanism. In et al. combined the synergistic effects of RT and
This type of charge transfer mechanism can prevent the oxidation of Cu the Cu2O Z-scheme heterostructures to achieve 7-cycle stability
if holes are not used efficiently. However, contrary to this unfavorable (Fig. 16c). They attributed the excellent stability of the heterostructures
charge transfer, the Z-scheme charge transfer mechanism is an

Fig. 16. Stability exhibited by the Z-scheme heterostructures of (a) Cu2O/g-C3N4 [137] Copyright 2020, American Chemical Society, (b) Cu2O-TiO2 [82] Copyright
2017, Elsevier, and (c) Cu2O-RT [40] Copyright 2020, Elsevier.

18
S. Ali et al. Chemical Engineering Journal 429 (2022) 131579

to the well-designed interface of Cu2O with an amorphous shell of RT. of hydrazine [73]. In addition, various other hole scavengers, such as
This electron-rich defected shell at the interface resisted the interfacial Na2SO3, can be utilized to improve the stability of Cu-based photo­
oxidation by scavenging the Cu2O holes. In addition to the favorable catalysts [73,131,141]. Fig. 18 summarizes the techniques used to
features of the Z-scheme charge transfer, its well-designed geometry maintain the stability of Cu-based photocatalysts.
utilizes the merits of both the photocatalysts by allowing them to harvest Photocatalytic CO2 reduction is always accompanied by the pro­
light and participate in the chemical reactions [40]. duction of oxygen, and Cu-based photocatalysts undergo oxidation/
deactivation upon exposure to this oxygen, as reported by Biswas et al.
4.3. Other methods and Li et al. [41,130]. Therefore, the removal of oxygen is essential and
can be realized using a continuous flow reactor. However, to the best of
A few reports are available on the utilization of noble metals in our knowledge, there is no specific study on this [127]. In contrast, for
combination with Z-scheme heterostructures to improve their stability. their Cu-TiO2 photocatalyst, Beller et al. found that the addition of a
In one such study, Xu et al. clearly demonstrated the potential of Ag to particular amount of oxygen to the feed stream helped in achieving
improve the stability of Cu2O-ZnO. The stability loss of pristine Cu2O- higher activities [140]. They found that oxygen can stabilize the active
ZnO could be avoided by adding Ag to it (Fig. 17a). They attributed this oxidation state (Cu1+) of Cu. Similar findings were reported by Crumlin
to the electron trapping ability of Ag and the Z-scheme charge transfer et al. for their study, in which oxygen facilitated the adsorption and
mechanism [76]. In addition, the VB of CuO/Cu2O can be engineered to stabilization of CO2 over the surface of Ag and Cu [27]. Therefore, it can
effectively use the photogenerated holes for water oxidation. Yu et al. be concluded that the destabilization/activation of the photocatalyst by
reported that Cl- doping improved the stability of Cu2O by lowering the oxygen depends on the photocatalytically active oxidation state of Cu.
band bending, because of which the holes present in Cu2O oxidize water
instead of oxidizing Cu2O itself. In addition, Cl- doping enhances the 5. Summary and outlook
conductivity of the photogenerated charges, which in turn transfer these
charges to the reactants for the redox reaction prior to their recombi­ The findings discussed in this review underscore the efficacious role
nation or the oxidation/reduction of Cu2O. Owing to these attributes, of Cu in improving the activity and selectivity of photocatalysts for CO2
the oxidation state and crystal structure of the Cl− -doped Cu2O, which reduction. Furthermore, the C–C coupling ability of Cu, which prevents
were investigated using XPS and XRD, respectively, did not change, and the formation of less valuable C1 products and facilitates the formation
Cu2O maintained stability for four cycles, as shown in Fig. 17b [78]. of expensive C2 + products, has opened a new avenue to overcome the
From this discussion, it can be concluded that the efficient scav­ barriers in the commercialization of Cu-based photocatalysts. The STF
enging of holes maintains the stability of Cu. In addition to using Z- efficiency of Cu-based photocatalysts has reached the landmark figure of
scheme heterostructures and precious metals, hole scavengers can be 10% for C1 chemicals, which is considered sufficient for their commer­
used to maintain the stability of Cu-based photocatalysts. The hydroxyl cialization. However, maintaining that efficiency for extended time is
group is the most common hole scavenger. Its photooxidation by holes still a challenge because of the irreversible changes in the oxidation state
also produces the necessary protons required for the photocatalytic of Cu under irradiation. Therefore, such strategies are highly desirable to
reduction reaction in addition to the important aspect of maintaining ensure the complete utilization of the photogenerated charges for the
stability [140]. Li et al. reported that the activity of a Cu-TiO2 photo­ photocatalytic CO2 reduction instead of the Cu redox reaction.
catalyst can be re-established by treating it with H2O2. This is because of The weak oxidation potential of the holes in CuO/Cu2O photo­
the additional hydroxyl groups generated by the photoirradiation of catalysts is critical for maintaining their stability, and the various stra­
H2O2 [130]. Similarly, Li et al. also demonstrated the role of hydroxyl tegies used for maintaining the stability of these photocatalysts have
groups in improving the activity of Cu-TiO2 photocatalysts [92]. Li et al. been discussed in this review. In addition, when a heterostructure is
also reported the use of methanol as a hole scavenger in improving the formed with Cu and other photocatalysts, such as TiO2, the strong •OH
stability of de-activated Cu-TiO2 [130]. Another approach to maintain radicals produced in the heterostructure also contribute to the destabi­
the stability of Cu is to use reducing agents in the photocatalytic reaction lization of the photocatalyst. In such situations, using either Cu2O or
medium. Ye et al. utilized hydrous hydrazine (N2H4⋅H2O) as the H CuO as a photocatalyst and enabling its photogenerated holes to drive
source and electron donor to prevent the oxidation of Cu. The EPR re­ water oxidation could be a potential remedy. To achieve this, the
sults showed that Cu oxidized during the photoreduction test with downward band bending of Cu2O/CuO may provide a VB with a suitable
water. On the other hand, the reduction of Cu occurred in the presence oxidation potential. Downward band bending seems to be a unique

Fig. 17. Stability profiles of (a) Ag-Cu2O-ZnO Z-scheme heterostructures [76] Copyright 2019, Elsevier and (b) Cl- -doped Cu2O [78] Copyright 2019, Elsevier.

19
S. Ali et al. Chemical Engineering Journal 429 (2022) 131579

Fig. 18. Factors affecting the stability of Cu-based photocatalysts and potential stability preservation strategies.

approach for the development of photocatalysts composed only of CuO/ the work reported in this paper.
Cu2O. Therefore, further studies should be carried out in this direction,
and the halogen doping of CuO/Cu2O can be a starting point. Further­ Acknowledgements
more, the use of hole-scavengers, especially those derived from green
processes, such as glycerol from biodiesel, with bimetallic Cu-based The authors thankfully acknowledge the support of the Ministry of
photocatalysts can further improve their stability and activity. This is Science and ICT (2017R1E1A1A01074890, 2021M3I3A1085039 and
because these hole scavengers prevent the oxidation of the photocatalyst 2021R1A2C2009459).
and solar products.
The Cu oxidation state affects the photocatalytic reduction of CO2.
References
Cu2+ facilitates the CO2 adsorption, Cu1+ stimulates the CO2 photore­
duction reaction, and Cu0 facilitates the charge separation and exhibits [1] X. Jiang, D. Guan, Determinants of global CO2 emissions growth, Appl. Energy
the LSPR response. Therefore, the suitable oxidation state of Cu is 184 (2016) 1132–1141.
[2] A. Razzaq, A. Sinhamahapatra, T.H. Kang, C.A. Grimes, J.S. Yu, S.-I. In, Efficient
indispensable for the efficient photoreduction of CO2, and reclaiming
solar light photoreduction of CO2 to hydrocarbon fuels via magnesiothermally
the photocatalyst after the photoreaction is essential for improving the reduced TiO2 photocatalyst, Appl. Catal. B Environ. 215 (2017) 28–35, https://
durability of the photocatalyst. To this end, the use of Cu SACs has doi.org/10.1016/j.apcatb.2017.05.028.
emerged as a promising approach. This is because the lost oxidation [3] P. Friedlingstein, M. Allen, J.G. Canadell, G.P. Peters, S.I. Seneviratne, Comment
on “The global tree restoration potential”, Science 366 (2019) 76–79, https://doi.
state of Cu can be easily regained by exposing the irradiated Cu SAC to org/10.1126/science.aay8060.
the ambient environment. This behavior of SACs not only provides a [4] R. Snoeckx, A. Bogaerts, Plasma technology–a novel solution for CO2 conversion?
facile approach for regeneration without requiring the removal of the Chem. Soc. Rev. 46 (19) (2017) 5805–5863.
[5] J. van Walsem, J. Roegiers, B. Modde, S. Lenaerts, S. Denys, Proof of concept of
photocatalyst from the reactor but also does not necessitate other an upscaled photocatalytic multi-tube reactor: a combined modelling and
regeneration methods, such as the use of hole scavengers and thermal experimental study, Chem. Eng. J. 378 (2019) 122038.
treatments. However, the selectivity of Cu SACs is still confined to C1 [6] R.K. de_Richter, T. Ming, S. Caillol, Fighting global warming by photocatalytic
reduction of CO2 using giant photocatalytic reactors, Renew. Sustain. Energy Rev.
products, and we believe that it can be tuned to C2 + products by 19 (2013) 82–106.
designing bimetallic SACs. The intermediate stabilization effect of Cu in [7] C.B. Hiragond, J. Lee, H. Kim, J.-W. Jung, C.-H. Cho, S.-I. In, A novel N-doped
conjunction with the abundant electrons on the adjacent bimetallic site, graphene oxide enfolded reduced titania for highly stable and selective gas-phase
photocatalytic CO2 reduction into CH4: An in-depth study on the interfacial
that is Pt, ensures the formation of C2 + products.
charge transfer mechanism, Chem. Eng. J. 416 (2021) 127978, https://doi.org/
The LSPR effect of the Cu-based photocatalysts is also very impor­ 10.1016/j.cej.2020.127978.
tant, as plasmonic electrons can be transferred to the CB of the photo­ [8] A.K. Singh, J.H. Montoya, J.M. Gregoire, K.A. Persson, Robust and synthesizable
photocatalysts for CO2 reduction: a data-driven materials discovery, Nat.
catalyst or to the reactants/intermediates formed during the
Commun. 10 (2019) 443.
photocatalytic CO2 reduction. Therefore, a higher population of elec­ [9] P.V. Danckwerts, Z-Scheme Photocatalytic Systems for Carbon Dioxide
trons can be achieved in the visible or NIR region for the efficient Reduction: Where Are We Now? Chem. Eng. Sci. 17 (1962) 955, https://doi.org/
photocatalytic reduction of CO2. In addition, the thermal effects induced 10.1016/0009-2509(62)87032-8.
[10] M. Khalil, J. Gunlazuardi, T.A. Ivandini, A. Umar, Photocatalytic conversion of
by the LSPR effect also contribute to enhancing the CO2 reduction. Apart CO2 using earth-abundant catalysts: a review on mechanism and catalytic
from stimulating the photocatalytic reaction, the LSPR effect also gen­ performance, Renew. Sustain. Energy Rev. 113 (2019) 109246, https://doi.org/
erates reaction sites for C–C coupling. Therefore, designing plasmonic 10.1016/j.rser.2019.109246.
[11] M. Marx, A. Mele, A. Spannenberg, C. Steinlechner, H. Junge, P. Schollhammer,
Cu photocatalysts can help in achieving higher C2 + production rates. M. Beller, Addressing the reproducibility of photocatalytic carbon dioxide
Perceiving the current trends of overwhelming interest in the Cu-based reduction, ChemCatChem 12 (6) (2020) 1603–1608.
photocatalysts, we anticipate that these will be pioneer materials for [12] S. Ali, A. Razzaq, S.-I. In, Development of graphene based photocatalysts for CO2
reduction to C1 chemicals: A brief overview, Catal. Today 335 (2018) 39–54. doi:
commercial-scale applications. 10.1016/ j.cattod.2018.12.003.
[13] S.-H. Li, M.-Y. Qi, Y.-Y. Fan, Y. Yang, M. Anpo, Y.M.A. Yamada, Z.-R. Tang, Y.-
J. Xu, Modulating photon harvesting through dynamic non-covalent interactions
Declaration of Competing Interest for enhanced photochemical CO2 reduction, Appl. Catal. B Environ. 292 (2021)
120157.
[14] Y.-H. Chen, M.-Y. Qi, Y.-H. Li, Z.-R. Tang, T. Wang, J. Gong, Y.-J. Xu, Activating
The authors declare that they have no known competing financial two-dimensional Ti3C2Tx-MXene with single-atom cobalt for efficient CO2
interests or personal relationships that could have appeared to influence photoreduction, Cell Rep. Phys. Sci. 2 (2021) 100371.

20
S. Ali et al. Chemical Engineering Journal 429 (2022) 131579

[15] K.-Q. Lu, Y.-H. Li, F. Zhang, M.-Y. Qi, X. Chen, Z.-R. Tang, Y.M.A. Yamada, [38] Q. Xu, L. Zhang, J. Yu, S. Wageh, A.A. Al-Ghamdi, M. Jaroniec, Direct Z-scheme
M. Anpo, M. Conte, Y.-J. Xu, Rationally designed transition metal hydroxide photocatalysts: principles, synthesis, and applications, Mater. Today 21 (10)
nanosheet arrays on graphene for artificial CO2 reduction, Nat. Commun. 11 (2018) 1042–1063, https://doi.org/10.1016/j.mattod.2018.04.008.
(2020) 5181. [39] J. Low, C. Jiang, B. Cheng, S. Wageh, A.A. Al-Ghamdi, J. Yu, A review of direct Z-
[16] Y. Zhang, B. Xia, J. Ran, K. Davey, S.Z. Qiao, Atomic-level reactive sites for scheme photocatalysts, Small Methods 1 (2017) 1700080.
semiconductor-based photocatalytic CO2 reduction, Adv. Energy Mater. 10 [40] S. Ali, J. Lee, H. Kim, Y. Hwang, A. Razzaq, J.-W. Jung, C.-H. Cho, S.-I. In,
(2020) 1903879. Sustained, photocatalytic CO2 reduction to CH4 in a continuous flow reactor by
[17] S. Sorcar, S. Yoriya, H. Lee, C.A. Grimes, S.P. Feng, A review of recent progress in earth-abundant materials: reduced titania-Cu2O Z-scheme heterostructures, Appl.
gas phase CO2 reduction and suggestions on future advancement, Mater. Today Catal. B Environ. 279 (2020) 119344, https://doi.org/10.1016/j.
Chem. 16 (2020) 100264. apcatb.2020.119344.
[18] S. Sorcar, Y. Hwang, J. Lee, H. Kim, K.M. Grimes, C.A. Grimes, J.-W. Jung, C.- [41] Y. Li, W.-N. Wang, Z. Zhan, M.-H. Woo, C.-Y. Wu, P. Biswas, Photocatalytic
H. Cho, T. Majima, M.R. Hoffmann, S.-I. In, CO2, water, and sunlight to reduction of CO2 with H2O on mesoporous silica supported Cu/TiO2 catalysts,
hydrocarbon fuels: a sustained sunlight to fuel (Joule-to-Joule) photoconversion Appl. Catal. B Environ. 100 (1-2) (2010) 386–392, https://doi.org/10.1016/j.
efficiency of 1%, Energy Environ. Sci. 12 (9) (2019) 2685–2696, https://doi.org/ apcatb.2010.08.015.
10.1039/C9EE00734B. [42] B.-H. Lee, S. Park, M. Kim, A.K. Sinha, S.C. Lee, E. Jung, W.J. Chang, K.-S. Lee, J.
[19] Y.A. Wu, I. McNulty, C. Liu, K.C. Lau, Q.i. Liu, A.P. Paulikas, C.-J. Sun, Z. Cai, J. H. Kim, S.-P. Cho, H. Kim, K.T. Nam, T. Hyeon, Reversible and cooperative
R. Guest, Y. Ren, V. Stamenkovic, L.A. Curtiss, Y. Liu, T. Rajh, Facet-dependent photoactivation of single-atom Cu/TiO2 photocatalysts, Nat. Mater. 18 (6) (2019)
active sites of a single Cu2O particle photocatalyst for CO2 reduction to methanol, 620–626, https://doi.org/10.1038/s41563-019-0344-1.
Nat. Energy 4 (11) (2019) 957–968, https://doi.org/10.1038/s41560-019-0490- [43] Z. Jiang, W. Sun, W. Miao, Z. Yuan, G. Yang, F. Kong, T. Yan, J. Chen, B. Huang,
3. C. An, G.A. Ozin, Living atomically dispersed Cu ultrathin TiO2 nanosheet CO2
[20] S. Ali, M.C. Flores, A. Razzaq, S. Sorcar, C.B. Hiragond, H.R. Kim, Y.H. Park, reduction photocatalyst, Adv. Sci. 6 (15) (2019) 1900289, https://doi.org/
Y. Hwang, H.S. Kim, H. Kim, E.H. Gong, J. Lee, D. Kim, S.-I. In, Gas Phase 10.1002/advs.v6.1510.1002/advs.201900289.
Photocatalytic CO2 Reduction, “A Brief Overview for Benchmarking”, Catalysts 9 [44] F. Zhang, Y.-H. Li, M.-Y. Qi, Y.M.A. Yamada, M. Anpo, Z.-R. Tang, Y.-J. Xu,
(2019) 727. Photothermal catalytic CO2 reduction over nanomaterials, Chem. Catal. 1 (2)
[21] J. Albero, Y. Peng, H. García, Photocatalytic CO2 reduction to C2+ products, ACS (2021) 272–297, https://doi.org/10.1016/j.checat.2021.01.003.
Catal. 10 (10) (2020) 5734–5749. [45] J.-Y. Li, L. Yuan, S.-H. Li, Z.-R. Tang, Y.-J. Xu, One-dimensional copper-based
[22] T. Wu, C. Zhu, D. Han, Z. Kang, L.i. Niu, Highly selective conversion of CO2 to heterostructures toward photo-driven reduction of CO2 to sustainable fuels and
C2H6 on graphene modified chlorophyll Cu through multi-electron process for feedstocks, J. Mater. Chem. A 7 (15) (2019) 8676–8689.
artificial photosynthesis, Nanoscale 11 (47) (2019) 22980–22988. [46] H. Xie, J. Wang, K. Ithisuphalap, G. Wu, Q. Li, Recent advances in Cu-based
[23] A. Loiudice, P. Lobaccaro, E.A. Kamali, T. Thao, B.H. Huang, J.W. Ager, nanocomposite photocatalysts for CO2 conversion to solar fuels, J. Energy Chem.
R. Buonsanti, Tailoring copper nanocrystals towards C2 products in 26 (6) (2017) 1039–1049.
electrochemical CO2 reduction, Angew. Chemie Int. Ed. 55 (19) (2016) [47] S.P. Meshram, P.V. Adhyapak, U.P. Mulik, D.P. Amalnerkar, Facile synthesis of
5789–5792. CuO nanomorphs and their morphology dependent sunlight driven photocatalytic
[24] W. Wang, C. Deng, S. Xie, Y. Li, W. Zhang, H. Sheng, C. Chen, J. Zhao, properties, Chem. Eng. J. 204-206 (2012) 158–168.
Photocatalytic C− C coupling from carbon dioxide reduction on copper oxide with [48] H. Xu, W. Wang, W. Zhu, Shape evolution and size-controllable synthesis of Cu2O
mixed-valence copper(I)/copper(II), J. Am. Chem. Soc. 143 (7) (2021) octahedra and their morphology-dependent photocatalytic properties, J. Phys.
2984–2993, https://doi.org/10.1021/jacs.1c0020610.1021/jacs.1c00206.s001. Chem. B. 110 (28) (2006) 13829–13834.
[25] M.B. Gawande, A. Goswami, F.-X. Felpin, T. Asefa, X. Huang, R. Silva, X. Zou, [49] S.-M. Park, A. Razzaq, Y.H. Park, S. Sorcar, Y. Park, C.A. Grimes, S.-I. In, Hybrid
R. Zboril, R.S. Varma, Cu and Cu-based nanoparticles: synthesis and applications CuxO–TiO2 heterostructured composites for photocatalytic CO2 reduction into
in catalysis, Chem. Rev. 116 (6) (2016) 3722–3811, https://doi.org/10.1021/acs. methane using solar irradiation: sunlight into fuel, ACS Omega 1 (5) (2016)
chemrev.5b00482. 868–875.
[26] J. Graciani, K. Mudiyanselage, F. Xu, A.E. Baber, J. Evans, S.D. Senanayake, D. [50] J. Yu, S. Zhuang, X. Xu, W. Zhu, B. Feng, J. Hu, Photogenerated electron reservoir
J. Stacchiola, P. Liu, J. Hrbek, J.F. Sanz, J.A. Rodriguez, Highly active copper- in hetero-p–n CuO–ZnO nanocomposite device for visible-light-driven
ceria and copper-ceria-titania catalysts for methanol synthesis from CO2, Science photocatalytic reduction of aqueous Cr(VI), J. Mater. Chem. A 3 (2015)
345 (6196) (2014) 546–550. 1199–1207.
[27] Y. Ye, H. Yang, J. Qian, H. Su, K.J. Lee, T. Cheng, H. Xiao, J. Yano, W.A. Goddard, [51] T. Sun, E. Source, Chapter 2 Solar radiation at the top of the atmosphere, Int.
E.J. Crumlin, Dramatic differences in carbon dioxide adsorption and initial steps Geophys. 84 (2002) 37–64. doi:10.1016/S0074-6142(02)80017-1.
of reduction between silver and copper, Nat. Commun. 10 (2019) 1875, https:// [52] M.D. Hernández-Alonso, F. Fresno, S. Suárez, J.M. Coronado, Development of
doi.org/10.1038/s41467-019-09846-y. alternative photocatalysts to TiO2: challenges and opportunities, Energy Environ.
[28] K. Teramura, K. Hori, Y. Terao, Z. Huang, S. Iguchi, Z. Wang, H. Asakura, Sci. 2 (12) (2009) 1231, https://doi.org/10.1039/b907933e.
S. Hosokawa, T. Tanaka, Which is an intermediate species for photocatalytic [53] Z. Lou, Z. Wang, B. Huang, Y. Dai, Synthesis and activity of plasmonic
conversion of CO2 by H2O as the electron donor: CO2 molecule, carbonic acid, photocatalysts, ChemCatChem 6 (9) (2014) 2456–2476.
bicarbonate, or carbonate ions? J. Phys. Chem. C 121 (16) (2017) 8711–8721, [54] M.M. Abouelela, G. Kawamura, A. Matsuda, A review on plasmonic nanoparticle-
https://doi.org/10.1021/acs.jpcc.6b1280910.1021/acs.jpcc.6b12809.s001. semiconductor photocatalysts for water splitting, J. Clean. Prod. 294 (2021)
[29] A. Gankanda, D.M. Cwiertny, V.H. Grassian, Role of atmospheric CO2 and H2O 126200.
adsorption on ZnO and CuO nanoparticle aging: formation of new surface phases [55] P. Zhang, G. Zeng, T. Song, S. Huang, T. Wang, H. Zeng, Synthesis of a plasmonic
and the impact on nanoparticle dissolution, J. Phys. Chem. C 120 (34) (2016) CuNi bimetal modified with carbon quantum dots as a non-semiconductor-driven
19195–19203, https://doi.org/10.1021/acs.jpcc.6b0593110.1021/acs. photocatalyst for effective water splitting, J. Catal. 369 (2019) 267–275.
jpcc.6b05931.s001. [56] P. Zheng, H. Tang, B. Liu, S. Kasani, L. Huang, N. Wu, Origin of strong and narrow
[30] K.C. Christoforidis, P. Fornasiero, Photocatalysis for hydrogen production and localized surface plasmon resonance of copper nanocubes, Nano Res. 12 (1)
CO2 reduction: the case of copper-catalysts, ChemCatChem 11 (1) (2019) (2019) 63–68.
368–382, https://doi.org/10.1002/cctc.v11.110.1002/cctc.201801198. [57] Z.-J. Wang, H. Song, H. Pang, Y. Ning, T.D. Dao, Z. Wang, H. Chen, Y. Weng,
[31] J. Zhu, G. Cheng, J. Xiong, W. Li, S. Dou, Recent advances in Cu-based cocatalysts Q. Fu, T. Nagao, Y. Fang, J. Ye, Photo-assisted methanol synthesis via CO2
toward solar-to-hydrogen evolution: categories and roles, Sol. RRL 3 (10) (2019) reduction under ambient pressure over plasmonic Cu/ZnO catalysts, Appl. Catal.
1900256, https://doi.org/10.1002/solr.v3.1010.1002/solr.201900256. B Environ. 250 (2019) 10–16.
[32] C.Y. Toe, Z. Zheng, H. Wu, J. Scott, R. Amal, Y.H. Ng, Photocorrosion of cuprous [58] J.S. DuChene, B.C. Sweeny, A.C. Johnston-Peck, D. Su, E.A. Stach, W.D. Wei,
oxide in hydrogen production: rationalising self-oxidation or self-reduction, Prolonged hot electron dynamics in plasmonic-metal/semiconductor
Angew. Chemie Int. Ed. 57 (2018) 13613–13617. heterostructures with implications for solar photocatalysis, Angew. Chemie Int.
[33] S.J. Hoseini, R.H. Fath, Formation of nanoneedle Cu(0)/CuS nanohybrid thin film Ed. 53 (30) (2014) 7887–7891.
by the disproportionation of a copper(i) complex at an oil–water interface and its [59] K. Wu, J. Chen, J.R. McBride, T. Lian, Efficient hot-electron transfer by a
application for dye degradation, RSC Adv. 6 (2016) 76964–76971. plasmon-induced interfacial charge-transfer transition, Science 349 (6248)
[34] Y. Lou, Y. Zhang, L. Cheng, J. Chen, Y. Zhao, A stable plasmonic Cu@ Cu2O/ZnO (2015) 632–635.
heterojunction for enhanced photocatalytic hydrogen generation, ChemSusChem [60] M. Sayed, L. Zhang, J. Yu, Plasmon-induced interfacial charge-transfer transition
11 (9) (2018) 1505–1511. prompts enhanced CO2 photoreduction over Cu/Cu2O octahedrons, Chem. Eng. J.
[35] M. Alvaro, E. Carbonell, V. Fornés, H. García, Enhanced photocatalytic activity of 397 (2020) 125390.
zeolite-encapsulated TiO2 clusters by complexation with organic additives and N- [61] O.K. Varghese, M. Paulose, T.J. LaTempa, C.A. Grimes, High-rate solar
doping, ChemPhysChem 7 (1) (2006) 200–205. photocatalytic conversion of CO2 and water vapor to hydrocarbon fuels, Nano
[36] S.D. Pike, E.R. White, A. Regoutz, N. Sammy, D.J. Payne, C.K. Williams, M.S. Lett. 9 (2) (2009) 731–737.
P. Shaffer, Reversible redox cycling of well-defined, ultrasmall Cu/Cu2O [62] Y. Xin, K. Yu, L. Zhang, Y. Yang, H. Yuan, H. Li, L. Wang, J. Zeng, Copper-based
nanoparticles, ACS Nano 11 (3) (2017) 2714–2723, https://doi.org/10.1021/ plasmonic catalysis: recent advances and future perspectives, Adv. Mater. 33 (32)
acsnano.6b0769410.1021/acsnano.6b07694.s001. (2021) 2008145, https://doi.org/10.1002/adma.v33.3210.1002/
[37] Q.i. Wang, Z. Zhao, Y. Jia, M. Wang, W. Qi, Y. Pang, J. Yi, Y. Zhang, Z. Li, adma.202008145.
Z. Zhang, Unique Cu@CuPt core–shell concave octahedron with enhanced [63] G.-Y. Yao, Z.-Y. Zhao, Unraveling the role of cuprous oxide and boosting solar
methanol oxidation activity, ACS Appl. Mater. Interfaces 9 (42) (2017) energy conversion: via interface engineering in a Cu/TiO2 plasmonic
36817–36827. photocatalyst, J. Mater. Chem. C 8 (25) (2020) 8567–8578, https://doi.org/
10.1039/D0TC01127D.

21
S. Ali et al. Chemical Engineering Journal 429 (2022) 131579

[64] C. Nethravathi, R.N. R, J.T. Rajamathi, M. Rajamathi, Microwave-assisted selective sites for photocatalytic conversion of CO2 to CH4, J. Am. Chem. Soc. 139
synthesis of porous aggregates of CuS nanoparticles for sunlight photocatalysis, (12) (2017) 4486–4492, https://doi.org/10.1021/jacs.7b0045210.1021/
ACS Omega 4 (3) (2019) 4825–4831. jacs.7b00452.s001.
[65] X. Li, L. Liang, Y. Sun, J. Xu, X. Jiao, X. Xu, H. Ju, Y. Pan, J. Zhu, Y.i. Xie, [90] S. Docao, A.R. Koirala, M.G. Kim, I.C. Hwang, M.K. Song, K.B. Yoon, Solar
Ultrathin conductor enabling efficient IR light CO2 reduction, J. Am. Chem. Soc. photochemical-thermal water splitting at 140◦ C with Cu-loaded TiO2, Energy
141 (1) (2019) 423–430, https://doi.org/10.1021/jacs.8b1069210.1021/ Environ. Sci. 10 (2) (2017) 628–640, https://doi.org/10.1039/C6EE02974D.
jacs.8b10692.s001. [91] N. Vu, S. Kaliaguine, T. Do, Plasmonic photocatalysts for sunlight-driven
[66] L. Jiang, K. Wang, X. Wu, G. Zhang, Highly enhanced full solar spectrum-driven reduction of CO2: details, developments, and perspectives, ChemSusChem 13
photocatalytic CO2 reduction performance in Cu2–xS/g-C3N4 composite: efficient (2020) 3967–3991.
charge transfer and mechanism insight, Sol. RRL 5 (2) (2021) 2000326, https:// [92] S. Yu, A.J. Wilson, J. Heo, P.K. Jain, Plasmonic control of multi-electron transfer
doi.org/10.1002/solr.v5.210.1002/solr.202000326. and C-C coupling in visible-light-driven CO2 reduction on Au nanoparticles, Nano
[67] H. Xu, S. Ouyang, L. Liu, D. Wang, T. Kako, J. Ye, Porous-structured Cu2O/TiO2 Lett. 18 (4) (2018) 2189–2194.
nanojunction material toward efficient CO2 photoreduction, Nanotechnology 25 [93] S. Zhu, X. Chen, Z. Li, X. Ye, Y. Liu, Y. Chen, L.i. Yang, M. Chen, D. Zhang, G. Li,
(2014) 165402. H. Li, Cooperation between inside and outside of TiO2: Lattice Cu+ accelerates
[68] K.L. Bae, J. Kim, C.K. Lim, K.M. Nam, H. Song, Colloidal zinc oxide-copper(I) carrier migration to the surface of metal copper for photocatalytic CO2 reduction,
oxide nanocatalysts for selective aqueous photocatalytic carbon dioxide Appl. Catal. B Environ. 264 (2020) 118515, https://doi.org/10.1016/j.
conversion into methane, Nat. Commun. 8 (2017) 1156, https://doi.org/ apcatb.2019.118515.
10.1038/s41467-017-01165-4. [94] A. Wang, J. Li, T. Zhang, Heterogeneous single-atom catalysis, Nat. Rev. Chem. 2
[69] K. Wang, R. Jiang, T. Peng, X. Chen, W. Dai, X. Fu, Modeling the effect of Cu (6) (2018) 65–81.
doped TiO2 with carbon dots on CO2 methanation by H2O in a photo-thermal [95] J. Wang, T. Heil, B. Zhu, C.-W. Tung, J. Yu, H.M. Chen, M. Antonietti, S. Cao,
system, Appl. Catal. B Environ. 256 (2019) 117780. A single Cu-center containing enzyme-mimic enabling full photosynthesis under
[70] G. Colón, M. Maicu, M.C. Hidalgo, J.A. Navío, Cu-doped TiO2 systems with CO2 reduction, ACS Nano 14 (7) (2020) 8584–8593, https://doi.org/10.1021/
improved photocatalytic activity, Appl. Catal. B Environ. 67 (1-2) (2006) 41–51, acsnano.0c0294010.1021/acsnano.0c02940.s001.
https://doi.org/10.1016/j.apcatb.2006.03.019. [96] M. Wang, M. Shen, X. Jin, J. Tian, M. Li, Y. Zhou, L. Zhang, Y. Li, J. Shi, Oxygen
[71] M.-C. Wu, P.-Y. Wu, T.-H. Lin, T.-F. Lin, Photocatalytic performance of Cu-doped vacancy generation and stabilization in CeO2-x by Cu introduction with improved
TiO2 nanofibers treated by the hydrothermal synthesis and air-thermal treatment, CO2 photocatalytic reduction activity, ACS Catal. 9 (5) (2019) 4573–4581,
Appl. Surf. Sci. 430 (2018) 390–398. https://doi.org/10.1021/acscatal.8b0397510.1021/acscatal.8b03975.s001.
[72] B.-R. Chen, V.-H. Nguyen, J.C.S. Wu, R. Martin, K. Kočí, Production of renewable [97] Z. Xiong, Z. Xu, Y. Li, L. Dong, J. Wang, J. Zhao, X. Chen, Y. Zhao, H. Zhao,
fuels by the photohydrogenation of CO2: Effect of the Cu species loaded onto TiO2 J. Zhang, Incorporating highly dispersed and stable Cu+ into TiO2 lattice for
photocatalysts, Phys. Chem. Chem. Phys. 18 (6) (2016) 4942–4951, https://doi. enhanced photocatalytic CO2 reduction with water, Appl. Surf. Sci. 507 (2020)
org/10.1039/C5CP06999H. 145095, https://doi.org/10.1016/j.apsusc.2019.145095.
[73] I. Shown, H.-C. Hsu, Y.-C. Chang, C.-H. Lin, P.K. Roy, A. Ganguly, C.-H. Wang, J.- [98] Z. Xiong, Z. Lei, C.C. Kuang, X. Chen, B. Gong, Y. Zhao, J. Zhang, C. Zheng, J.C.
K. Chang, C.-I. Wu, L.-C. Chen, K.-H. Chen, Highly efficient visible light S. Wu, Selective photocatalytic reduction of CO2 into CH4 over Pt-Cu2O TiO2
photocatalytic reduction of CO2 to hydrocarbon fuels by cu-nanoparticle nanocrystals: the interaction between Pt and Cu2O cocatalysts, Appl. Catal. B
decorated graphene oxide, Nano Lett. 14 (11) (2014) 6097–6103, https://doi. Environ. 202 (2017) 695–703, https://doi.org/10.1016/j.apcatb.2016.10.001.
org/10.1021/nl503609v. [99] D. Ferrah, A.R. Haines, R.P. Galhenage, J.P. Bruce, A.D. Babore, A. Hunt,
[74] Q. Kang, T. Wang, P. Li, L. Liu, K. Chang, M.u. Li, J. Ye, Photocatalytic reduction I. Waluyo, J.C. Hemminger, Wet chemical growth and thermocatalytic activity of
of carbon dioxide by hydrous hydrazine over Au-Cu alloy nanoparticles supported Cu-based nanoparticles supported on TiO2 nanoparticles/HOPG. In situ ambient
on SrTiO3/TiO2 coaxial nanotube arrays, Angew. Chemie. 127 (3) (2015) pressure XPS study of the CO2 hydrogenation reaction, ACS Catal. 9 (8) (2019)
855–859, https://doi.org/10.1002/ange.201409183. 6783–6802, https://doi.org/10.1021/acscatal.9b0141910.1021/
[75] G. Li, N.M. Dimitrijevic, L.e. Chen, T. Rajh, K.A. Gray, Role of surface/interfacial acscatal.9b01419.s001.
Cu2+ sites in the photocatalytic activity of coupled CuO-TiO2 nanocomposites, [100] L. Liu, F. Gao, H. Zhao, Y. Li, Tailoring Cu valence and oxygen vacancy in Cu/
J. Phys. Chem. C 112 (48) (2008) 19040–19044, https://doi.org/10.1021/ TiO2 catalysts for enhanced CO2 photoreduction efficiency, Appl. Catal. B
jp8068392. Environ. 134-135 (2013) 349–358.
[76] E. Kalamaras, M.M. Maroto-Valer, J.M. Andresen, H. Wang, J. Xuan, [101] S. Zhu, S. Liang, Y. Tong, X. An, J. Long, X. Fu, X. Wang, Photocatalytic reduction
Thermodynamic analysis of the efficiency of photoelectrochemical CO2 reduction of CO2 with H2O to CH4 on Cu(i) supported TiO2 nanosheets with defective 001
to ethanol, Energy Procedia 158 (2019) 767–772. facets, Phys. Chem. Chem. Phys. 17 (2015) 9761–9770, https://doi.org/10.1039/
[77] F. Zhang, Y.-H. Li, M.-Y. Qi, Z.-R. Tang, Y.-J. Xu, Boosting the activity and c5cp00647c.
stability of Ag-Cu2O/ZnO nanorods for photocatalytic CO2 reduction, Appl. Catal. [102] I.-H. Tseng, J.-S. Wu, Chemical states of metal-loaded titania in the
B Environ. 268 (2020) 118380, https://doi.org/10.1016/j.apcatb.2019.118380. photoreduction of CO2, Catal. Today 97 (2-3) (2004) 113–119, https://doi.org/
[78] A. Kudo, Y. Miseki, Heterogeneous photocatalyst materials for water splitting, 10.1016/j.cattod.2004.03.063.
Chem. Soc. Rev. 38 (1) (2009) 253–278. [103] C. Chen, X. Sun, X. Yan, Y. Wu, M. Liu, S. Liu, Z. Zhao, B. Han, A strategy to
[79] L. Yu, X. Ba, M. Qiu, Y. Li, L. Shuai, W. Zhang, Z. Ren, Y. Yu, Visible-light driven control the grain boundary density and Cu+/Cu0 ratio of Cu-based catalysts for
CO2 reduction coupled with water oxidation on Cl-doped Cu2O nanorods, Nano efficient electroreduction of CO2 to C2 products, Green Chem. 22 (5) (2020)
Energy 60 (2019) 576–582, https://doi.org/10.1016/j.nanoen.2019.03.083. 1572–1576.
[80] A.D. Handoko, J. Tang, Controllable proton and CO2 photoreduction over Cu2O [104] S.C. Roy, O.K. Varghese, M. Paulose, C.A. Grimes, Toward solar fuels:
with various morphologies, Int. J. Hydrogen Energy 38 (29) (2013) Photocatalytic conversion of carbon dioxide to hydrocarbons, ACS Nano 4 (3)
13017–13022. (2010) 1259–1278, https://doi.org/10.1021/nn9015423.
[81] W. Chen, Y. Wang, S. Liu, L.i. Gao, L. Mao, Z. Fan, W. Shangguan, Z. Jiang, Non- [105] J. Fu, K. Jiang, X. Qiu, J. Yu, M. Liu, Product selectivity of photocatalytic CO2
noble metal Cu as a cocatalyst on TiO2 nanorod for highly efficient photocatalytic reduction reactions, Mater. Today 32 (2020) 222–243.
hydrogen production, Appl. Surf. Sci. 445 (2018) 527–534. [106] X. Zhang, X. Li, D. Zhang, N.Q. Su, W. Yang, H.O. Everitt, J. Liu, Product
[82] P. Wang, Y. Xia, P. Wu, X. Wang, H. Yu, J. Yu, Cu(II) as a general cocatalyst for selectivity in plasmonic photocatalysis for carbon dioxide hydrogenation, Nat.
improved visible-light photocatalytic performance of photosensitive Ag-based Commun. 8 (2017) 14542.
compounds, J. Phys. Chem. C 118 (17) (2014) 8891–8898. [107] C. Dong, M. Xing, J. Zhang, Economic hydrophobicity triggering of CO2
[83] M.E. Aguirre, R. Zhou, A.J. Eugene, M.I. Guzman, M.A. Grela, Cu2O/TiO2 photoreduction for selective CH4 generation on noble-metal-free TiO2–SiO2,
heterostructures for CO2 reduction through a direct Z-scheme: Protecting Cu2O J. Phys. Chem. Lett. 7 (15) (2016) 2962–2966.
from photocorrosion, Appl. Catal. B Environ. 217 (2017) 485–493. [108] C. Hiragond, S. Ali, S. Sorcar, S.-I. In, Hierarchical nanostructured photocatalysts
[84] P. Zhou, J. Yu, M. Jaroniec, All-solid-state Z-scheme photocatalytic systems, Adv. for CO2 photoreduction, Catalysts. 9 (2019) 370.
Mater. 26 (29) (2014) 4920–4935, https://doi.org/10.1002/adma.201400288. [109] Y. Lan, Y. Xie, J. Chen, Z. Hu, D. Cui, Selective photocatalytic CO2 reduction on
[85] Z. Zeng, Y. Yan, J. Chen, P. Zan, Q. Tian, P. Chen, Boosting the photocatalytic copper–titanium dioxide: a study of the relationship between CO production and
ability of Cu2O nanowires for CO2 conversion by MXene quantum dots, Adv. H2 suppression, Chem. Commun. 55 (2019) 8068–8071.
Funct. Mater. 29 (2) (2019) 1806500, https://doi.org/10.1002/adfm. [110] T. Kong, Y. Jiang, Y. Xiong, Photocatalytic CO2 conversion: what can we learn
v29.210.1002/adfm.201806500. from conventional COx hydrogenation? Chem. Soc. Rev. 49 (2020) 6579–6591.
[86] J.-C. Wang, L. Zhang, W.-X. Fang, J. Ren, Y.-Y. Li, H.-C. Yao, J.-S. Wang, Z.-J. Li, [111] X. Wang, L. Lu, B. Wang, Z. Xu, Z. Xin, S. Yan, Z. Geng, Z. Zou, Frustrated Lewis
Enhanced photoreduction CO2 activity over direct Z-Scheme α-Fe2O3/Cu2O Pairs Accelerating CO2 Reduction on Oxyhydroxide Photocatalysts with Surface
heterostructures under visible light irradiation, ACS Appl. Mater. Interfaces 7 (16) Lattice Hydroxyls as a Solid-State Proton Donor, Adv. Funct. Mater. 28 (2018)
(2015) 8631–8639, https://doi.org/10.1021/acsami.5b00822. 1–9. 1804191 doi:10.1002/adfm.201804191.s.
[87] L. Liu, C. Zhao, Y. Li, Spontaneous dissociation of CO2 to CO on defective surface [112] Y. Pu, Y. Luo, X. Wei, J. Sun, L. Li, W. Zou, L. Dong, Synergistic effects of Cu2O-
of Cu(I)/TiO2-x nanoparticles at room temperature, J. Phys. Chem. C. 116 (14) decorated CeO2 on photocatalytic CO2 reduction: Surface Lewis acid/base and
(2012) 7904–7912, https://doi.org/10.1021/jp300932b. oxygen defect, Appl. Catal. B Environ. 254 (2019) 580–586, https://doi.org/
[88] L. Yuan, S.-F. Hung, Z.-R. Tang, H.M. Chen, Y. Xiong, Y.-J. Xu, Dynamic evolution 10.1016/j.apcatb.2019.04.093.
of atomically dispersed Cu species for CO2 photoreduction to solar fuels, ACS [113] W.N.R.W. Isahak, Z.A.C. Ramli, M.W. Ismail, K. Ismail, R.M. Yusop, M.W.
Catal. 9 (6) (2019) 4824–4833, https://doi.org/10.1021/ M. Hisham, M.A. Yarmo, Adsorption-desorption of CO2 on different type of
acscatal.9b0086210.1021/acscatal.9b00862.s001. copper oxides surfaces: Physical and chemical attractions studies, J. CO2 Util. 2
[89] R. Long, Y.u. Li, Y. Liu, S. Chen, X. Zheng, C. Gao, C. He, N. Chen, Z. Qi, L.i. Song, (2013) 8–15, https://doi.org/10.1016/j.jcou.2013.06.002.
J. Jiang, J. Zhu, Y. Xiong, Isolation of Cu atoms in Pd lattice: forming highly

22
S. Ali et al. Chemical Engineering Journal 429 (2022) 131579

[114] N. Li, B. Wang, Y. Si, F. Xue, J. Zhou, Y. Lu, M. Liu, Toward High-Value [129] N. Li, X. Liu, J. Zhou, W. Chen, M. Liu, Encapsulating CuO quantum dots in MIL-
Hydrocarbon Generation by Photocatalytic Reduction of CO2 in Water Vapor, ACS 125(Ti) coupled with g-C3N4 for efficient photocatalytic CO2 reduction, Chem.
Catal. 9 (6) (2019) 5590–5602, https://doi.org/10.1021/ Eng. J. 399 (2020) 125782, https://doi.org/10.1016/j.cej.2020.125782.
acscatal.9b0022310.1021/acscatal.9b00223.s001. [130] H. Ge, B. Zhang, H. Liang, M. Zhang, K. Fang, Y. Chen, Y. Qin, Photocatalytic
[115] S.J. Datta, C. Khumnoon, Z.H. Lee, W.K. Moon, S. Docao, T.H. Nguyen, I. conversion of CO2 into light olefins over TiO2 nanotube confined Cu clusters with
C. Hwang, D. Moon, P. Oleynikov, O. Terasaki, K.B. Yoon, CO2 capture from high ratio of Cu+, Appl. Catal. B Environ. 263 (2020) 118133, https://doi.org/
humid flue gases and humid atmosphere using a microporous coppersilicate, 10.1016/j.apcatb.2019.118133.
Science. 350 (6258) (2015) 302–306. [131] L. Liu, C. Zhao, J.T. Miller, Y. Li, Mechanistic study of CO2 photoreduction with
[116] L. Yuan, K.Q. Lu, F. Zhang, X. Fu, Y.J. Xu, Unveiling the interplay between light- H2O on Cu/TiO2 nanocomposites by in situ x-ray absorption and infrared
driven CO2 photocatalytic reduction and carbonaceous residues decomposition: A spectroscopies, J. Phys. Chem. C 121 (1) (2017) 490–499.
case study of Bi2WO6-TiO2 binanosheets, Appl. Catal. B Environ. 237 (2018) [132] F. Li, L.i. Zhang, J. Tong, Y. Liu, S. Xu, Y. Cao, S. Cao, Photocatalytic CO2
424–431, https://doi.org/10.1016/j.apcatb.2018.06.019. conversion to methanol by Cu2O/graphene/TNA heterostructure catalyst in a
[117] H. Zhang, T. Itoi, T. Konishi, Y. Izumi, Dual Photocatalytic Roles of Light: Charge visible-light-driven dual-chamber reactor, Nano Energy 27 (2016) 320–329.
Separation at the Band Gap and Heat via Localized Surface Plasmon Resonance to [133] X. An, K. Li, J. Tang, Cu2O/reduced graphene oxide composites for the
Convert CO2 into CO over Silver-Zirconium Oxide, J. Am. Chem. Soc. 141 (15) photocatalytic conversion of CO2, ChemSusChem 7 (4) (2014) 1086–1093.
(2019) 6292–6301, https://doi.org/10.1021/jacs.8b1389410.1021/ [134] H. Li, X. Zhang, D.R. MacFarlane, Carbon quantum dots/Cu2O heterostructures
jacs.8b13894.s001. for solar-light-driven conversion of CO2 to methanol, Adv. Energy Mater. 5 (5)
[118] X. Zhang, F. Han, B.o. Shi, S. Farsinezhad, G.P. Dechaine, K. Shankar, (2015) 1401077, https://doi.org/10.1002/aenm.201401077.
Photocatalytic conversion of diluted CO2 into light hydrocarbons using [135] L. Yu, G. Li, X. Zhang, X. Ba, G. Shi, Y. Li, P.K. Wong, J.C. Yu, Y. Yu, Enhanced
periodically modulated multiwalled nanotube arrays, Angew. Chemie - Int. Ed. 51 Activity and Stability of Carbon-Decorated Cuprous Oxide Mesoporous Nanorods
(51) (2012) 12732–12735, https://doi.org/10.1002/anie.201205619. for CO2 Reduction in Artificial Photosynthesis, ACS Catal. 6 (10) (2016)
[119] Q. Zhai, S. Xie, W. Fan, Q. Zhang, Y.u. Wang, W. Deng, Y.e. Wang, Photocatalytic 6444–6454, https://doi.org/10.1021/acscatal.6b0145510.1021/
conversion of carbon dioxide with water into methane: Platinum and Copper(I) acscatal.6b01455.s001.
oxide co-catalysts with a core-shell structure, Angew. Chemie - Int. Ed. 52 (22) [136] C. Kim, K.M. Cho, A. Al-Saggaf, I. Gereige, H.-T. Jung, Z-scheme Photocatalytic
(2013) 5776–5779, https://doi.org/10.1002/anie.201301473. CO2 Conversion on Three-Dimensional BiVO4/Carbon-Coated Cu2O Nanowire
[120] M.L. Ovcharov, V.V. Shvalagin, V.M. Granchak, Photocatalytic Reduction of CO2 Arrays under Visible Light, ACS Catal. 8 (5) (2018) 4170–4177, https://doi.org/
on Mesoporous TiO2 Modified with Ag/Cu Bimetallic Nanostructures, Theor. Exp. 10.1021/acscatal.8b0000310.1021/acscatal.8b00003.s001.
Chem. 50 (3) (2014) 175–180, https://doi.org/10.1007/s11237-014-9362-x. [137] J. Jin, J. Yu, D. Guo, C. Cui, W. Ho, A Hierarchical Z-Scheme CdS–WO3
[121] M. Leclerc, A. Etxebarria, Y. Ye, E.J. Crumlin, G.M. Brisard, An APXPS Probe of Photocatalyst with Enhanced CO2 Reduction Activity, Small 11 (39) (2015)
Cu/Pd Bimetallic Catalyst Surface Chemistry of CO2 Toward CO in the Presence of 5262–5271.
H2O and H2, J. Phys. Chem. C. 124 (31) (2020) 17085–17094, https://doi.org/ [138] X. Zhao, Y. Fan, W. Zhang, X. Zhang, D. Han, L.i. Niu, A. Ivaska, Nanoengineering
10.1021/acs.jpcc.0c0471710.1021/acs.jpcc.0c04717.s001. Construction of Cu2O Nanowire Arrays Encapsulated with g-C3N4 as 3D Spatial
[122] T. Yui, Y. Tamaki, K. Sekizawa, O. Ishitani, Photocatalytic reduction of CO2: from Reticulation All-Solid-State Direct Z-Scheme Photocatalysts for Photocatalytic
molecules to semiconductors, Photocatalysis. 303 (2011) 151–184. Reduction of Carbon Dioxide, ACS Catal. 10 (11) (2020) 6367–6376, https://doi.
[123] Z. Gu, H. Shen, Z. Chen, Y. Yang, C. Yang, Y. Ji, Y. Wang, C. Zhu, J. Liu, J. Li, T.- org/10.1021/acscatal.0c0103310.1021/acscatal.0c01033.s001.
K. Sham, X. Xu, G. Zheng, Efficient Electrocatalytic CO2 Reduction to C2+ [139] Y. Wang, X. Shang, J. Shen, Z. Zhang, D. Wang, J. Lin, J.C.S. Wu, X. Fu, X. Wang,
Alcohols at Defect-Site-Rich Cu Surface, Joule. 5 (2) (2021) 429–440. C. Li, Direct and indirect Z-scheme heterostructure-coupled photosystem enabling
[124] H. Yang, X. Wang, Q. Hu, X. Chai, X. Ren, Q. Zhang, J. Liu, C. He, Recent Progress cooperation of CO2 reduction and H2O oxidation, Nat. Commun. 11 (2020) 3043.
in Self-Supported Catalysts for CO2 Electrochemical Reduction, Small Methods. 4 [140] W. Shi, X. Guo, C. Cui, K. Jiang, Z. Li, L. Qu, J.C. Wang, Controllable synthesis of
(2020) 1900826. Cu2O decorated WO3 nanosheets with dominant (0 0 1) facets for photocatalytic
[125] J. Wu, Y. Huang, W. Ye, Y. Li, CO2 reduction: from the electrochemical to CO2 reduction under visible-light irradiation, Appl. Catal. B Environ. 243 (2019)
photochemical approach, Adv. Sci. 4 (2017) 1700194. 236–242, https://doi.org/10.1016/j.apcatb.2018.09.076.
[126] E. Vahidzadeh, S. Zeng, A.P. Manuel, S. Riddell, P. Kumar, K.M. Alam, K. Shankar, [141] S. Kreft, R. Schoch, J. Schneidewind, J. Rabeah, E.V. Kondratenko, V.
Asymmetric Multipole Plasmon-Mediated Catalysis Shifts the Product Selectivity A. Kondratenko, H. Junge, M. Bauer, S. Wohlrab, M. Beller, Improving selectivity
of CO2 Photoreduction toward C2+ Products, ACS Appl. Mater. Interfaces. 13 (6) and activity of CO2 reduction photocatalysts with oxygen, Chemistry 5 (7) (2019)
(2021) 7248–7258, https://doi.org/10.1021/acsami.0c2106710.1021/ 1818–1833, https://doi.org/10.1016/j.chempr.2019.04.006.
acsami.0c21067.s001. [142] H. Li, Y. Lei, Y. Huang, Y. Fang, Y. Xu, L.i. Zhu, X. Li, Photocatalytic reduction of
[127] L. Wan, Q. Zhou, X. Wang, T.E. Wood, L.u. Wang, P.N. Duchesne, J. Guo, X. Yan, carbon dioxide to methanol by Cu2O/SiC nanocrystallite under visible light
M. Xia, Y.F. Li, A.A. Jelle, U. Ulmer, J. Jia, T. Li, W. Sun, G.A. Ozin, Cu2O irradiation, J. Nat. Gas Chem. 20 (2) (2011) 145–150.
nanocubes with mixed oxidation-state facets for (photo)catalytic hydrogenation [143] G.e. Yin, M. Nishikawa, Y. Nosaka, N. Srinivasan, D. Atarashi, E. Sakai,
of carbon dioxide, Nat. Catal. 2 (10) (2019) 889–898, https://doi.org/10.1038/ M. Miyauchi, Photocatalytic carbon dioxide reduction by copper oxide
s41929-019-0338-z. nanocluster-grafted niobate nanosheets, ACS Nano 9 (2) (2015) 2111–2119,
[128] S. Sorcar, Y. Hwang, C.A. Grimes, S.-I. In, Highly enhanced and stable activity of https://doi.org/10.1021/nn507429e.
defect-induced titania nanoparticles for solar light-driven CO2 reduction into CH4,
Mater. Today. 20 (9) (2017) 507–515.

23

You might also like