Review Articles
Review Articles
2 Review
6 Sunday Ogunlajae
a
7 Laboratoire de Chimie de Coordination du CNRS, UPR8241, Universite´ de Toulouse,
18 Africa
g
19 Department of Chemical Sciences, Tai Solarin University of Education, Ijebu Ode,
22 CA 92697-2175, USA
23 **
Corresponding author’s e-mail address: [email protected]
24 **
Corresponding author’s address: LCC-CNRS, Université de Toulouse, CNRS, UPS,
25 205 Route de Narbonne, BP44099, 31077 CEDEX 4 Toulouse, France
1
26 Corresponding author’s e-mail address: [email protected]
*
27 *
Corresponding author’s address: Department of Chemistry and Biochemistry, Florida
28 International University, 11200 SW 8th St, Miami, FL, 33199, USA
29
30 **
Corresponding author’s e-mail address: [email protected]
31 **
Corresponding author’s address: LCC-CNRS, Université de Toulouse, CNRS, UPS,
32 205 Route de Narbonne, BP44099, 31077 CEDEX 4 Toulouse, France
33 Abstract
34 Carbon dioxide (CO2) is the most important greenhouse gas (GHG), accounting for
35 76% of all GHG emissions. The atmospheric CO2 concentration has increased from 280
36 ppm in the pre-industrial era to about 417.86 ppm, and is projected to reach 570 ppm by
37 the end of the 21st century. In addition to reducing CO2 emissions from anthropogenic
38 activities, strategies to adequately address climate change must include CO2 capture. To
40 such as fuels and chemical feedstock. Due to their tunable chemistry (which allows
41 them to be selective) and high surface area (which allows them to be efficient),
42 engineered nanomaterials are promising for CO2 capturing and/or transformation. This
43 work critically reviewed the application of nanomaterials for the transformation of CO2
44 into various fuels, like formic acid, carbon monoxide, methanol, and ethanol. We
46 nanocomposites, as well as other routes suitable for CO2 conversion such as the
49 technologies were also discussed. Finally, we presented a section on the future outlook
50 of the field, which includes recommendations for how to continue to advance the use of
2
53 Table of contents
54 1. Introduction
72 4.3 Methanol
73 4.4 Ethanol
74 4.5 Methane
75 4.6 Formaldehyde
76 4.7 Hydrogen
3
78 5. Conclusions and future outlook
79
80 1. Introduction
82 climate change include global warming, increased frequency and intensity of natural
83 disasters and extreme weather events such as wildfires which was responsible for the
84 loss of ~$650 billion between 2016 to 2018, extreme droughts in Europe, heatwaves in
85 USA, and hurricanes in other parts of the world (Mustafa et al., 2020). Climate change
87 (Cassia et al., 2018; Solomon et al., 2010, 2009), which has led to the global surface
88 temperature of 0.88°C (1.58°F) in September 2022 which is above the average for the
89 20th century of 15.0°C (59.0°F), tying September 2021 as the sixth warmest for
90 September in the 143-year record. In addition, global CO2 levels was 280 ppm in the
91 pre-industrial era; it has now increased to 417.86 ppm as at October 19, 2022 according
92 to the Global Monitoring Laboratory (daily global CO2 trend) and it is projected to
93 reach 570 ppm by the end of the 21st century (Mustafa et al., 2020).
95 imperative that emission of CO2 is controlled, and the excess atmospheric CO2 is
96 captured. Several strategies exist for achieving atmospheric CO2 reduction. For
98 (Halder et al., 2014; Jansen et al., 2015; Li et al., 2013; Mondal et al., 2012). More so,
100 reduce the emissions of CO2 by putting taxes on carbon sources and fossil fuels (Atsbha
101 et al., 2021; Herzer, 2022; Omoregbe et al., 2020; Peiseler & Cabrera Serrenho, 2022).
102 Chemical CO2 conversion is also effective but is a herculean task due to the
4
103 characteristic thermodynamic stability of CO2 molecule. The stability of CO2 arises
104 from its structure (O=C=O) and possession of double covalent bonds with two oxygen
105 atoms (Ateka et al., 2022). The carbon in CO2 is extremely inactive and is kinetically
106 inert owing to its occurrence at the highest oxidation state in CO2 (Wang et al., 2022).
107 As a result, the transformation of CO2 to useful fuels and/or chemicals demands the
108 input of substantial amount of energy to cleave the C-O bond (Pawelec et al., 2021; Sha
110 The energy requirement for CO2 conversion is met through three major routes: (i) the
111 use of (nanomaterial-based) catalysts with active sites, (ii) the application of high
112 pressure and/or temperature, and (iii) application of energy from carbon-neutral and
113 renewable energy sources (Kamkeng et al., 2021). There are a number of technological
114 approaches that can be employed to execute transformation of CO2. In terms of using
115 catalysts for CO2 transformation, the processes and/or mechanisms involved may be
116 photocatalytic, electrocatalytic, thermal or non-thermal etc., to lower the energy barrier
117 and create alternative pathway for the catalytic CO2 reduction (Ateka et al., 2022). For
120 which interacts with CO2 and H2O molecules to produce hydrocarbons and other value-
121 added products. This is feasible only by appropriate catalyst selection, doping and
122 conditions (Low et al., 2017). Scheme 1 presents the various value added-products of
5
124
127 of CO2 have been published (Kamkeng et al., 2021; Low et al., 2017; Mustafa et al.,
128 2020; Pawelec et al., 2021; Sha et al., 2020; Wang et al., 2011; Wu et al., 2017).
129 Although a large proportion of them are detailed with respect to application and
130 transformation of CO2, recent advances in technologies for CO2 conversion have not
131 been sufficiently covered by the existing reviews. In addition, the technical challenges
132 facing the real-world implementation of these technologies are yet to the
133 comprehensively explored and presented. Therefore, this synthesis on CO2 conversion
6
134 to valued-added products attempts to provide current and comprehensive understanding
135 on nano-enabled CO2 conversion, routes of CO2 conversion, fuels generated from CO2
137 of CO2 into hydrocarbon fuels. The review also elaborates the research gaps that need
138 to be filled.
140 Nanomaterials have gotten a lot of attention in the research world because of their
141 fascinating small size (1-100 nm), which is responsible for their improved functionality
142 (Afolabi et al., 2022; Alli et al., 2022, 2021). In this regard, the use of nanomaterials
143 for direct transformation of CO2, to renewable organic fuels and other feedstocks has
144 become a promising solution to simultaneously mitigate the global energy crisis and
145 climatic change. Due to the thermodynamic stability, fully oxidized nature, and
146 inertness of CO2, various nanoscale catalysts have been developed to activate CO2
147 molecules into low (C1) and high hydrocarbon (C1-C9) molecules. Figure 1 depicts
148 possible catalytic pathways for CO2 transformation to fuels and chemical products.
151 epoxide to form cyclic and polycarbonates, and the emerging photocatalytic reduction
152 of CO2/H2O to CH4, CHOOH, CH3OH, etc. The nature and kinds of nanomaterials for
153 conversion of CO2 also play a vital role in product selectivity to get carbon-containing
154 fuels, in addition to process parameters such as solvent effects, reaction temperature,
155 H2/CO2 ratios, and partial pressure condition of CO2. The nanomaterials for catalytic
156 conversion of CO2 are range from plasmonic metal nanoparticles and metal oxides-
7
158 such as grapheme oxide (GO) reduced grapheme oxide (rGO) and carbon nanotubes
160
164 The natural process of photosynthesis, in which plants change CO2 and H2O into
165 oxygen and carbohydrates when exposed to sunlight, served as the inspiration for
166 photocatalytic CO2 reduction. This process transforms solar energy into chemical
167 bonds, which are then stored (carbohydrates). It combines light and dark reactions with
168 water oxidation and carbon dioxide reduction (or CO2 fixation) reactions. The process
170 semiconductor or metal complex as catalyst can be divided into four steps: (i)
171 absorption of light greater or equal to the band gap (ii) Separation of valence band (CB)
172 and conduction band (CB) (iii) migration of electrons from the VB to CB and migration
8
173 of electrons to the surface of cocatalyst (iv) adsorption and activation of CO2 by the
175 To understand the kinetic and thermodynamic point of view, the formal reduction
176 potentials for the reactions connected to the photoreduction of CO2 (at pH 7) is
177 presented in Figure 12a. Due to their extremely high negative reduction potentials
179 one-electron reduction processes of CO2, such as the generation of CO2 •, are not
180 practical (Lingampalli et al., 2017). A highly negative reduction potential results from
181 the conversion of C's sp2 to sp3 hybridization. Reduction potentials for HCOOH,
182 HCHO, CH3OH, and CH4 production are minimal (-0.81, -0.48, -0.38, and -0.246 V)
183 (Figure 12b) and positive relative to CB edges of various semiconductors. Proton
184 coupled electron transfer (PCET), in which electron transfer to CO2 is accompanied
185 with proton transfer, is thus preferable (Kovačič et al., 2020; Lingampalli et al., 2017).
186 A good number of works has been done recently to functionalize semiconductors and
187 it’s like as novel and effective photocatalysts. Various single and composites
188 nanomaterial have been reported for photocatalytic transformation of CO2. This section
190
191 Figure 2: (a) Energy diagram for CO2 reduction and water oxidation on a
9
193 process, demonstrating elements that may influence photocatalytic performance.
194 Reproduced with the permission from (Shen et al., 2020). Copy right 2020 Willey
195 Library
197 Plasmonic metal nanoparticles (NPs) such as Au, Ag, and Cu have a large optical
198 extinction cross section and the ability to absorb significant visible light which accounts
199 for up to 40% of solar irradiation to catalyze a wide range of photochemical reactions.
200 As a result, , they offer several advantages over other conventional catalysts like Pt, Pd,
201 Ru, and Rh, whose localized surface plasmon resonances (LSPRs) fall in the UV region
202 of the solar spectrum. A single silver nanoparticles (AgNPs) photocatalyst was excited
203 in a microfluidic reaction cell filled with CO2 led to CO2 reduction reaction CO2RR, to
204 CO and HCOOH. The synergy between photoexcited CO2 and catalytic active sites,
205 which is not clearly understood at the molecular level, was evidenced by in-situ
206 surface-enhanced Raman spectroscopy (SERS) (Kumari et al., 2018). The SERS was
209 HCOOH, HCHO and CH3OH), (see Figure 3). The real-time monitoring of CO2RR
210 activity by SERS shows that plasmonic AgNPs are selective for CO and HCOOH, with
214 CO (Yu et al., 2018). The CO2RR was conducted in isopropyl alcohol (IPA) saturated
215 CO2-aqueous solution to prevent h+/e- recombination during photoredox reaction, and
216 the CO2RR selectively produced CH4 (6.8 NP-1) and C2H6 (5.6 NP-1) after 10 h of light
217 irradiation. The results show that catalytic and selectivity of plasmonic gold
10
218 nanoparticles (AuNPs) for CO2 transformation to C1 and C2 is modulated by laser
219 excitation wavelength (photon energy), and photon flux (light intensity). Thus, the CH4
220 production rate increased with decreasing excitation wavelength and with overall 8e-
221 process (8H+ + CO2 → CH4 + 2H2O), while C2H6 production was only observed at
222 lower excitation wavelength with 14e- process (14H+ + 2CO2 → C2H6 + 4H2O). Zeng
223 et al. (Zeng et al., 2021) embedded uniformly distributed plasmonic CuNPs into a
224 triphenothiazine benzene ring (TPB) for photocatalytic CO2 reduction to CO, H2 and
225 CH4. CO evolution rates over TPB-Cu increased as the wt% of Cu bonded to TBP units
226 increased, and CuNPs with 1wt% content yielded a CO evolution rate 3.2 times higher
227 than bare CuNPs. Interestingly, the evolution rates of H2 and CH4 on TPB-Cu remained
228 relatively constant, implying that incorporating plasmonic CuNPs into TPB increased
229 the directional selectivity for CO formation over H2 and CH4. Plasmonic cupper NPs
230 conjugated with TBP could provide a large contact area for hot electron migration,
231 opening new doors for energy and carbon cycle processes.
11
232
233 Figure 3: Representative SERS spectra that capture specific adsorbates or products
234 observed in photocatalytic CO2 reduction on Ag NPs. Specific species captured include
235 (a) physisorbed CO2, (b) HOCO* reaction intermediate, (c) carbon monoxide or CO, (d)
237 hydroxide or OH*, and H2, and (g) Histogram of prevalence by the number of catalytic
238 NPs on which each type of species was observed. Reproduced with permission from the
242 photocatalysts and to expand their light harvesting to UV region, researchers have
12
243 turned to multi-component semiconductors for a variety of applications. Decoration of
244 metal oxide semiconductors with transition or plasmonic metals has been a huge
245 success. Both plasmonic and transition metals on semiconductors can help harness UV
246 and visible light by reducing energy band gaps, which play key roles in CO2 activation
248 For many years, the selective absolute conversion of CO2 to CO has been a source of
249 contention. In 2020, Nguyen et al. (Nguyen et al., 2020) constructed air-stable Z-
252 byproducts) and with an evolution rate of 1166.72 μmol g-1 h-1 after 7 h of light
253 illumination. On one hand, the plasmonic Ag (1wt% Ag) increased the photocatalytic
254 electron reaction rate up to 2333.44 μmol g-1 h-1, and facilitated intense absorption light
255 in the visible region of the catalysts due to its strong LSPR effect. The presence of Ti 3+
256 states in the catalysts, on the other hand, facilitated the physisorption, stabilization, and
257 reorientation of the intermediates formed during CO2 reduction. Other gas products
258 such as CH4 and H2 were produced in the absence of Ag deposition, suggesting the
260 Zhang et al. (Zhang et al., 2021) reported p-type Cu2O photocatalysts with phase
261 structures of 100 (100Cu2O) and 111 (100Cu2O) modified with Pd nanoparticles for
262 CO2RR activities. The well-matched band alignments between Pd and 100Cu2O
263 increased the work functions (Ws) of the 100Cu2O-Pd composite, resulting in an
265 3.0 times higher than bare 100Cu2O nanoparticles. Pd nanoparticles (in 100Cu2O-Pd)
13
268 However, 111Cu2O-Pd composites lacked electron-hole mobility due to mismatched
269 band structures between 111Cu2O and Pd species, resulting in lower work function and
271 importance of crystal surface structure and matched band structure construction for
273 Wang and co-workers demonstrated LSPR photo-assisted electron migration from
274 plasmonic CuNPs to ZnO via metal-support interfaces in Cu/ZnO catalysts for CO2
275 reduction to MeOH (Wang et al., 2019). In the photothermal process, the rate of MeOH
276 formation was 2.13 μmol/g. min, compared to 1.38 μmol/g. min recorded in pure
277 thermal process, and the apparent activation energy decreased from 82.4 to 49.4 kJ mol-
1
278 . The rate-determining step in the thermal process was the formation of reaction
281 followed the same reaction pathway as thermal process, with the exception that induced
283 (MeOH). Under visible light irradiation, the product selectivity over Cu/ZnO remained
284 relatively constant and the decreased activation energy had no effect. Small amounts of
285 Pt nanoparticles have also been used as co-catalyst on ZnMn2O4 nanorods as visible-
286 light-driven catalysts for promoting CO2 reduction to produce methane, as was the case
287 in other studies (Alhaddad and Shawky, 2021) (see Figure 4a). The wt% of Pt species
288 on ZMO nanorods influenced the catalytic activity and selectivity of Pt/ZMO catalysts,
289 with 1.5% Pt/ZMO exhibiting the best photocatalytic performance for CO2RR activity
290 (Figure 4b). Band connection and surficial interactions between Pt species and ZMO
291 nanorods with optimal Pt/ZMO at 1.5% Pt enhanced selective reduction of CO2 to
14
293
295 different illumination time, and (b) impact of Pt content in Pt/ZMO composite for a
296 dose of 1.0 g/L for each within 9 h of light illumination. Reproduced from (Alhaddad
298
299 V, Cr, and Co loaded on TiO2 were studied for CO2RR to produce H2, C1 and C2+.
300 Higher photoconversion rates were observed compared to pristine TiO2, which were
301 found to be sensitive to the metal concentrations and TiO2 crystallite size modification
302 (Ola and Maroto-Valer, 2015). After 4 h of illumination, 1wt% Co-TiO2 demonstrated
303 the highest photoconversion rates of 62.91 μmol/gcath for H2 and 26.12 μmol/gcath for
304 MeOH (Figure 5). Ni-modified CeO2, MgO, α-Al2O3 and TiO2 NPs were studied for
305 CO2 methanation, and the effects of the solid supports were elucidated (Tada et al.,
306 2012). CO2 methanation pathways followed two steps: (i) reaction of CO2 and H2 at the
307 meta/metal oxide interfaces to afford CO through reverse water gas shift reaction (CO2
308 + H2 → CO + H2O), and (ii) prompt methanation of CO with H2 (CO + 3H2 → CH4 +
309 H2O) at low temperature via the Fisher–Tropsch synthesis (FTs). CeO2 has the highest
310 number of CO2 adsorption sites and exhibits the highest CO2 methanation with near-
311 unity CH4 selectivity. Li and co-workers showed that partially oxidized metallic Fe NPs
15
312 (Fe and FeOx) deposited on MgO-Al2O3 and then reduced under H2 at different
313 temperatures photothermally hydrogenated CO2 into C1 and C2+ hydrocarbons (Li et al.,
314 2021). The Fe-500 sample (reduction performed at 500°C) showed the highest CO2
315 conversion and excellent selectivity for C2+ hydrocarbons with 0% CO. The FeOx phase
316 promotes CO2 reduction to CO via RGSW reaction and modulates the electronic
317 properties of Fe phase to suppress CH2 and CH3 hydrogenation, thus enhancing C2+
319
320 Figure 5: Irradiation time profile of 1 wt%Co–TiO2 for CO2 reduction to fuels.
322
323 The conjoining of metal, e.g. Ag and narrow band semiconductor e.g. InS2 has been
326 excellent photoreduction activity of carbon dioxide in H2O with the formation of
327 mostly methanol, formaldehyde and formic acid. AgCuInS2 combined with TiO2
328 semiconductors (AgCuInS2−TiO2) absorb photon energy, which led to the excitation of
329 electrons from valence band (VB) to conduction band (CB), creating electron−hole
330 pair, and the graphene act as electron assimilates (Figure 6).
16
331
332
333 Figure 6: Mechanism of photocatalytic CO2 reduction. Re-used with permission from
336 Many metal-functionalized mesoporous aluminas and silicas have been studied as
338 Tropsch (FT) reactions. In terms of activity and selectivity, mesoporous materials with
339 high Brunauer-Emett-Teller (BET) surface areas are important features for CO2
340 adsorption and conversion (Petrovic et al., 2021). The nature of metal species
341 immobilized on porous oxides also influences the CO2RR selectivity for either CO or
342 CH4. Deposition of Ni on CaO-Al2O3 showed CO2 conversion with 81% selectivity to
343 CH4; however, lower performance of the Ni-based catalyst in the subsequent
344 methanation reaction (due to partial oxidation of Ni fractions) suggests that a better
345 approach for CO2RR on metal/mesoporous materials is required (Mutz et al., 2015).
346 Addition of La as a second metal to Ni/-Al2O3, enhanced direct CO2 methanation (CO2
347 + H2 → CO + H2O) via the Sabatier reaction, increasing the CO2 conversion rate to
348 CH4 from 42% to 80.3% (Rahmani et al., 2019). CO2 methanation over mono and
17
349 bimetallic catalysts was temperature, feedstock (H2/CO2 ratio), and gas hourly space
350 velocity (GHSV) sensitive. CO2 was transformed into CO as a result of the RWGS
351 reaction which occurred at higher GHSV values (between 6000-18,000 mL/(gcath)),
352 350oC, and 3.5 H2/CO2 ratio, resulting in decreased CH4 selectivity. In contrast, Bashiri
353 and coworkers demonstrated that lowering the GHSV from 2000 to 1250 nml/min over
355 methanation significantly (Bashiri et al., 2018). The results also show that adding
356 HZSM5 to Al2O3 increases the rate of CO2 conversion to light hydrocarbons, CO, and
357 CH4 by 15.53%, indicating that combined supports have the potential to improve CO2
359
361 surface (b) Activation energy calculation for CO2 conversion reaction over Cu/fumed-
362 SiO2 and CuNi/fumed-SiO2. Reproduced from (Kumar et al., 2022) with permission.
363
364 Mesoporous siliceous materials are valuable supports because of their large number of
365 active sites for supporting various metal ions to catalyze CO2RR to renewable
366 feedstocks. Single Cu-based catalytic systems are not suitable for direct CO2
18
367 hydrogenation to either CH4 or CH3OH because they lead preferentially to the
369 enhanced activity and selectivity for CO, as well as a tendency for methanation to CH4
372 Figure 7a (Kumar et al., 2022). The incorporation of Ni increases the system's catalytic
373 active sites and lowers the activation barriers for CO2RR to CO and CH4, Figure 7b.
374 Mesoporous Cu/SiO2 doped with Ga and/or Zn had greater catalytic activity for direct
375 CO2 reduction to MeOH than undoped Cu/SiO2 (Paris et al., 2020).
376 The strong promoting effects of dopants restricted the occurrence of RWGS reaction
377 for CO formation while increasing selectivity for MeOH one. Both dopants provide
378 more active sites on the catalysts, lowering the activation barrier of MeOH synthesis
380 When mesoporous TiO2/SiO2 was irradiated with visible light without Cu species, CO
381 was the main product of photo-transformation of CO2/H2O (Li et al., 2010). Cu
382 deposition on TiO2/SiO2 improved photoreduction of CO2 to CO, and CH4 selectivity
383 increased as well, peaking at 0.5wt% optimum Cu specie on TiO2/SiO2 (Figure 8). XPS
384 allowed to identify the deposited Cu species as Cu2O, which acts as a photoexcited
385 electron sink, limiting the recombination of photoinduced holes and excited electrons
386 and thus enhancing multi-electron photoredox reactions for CO2 to CO and CH4. Fu et
387 al. designed photocatalytic systems based on CoO and Co3O4 nanoparticles to convert
388 CO2 to CO using bare CoONPs, CoO/s-SiO2, CoO/s-SBA-15, and Co3O4/s-SBA-15 (Fu
389 et al., 2019). Under similar photoredox conditions, CoO/s-SBA-15 exhibited the best
390 activity for CO2RR to CO with 85% selectivity at 14,147 μmol/hg after 2.5 h.
391 Interfacial interactions between porous silica -OH groups and reaction substrates
19
392 enhance sorption and photocatalytic CO2 reduction to CO over CoO/s-SBA-15 catalyst
393 system.
394
395 Figure 8: (a) Peak production rates of CO and CH4 on various TiO2-containing
396 catalysts, and (b) Peak production rates of CO and CH4 as a function of Cu loading in
397 Cu/TiO2–SiO2 catalysts. Reproduced from (Li et al., 2010) with permission.
400 MOFs are recognized as catalytic systems of choice for CO2RR conversion to
401 renewable energy feedstocks due to the large panel of surface functionalities and
402 geometries, porosity as well as the possible customization these materials offer
403 specifically for CO2 capture and utilizations (CCUs). Architectural flexibility and ease
404 of structural modification of MOFs through insertion of different metal ions (or metal
405 clusters) together with interconnection of multidentate organic linkers with different
407 photocatalysts for CO2 reduction (Ejeromedoghene et al., 2022). Wang and coworkers
408 investigated the photocatalytic conversion of CO2 to formate over a variety of amino-
409 functionalized and unfunctionalized Fe-based MOFs. They found that amine-
20
410 functionalized Fe-based MOFs outperformed their respective parent MOFs in this
413 exhibits the best catalytic activity, to generate HCOO- at the evolution rate of 178
414 µmol/g (Wang et al., 2014). Mechanistic pathways for enhanced activities include
415 excitation of NH2 functionality, electron transfer to the Fe center and direct excitation
416 of Fe-O cluster (Scheme 2). Wei et al. reported visible-light induced CO evolution
417 from CO2RR over a series of UiO-based MOFs photocatalysts modified with different
418 functional groups (F, NH2, OCH3, CH3) (Wei et al., 2019). They observed an influence
420 functional groups, CO evolution following the order UiO-68-F < UiO-68-NH2 < UiO-
421 68-CH3 < UiO-68-OCH3. The superior performance of UiO-68-OCH3 was attributed to
422 the pronounced charge separation efficiency of OCH3. Crake et al. investigated
424 nanocomposites in photoreduction of CO2 (Crake et al., 2019). They found that
425 TiO2/CNNS possessed improved CO2 conversion and adsorption when compared to the
426 individual materials used to make the composites (> 10 times). In addition, it was
427 reported that the composite material with more {001} TiO2 facets was the most
428 catalytically active and this is attributed to enhanced charge transfer at the TiO2/CNNS
429 heterojunction interface. Previously, Crake et al (Crake et al. 2017) coupled NH2-UiO-
430 66 onto pre-formed TiO2 nanosheets via in situ process to develop TiO2/NH2-UiO-66
431 heterojunction with permanent pore sizes and exposed catalytic actives sites for
432 enhance CO2 uptake and photoreduction of adsorbed CO2 to CO under UV-visible
433 light. The evolution rate of CO2-to-CO conversion over the bound nanocomposite was
434 1.9 times more than that of pure TiO2. Interestingly, the photoreduction efficiency of
21
435 mechanically blended TiO2 and NH2-UiO-66 is comparable to pure TiO2, proving that
436 in situ growth method was necessary for the enhanced catalytic activity. Transient
438 66 was due to the efficient charge transfer at the interface of heterojunction.
439
441 photocatalytic reduction of CO2 to formate. Reproduced from (Wang et al., 2014) with
442 permission.
443 Besides the use of MOFs with multidentate linkers and different functional groups,
444 multi-metallic MOFs can also offer improved stability and overall catalytic efficiency
445 for CO2 reduction in comparison to monometallic counterparts. Different MOFs with
446 multi-metallic nodes have been explored for CO2RR. For instance, bimetallic ZnCu
448 synthesis (Stolar et al., 2021). Selectivity for MeOH formation increased by 4.7 times
451 findings highlight the significance of using heterometal nodes and amorphous MOF to
452 get improved activity and selectivity. Fe-N-TiO2/CPO-Cu MOF nanocomposites were
22
454 leading to effective photocatalytic reduction of CO2 (khalilzadeh and Shariati, 2019).
456 CPO-Cu-27 (CPO) in composites from 8 to 65wt.%, and the Fe-N-Ti@50CPO sample
457 with 50wt% CPO yields the highest conversion of CO2 to methane. The performance
458 achieved is ca. 1.7 times higher than in the case of CO2 methanation over Fe-NTiO2.
460 metal NPs-based MOFs (Liu et al., 2021). Their results show that Co-MOF-3 with two
461 stacked piezoelectric phononic crystal nanoplates had a 2.5-fold higher CO2
462 photoreduction efficiency than Co-MOF-1 with single nanoplate ( Figure 9). The Co-
463 MOF-3 homojunction system demonstrated a high CO2 to CO conversion rate with
464 100% selectivity, outperforming most of the previously reported semiconductors for
466 Polyoxometalates-based MOFs (POMOF) composites have also been studied for
467 CO2RR due to POM's unique ability to act as nodes or pillars, and more importantly, as
468 high-efficiency catalytic sites in the MOF cages to adsorb reactants and improve MOF
469 acidity, stability, and redox activity. Li et al. (Li et al., 2019) reported the use of a
472 system. The presence of a hydrophobic ligand improves the chemical stability of NNU-
473 29 and reduces H2 generation due to water attack to some extent, resulting in 98%
474 selectivity for CO2 reduction to HCOOH in aqueous solution after 16h of light
475 irradiation. Wang et al. (Wang et al., 2021) recently proposed SO3H-MIL-101-Cr
476 (SO3H-MOF) supported silicomolybdic acid (SMA) as a host-guest composite for CO2
477 photoreduction. Pristine POM and MOF have negligible CO2-to-CO activity compared
23
479 remarkable yield of 829.46 mol/g and selectivity of 93.2%. The superior reductive
480 activity of POMOF was attributed to increased CO2 affinity for silicomolybdic acid
481 (SMA) and synergy between POF and MOF, resulting in enhanced charge carrier
482 separation.
483
484 Figure 9: (a) Schematic view for the hollow Au/Ag directed formation of Co-MOF-3
485 nanocomposites. (b) Evolution of CO and H2 as a function of reaction time during the
486 CO2 reduction under visible-light irradiation for Co-MOF-1 and Co-MOF-3
487 photocatalysts. (c) Stability of Co-MOF-3 during the continuous CO2 reduction
488 process. Reproduced with permission from ref. (Liu et al., 2021). Copyright 2021
489 Nature.
492 deactivation in catalytic systems, environmental concerns about toxicity (Fall et al.,
493 2022; Keller et al., 2017), and more often than not, their moderate catalytic
494 performance in terms of activity and selectivity, have limited their use in CCUs.
495 Graphene oxides, fullerenes, and graphitic carbon nitrides have recently emerged as
496 valuable metal-free platforms for CO2RR preparation, chemical inertness, and natural
497 abundance (Yang et al., 2022). These materials are frequently used as a standalone
499 their CO2RR performance (Shen et al., 2020). Several reports have shown that GO can
24
500 act by itself as a photocatalyst for CO2 reduction either by adapting its method of
502 2020). Hsu et al. (Hsu et al., 2013) studied the effect of volume of H3PO4 used as a
504 Their results show that modified GOs have increased visible light absorption and
505 improved CO2 photoreduction to MeOH compared to pristine GO. The CO2-to-MeOH
506 conversion rate of GO-3 with excess H3PO4 was found 6-fold higher than with pure
507 TiO2. Kuang et al. (Kuang et al., 2020) reported on the photoreduction of CO2 to CO
509 sunlight provided the highest CO2 reduction efficiency, with a CO yield of 1.3-fold
510 higher than the UV-activated GO and nearly 3 times higher than with pristine GO. The
511 observed efficiency is due to light-induced increased defects and restoration of large π-
512 conjugative systems. Compared to GO, reduced-GO (rGO) are superior solid supports
513 for other semiconducting materials due to their high thermal and electrical conductivity.
514 Yang et al. supported quinacridone (QA) NPs on rGO nanosheets through H-bonding
515 and with different rGO contents and observed that this metal-free photocatalysts
516 allowed photoreduction of CO2 to CO and CH4 under UV light (Yang et al., 2019). The
517 sample containing 2% rGO had the highest CO2 reduction efficiency and selectivity for
518 CO and CH4 due to the synergistic effects of QA as a photosensitizer and rGO as an
520 Based on the number of published articles on photocatalysis, graphitic carbon nitride
521 (gCN) materials is the highest promising among the reported carbon-based materials
522 for photoreduction of CO2, due to its facile tunable synthetic routes, structural
523 maneuverability, and tailorable morphology. However, they are limited due to their
524 narrow visible light absorption window, low specific surface area, poor electrical
25
525 conductivity, and rapid recombination of photoexcited excitons (Shi et al., 2018). A
526 variety of strategies, including structural modification via elemental doping, the
527 attachment of a semiconducting motif, functional group modulation of gCN, and others,
528 can significantly improve the photocatalytic CO2 reduction (Alaghmandfard and
529 Ghandi, 2022). Nonmetals like P, O, B, S, and N have become popular as interstitial or
530 substitutional dopants on gCN for CO2 photoreduction and other photocatalytic
531 reactions.
532 Arumugam and coworkers synthesized a series of porous gCN doped with various
534 (Arumugam et al., 2022) among which S-doped gCN showed the highest CO2-to-CH4
535 conversion rate compared to others. Imaging studies showed that bare g-C3N4 has a
536 thin-layered structure with irregular porosity, whereas S-doped g-C3N4 has a denser
537 structure with a homogenous distribution of S atoms on the g-C3N4 matrix, suggesting
538 that S atom can act as an electron sink for improved activity. Doping of S atoms into g-
539 C3N4 did not enhance the surface area or modify the energy band gap of g-C3N4; rather,
540 the remarkable boost in photoreduction efficiency was due to charge separation and
542 and hydroxylammonium chloride precursors structurally transformed bulk g-C3N4 from
544 and that subsequent heating of the intermediate under NH3 atmosphere led to the
545 formation of porous TCN(NH3) nanotubes with a 1D tubular structure. The structural
546 transition from bulk gCN to 1D tubular structure was confirmed using XRD, FTIR,
547 SEM, and TEM. The CO2-to-CO conversion rate of porous TCN(NH3) nanotubes was
548 17 and 15 times that of bulk g-C3N4 and TiO2, respectively. Tubular structure of
549 TCN(NH3) has permanent porosity and a large surface area, which enhances solar light
26
550 utilization and spatially resisted photoinduced charge recombination, resulting in a
551 better photoreduction efficiency than bulk g-C3N4. Huang et al. (Huang et al., 2015)
552 successfully photoreduced CO2 to CH4 and MeOH over bulk g-C3N4 and amine-
554 (MEA) solution. Both pure and functionalized g-C3N4 have a 2D lamellar structure
555 with equal porosity, indicating that the amine functionalization has no effect on g-C3N4
556 morphology. The functionalized g-C3N4 heated in an oil bath for 5 hours had the
557 highest CO2 to fuel conversion rate. The CO2 photoreduction efficiency over amine-
558 functionalized g-C3N4 was significantly higher than pristine g-C3N4 due to the
560
561 Figure 10: Schematic view and mechasism of CN/CTF heterostructure for
562 photocatalytic CO2 reduction. Reproduced with permission from ref. (He et al., 2022).
563
564 gCN has also been decorated with COFs and carbon dots (CDs) to fabricate metal-free
565 composites for improved visible-light-induced CO2 reduction. He et al. (He et al.,
566 2022) reported for the first-time gCN/covalent triazine framework (2D/2D CN/CTF)
567 heterostructures for CO2 photoreduction. The TEM micrograph of CN/CTF reveals an
27
569 morphology of CTF, validating successful formulation of 2D/2D CN/CTF
570 heterostructure. The optimal CN/CTF heterostructure is up to 25.5 and 2.5-fold more
571 efficient in photoreduction of CO2 to CO than CTF and CN, respectively. The enhanced
574 charges separation, cleaving O=C=O to selectively produce CO (Figure 10). Wang et
575 al. (Wang et al., 2022) reported van der Waals interfacial bonding between N-deficient
576 gCN and Tp-Tta COF, demonstrating exceptional stability and selectivity for CO2
577 photoreduction to CO. DFT calculations and EPR analysis confirmed the strong
578 electronic coupling between gCN and Tp-Tta COF, which was responsible for
579 improved photoreduction efficiency of the composite via large active sites for CO2
580 capture and, an accelerated S-scheme charge separation system. Wang et al. (Wang et
581 al., 2020) described carbon nanodots (CDs) electrostatically bounded to the surface of
583 to promote selective reduction of CO2 to nearly 100% methanol. Structural elucidation
584 reveals that the boundaries and edges of CN nanosheets, which possess a graphene-like
585 structure, were coherently decorated by graphitic structure of mCDs, indicating strong
586 electrostatic interactions between the two phases, which may have enhanced the
587 composite's charge transfer process. In contrast, the amorphous structure of sCD was
588 barely located in the phase structure of CN/sCD composite following heterojunction
m
589 formation, indicating weak interaction between the two phases. CD features near-
592 According to this study, mCD/CN photoreduces CO2 to 6e- product MeOH (Figure
593 11a), while sCD/CN favors formation of 2e- product CO (Figure 11b). The authors
28
m
594 attributed the improved CD/CN reduction efficiency to the hole-accepting role of
m
595 CD, which led to electron accumulation on CN surface and facilitated a multi-electron
596 reduction process to produce MeOH with near-unity selectivity (Figure 11c).
597
m
598 Figure 11: Photocatalytic activity of (a) CD/CN for methanol formation, and (b)
599 CD/CN for CO formation measured under visible light (λ > 420 nm). (c) Schematic
s
600 diagram of photocatalytic CO2 reduction by the mCD/CN and sCD/CN. Reproduced
601 with permission from ref (Wang et al., 2020). Copyright 2020 Nature.
602 Apart from all the aforementioned examples, this section also surveyed other
605
607 The electrochemical transformation of CO2 to value added fuels has received
608 considerable attention in recent years owing to its plethora benefits including energy
609 production, storage, and others. The conversion process is located at the interface
610 between the solid catalyst surface and the electrolyte, as is the case with all
611 electrochemical reactions. A compact layer of ions next to the electrode, followed by a
612 diffusion layer of solvated ions, can be used to describe the interface's structure (Rosen
613 and Hod, 2018). The more the potential an electroactive species has, the more the
29
614 distance it can travel to the interface, and its relative orientation to the interface are all
615 influenced by the bilayer structure which is thus essential for the rate and efficiency of
616 the electrochemical CO2RR. Electrochemical CO2RR can produce useful chemicals and
617 fuels including formaldehyde, CO, formic acid, ethylene, ethanol, methane, and
618 methanol, depending on the electrical voltage supplied to the electrode (Chen et al.,
619 2020).
620 The methods most frequently used for electroreduction of CO2 are cyclic voltammetry
621 (CV) and linear sweep voltammetry (LSV). The catalytic performance can also be
622 evaluated using chronoamperometry. These CO2 conversion operations are typically
623 carried out at a specific reaction rate and ambient conditions of temperature and
624 pressure (Nielsen et al., 2018). A coupled proton electron transfer that results in the
625 protonation of the oxygen atom O (eq 1) or the carbon atom C (eq 2) triggers the
626 adsorption of CO2 on the electrode surface (eq 4) (Nielsen et al., 2018; Zhu et al.,
627 2014). As such, this process requires the use of a large potential, which results in
628 competing reactions such as the hydrogen evolution reaction (HER). In lieu of the
629 above, integration of a catalyst in the system allows to reduce the overpotential and
632 electrocatalytic CO2 transformation. These catalysts will allow for a faster electron
633 transfer coupled with protons, resulting in the observed products HCOOH (eq 2 and eq
30
638 * + CO2 + H+ + e-→ *COOH (Eq4; a slow step)
641 As shown in Figure 12, various studies using different metal electrodes modified with
642 catalysts such as Sn, In, Au, Ag, Pd, Bi, Co, Mo, and others have been reported. As the
643 other side reactions (i.e. HER) can be almost suppressed, these materials exhibit
644 significant catalytic properties for the CO2 transformation (Lu et al., 2019).
645 Furthermore, they present an extremely low energy barrier for the formation of the
646 intermediate, which must then be protonated at the carbon or oxygen atom before being
648
31
649 Figure 12: Overview of some typical Au, Ag, Pd, Sn, and Ni-based nanocatalysts for
650 electrochemical reduction of CO2 to CO. All potentials are contrasted with an electrode
651 made of reversible hydrogen (RHE). NP stands for nanoparticle, whereas FE stands for
652 faradic efficiency. Reprinted with the permission from (Yin et al., 2019). Copyright
654 Yang et al also investigated the CO2RR on Ag, Pd, and Ag@Pd nanocubes. Pd
655 nanocubes led to CO with high selectivity as a result of Pd turning into Pd hydride
656 (PdH) during CO2RR (Yang et al., 2022). In contrast, Ag@Pd makes the formation of
657 PdH more difficult due to inhibited production of CO from the intermediate *HOCO
658 which thus adjusts the reaction pathway towards HCOOH. The authors concluded that
659 Ag nanocubes have high selectivity for H2 with no phase transition during CO2RR.
660 Li et al. (Z. Li et al., 2022) studied electrochemical CO2RR to formate, which not only
661 has the potential to reduce the issue of global warming brought on by CO2, but also has
662 the potential to be exploited for energy storage as intermittent renewable energy. A
663 bimetallic BiSn nanomaterials were electrodeposited on copper mesh and tested as
664 electrocatalysts for the conversion of CO2 to formate. The resulting electrodes
665 demonstrated excellent catalytic activity at 1.0 V vs. RHE, with a partial current density
666 of 34.0 mA/cm2 and an astonishingly high faradaic efficiency of roughly 94.8%. The
667 production of formate at a remarkable rate of 634.3 mol/cm2h -significantly higher than
668 most rates previously reported- is noteworthy. The quick electron transfer of BiSn
669 nanomaterials also contributed in the creation of the CO2•- intermediate. Additionally,
670 they produced a metastable metal oxide/metal interface that helped to stabilize the
671 CO2•-intermediate and suppress the HER side reaction. The pathway connected to the
672 OCHO• intermediate was improved because of the optimized electronic structure of
32
673 BiSn, promoting the conversion of CO2 to formate. According to these results, Bi/Sn
674 bimetallic catalysts are a viable option for very effective CO2 reduction.
675 In contrast, the conversion of CO2 to CO and formate is less difficult than CO2RR to C-
676 H hydrocarbons, which are more significant scientifically and technologically but are
677 more difficult to manage. Concerning the synthesis of ethylene, Cu and Cu2O are
678 regarded as the most common and effective electrocatalysts for CO2RR. Multi-carbon
679 oxygenates, such as ethanol (the primary product), acetate, and n-propanol, can also be
682 In order to achieve a lofty ultra-high selectivity of 82.4% for ethylene in an H-cell
683 system with good catalytic stability and material durability of 100 h, Sultan et al.
685 Al2CuO4 nanosheets uniformly coated with CuO nanoparticles (CuAl-1: CuO/Al2 CuO
686 4 -23) by phase. They found that Cu (CuAl-1) permitted high surface CO intermediate
687 coverages and improved CO adsorption for C-C coupling to *OCCO, a generation of
689 intermediates.
690 Cu catalysts for CO2RR can produce a wide range of hydrocarbons at a high catalytic
691 activity. For selective CO2RR, Yin et al. (Yin et al., 2019) used a catalyst based on
692 copper nitride (Cu3N) nanocubes (NC). With a FE of 60%, a mass activity of 34 A/g, a
693 C2H4/CH4 molar ratio over than 2000, and strong selectivity and stability of CO2RR
694 towards ethylene at -1.6 V vs. RHE, their 25 nm Cu3N NCs showed good performance.
695 This lacks of product selectivity is a significant barrier to the use of these catalysts for
33
696 selective CO2RR, and more research on the application of these substitute catalysts is
698 In a related piece of study, Liu and coworkers noted the formation of exceptional
699 oxygen-rich ultrathin CuO nanoplate arrays that spontaneously transform into Cu/Cu2O
700 heterogeneous surfaces after CO2RR (Liu et al., 2022). The electrocatalyst displayed a
701 high ethylene energy efficiency of 84.5%, steady electrolysis for 55h, and a total cell
702 ethylene energy efficiency of 27.6% at 200 mA.cm-2 in a flow cell with a neutral KCl
704 highlighted that stable nanostructures, stable Cu/Cu2O interfaces, and improved
705 adsorption of the intermediate *OCCOH can enhance selectivity and amount of C2H4
706 produced.
707 Wang et al. developed CuBi NPs that showed a CH4 FE as high as 70.6% at 1.2 Vs
708 RHE. According to their study, adding Bi NPs to the matrix (CuBi NPs) significantly
709 boosted the FE when compared to regular Cu NPs by almost 25 times (Wang et al.,
710 2020). DFT studies showed that alloying Cu with Bi lowers the energy required for the
713 In addition to the aforementioned systems, other metal-based nanomaterials have been
714 investigated towards the production of hydrocarbon via the electrochemical reduction
34
716 Table 1. Selected examples of electroactive metal nanoparticles for converting CO2 into hydrocarbon products
ial vs
Density
RHE (V)
(mA/c
m2 )
pass through
between g-C3N4
35
and CU2O-FeO
are responsible
CO2 reduction to
CO
of EtOH is
nanowire
attributed to the
arrays
synergistic effect
of morphology
36
of trace Fe atoms
to graphene
matrices
porphyrin serve as a
transfer from
ZnIn2S4 to the
FeTCPP.
activity of the Cu
37
nanocrystals
toward the
production of n-
propanol was
linked to the
number of defect
sites on their
surface.
electrochemical
38
performance
respectively tuned to
systematically
controlling
hydrocarbon
electrochemical
reduction of
CO2.
39
by Ni NPs active sites of
mesoporous
carbon is good
for selective
reduction of CO2
to CO.
about 70 h (in 3-
h electrolysis
experiments at
different
potentials).
717 FE - Faradaic Efficiency; EtOH – Ethanol; FETCPP - Tetra(4-carboxyphenyl)porphyrin iron(III) chloride; HCOO- - Formate ion; NPs –Nanoparticles; AuNPs –
40
719 In the quest for additional electro-active nanoparticles for CO2 reduction, non-metal
720 NPs have been explored in order to speed up the reaction rate and increase the
721 selectivity for products. For instance, boron phosphide NPs were shown by Mou et al.
722 (Mou et al., 2019) to function remarkably as a non-metal electrocatalyst for the highly
724 m KHCO3, this catalyst achieves a high FE of 92% for CH3OH at 0.5 V. According to
725 DFT calculations, B and P synergistically facilitate the binding and activation of CO2,
726 and the *CO + *OH to *CO + *H2O pathway, with a free energy change of 1.36 eV, is
727 the rate-determining step for the CO2RR. Liu et al. (Liu et al., 2022) reported B/N-
728 doped sp3/sp2 hybridized nanocarbon, which consists of extremely small carbon NPs
729 with an sp3 carbon core surrounded by a sp2 carbon shell, as an effective electrocatalyst
730 for CO2 to ethanol reduction at relatively low overpotentials. They claim that CO2
731 reduction occurs during the synthesis of ethanol and acetate at 0.5 0.6 V (vs. RHE),
732 with ethanol accounting for 51.6% - 56.0% of the total reduction. Due to the combined
733 effects of sp3/sp2 carbon and B/N doping, there is a strong ethanol selectivity.
734 To perform electrocatalytic CO2 reduction to liquid ethanol and acetone, Yuan and
735 colleagues presented their findings. Yuan and colleagues used five pyridine derivatives,
738 al., 2018). Surprisingly, the greatest catalyst for electrochemical CO2 conversion is
740 overall FE of 45.8%. Furthermore, the CO2 reduction potentials of the other four
741 pyridine derivatives functionalized GO varied with respect to ethanol and acetone. Li
742 and colleagues described non-metal nanoporous S-doped and S,N-codoped carbons as
743 catalysts for electrochemical transformation CO2 (Li et al., 2016). The team found that
41
744 S,N-doped carbon has greater FE for conversion to CO and CH4 compared to S-doped
745 counterpart. The former revealed a maximum FE of 11.3% and 0.18%. for CO and CH4
746 production, respectively. The S,N-nanoporous carbon was better at decreasing the
748 Two-dimensional (2D) boron nitride nanosheets or meshes (BNs) doped with beryllium
749 have been the subject of research by Azofra and coworkerss (Azofra et al., 2016).
751 network significantly lowered the energy barriers for CO2 fixation and the initial steps
752 in hydrogenation. As a result of the exceptionally large hole created on the Be-doped
753 surface environment, CO2 was spontaneously absorbed (in terms of the Gibbs free
755 Despite all the accomplishments in CO2RR, it is worth to mention that high progress is
756 still needed in the design of catalysts based on metal and non-metal for achieving a
757 more performant CO2 electrocatalytic transformation. The main challenge being the
758 energy requirement for reducing CO2 (due to its stability), an electrochemical potential
759 [E°] of 1.9 V vs normal hydrogen electrode (NHE) at pH. Furthermore, due to the high
760 reactivity of the one-electron reduction product (CO2•-), regulating product selectivity is
762 transfer, the potential for CO2 reduction is greatly reduced, resulting in a variety of
763 products depending on the number of electrons involved (Maeda, 2019). Finally, the
764 quantum yield and current density reported thus far are extremely low, particularly for
765 C2+ products, making commercial production extremely difficult. As a result, extensive
766 and intensive research is required to overcome all of the aforementioned challenges.
42
768 Photoelectrochemical (PEC) conversion of CO2 can be described as a procedure
769 mimicking the artificial photosynthesis technique. This method has been employed for
770 the conversion of CO2 to formate, formaldehyde, formic acid, methane, methanol, and
773 electrode surface under an applied electric field, thus reducing CO2 in the process
775 During the 80s, Fisher et al. (Fisher and Eisenberg, 2002)and Inoue et al. (Inoue et al.,
776 1979) studied the electrocatalytic reduction reactions that gave rise to studies of photo-
778 study using various metals and metal oxides requires large over potentials and suffers
779 from hydrogen evolution (Benson et al., 2009; Qiao et al., 2014; Shen et al., 2015).
780 Since then, it has become a promising technique for CO2 conversion. In 2020, Arai et
781 al. (Arai et al., 2010) reported the PEC conversion of CO2 into HCOOH using a p-InP-
782 Zn photocathode which has been modified with a Ru polymer. The conversion was
783 achieved in the presence of visible light/ solar energy. Similarly, Kumaravel and his
784 coleagues (Kumaravel et al., 2020) also synthesized a range of fuels and chemicals
785 (Formic acid, methanol, ethanol, methane and ethane) from CO2. They employed a
786 catalyst-embedded microfluidic PEC reactor for the synthesis. The catalyst was
787 activated in the presence of visible light irradiation. The schematic of the conversion is
43
789
790 Figure 13: PEC Conversion of CO2 into value-added product using solar irradiation.
791 Adapted with permission from reference (Kumaravel et al., 2020). Copyright [2020],
793
794 According to a recent study by Karim et al. (Rezaul Karim et al., 2018), single electron
797 vs NHE) and formate (E = -0.61 V vs NHE) while the 6e- pathway leads to the
798 production of methanol (E= -0.38 V vs NHE) (Qiao et al., 2014; Rezaul Karim et al.,
799 2018). The band varies depending on the type of metal coordination as shown in
802 activates the metal center, thus forming an oxidation-reduction environment. During
803 this process, electrons get excited from the VB to CB, leaving behind positive holes
804 that oxidize water on the surface to form OH- radicals with strong oxidative
805 decomposing power. Excited electrons from the CB get accepted by CO2 to form •CO2
806 radicals that react with the intermediate products of the oxidative reactions forming one
807 or more of the products in Table 2 depending on the bandgap of the photocatalysts
808 (Ekambaram et al., 2007). Metal oxides present desirable bandgap and bandgap edge
44
809 positions, and their electronic structure plays a major role where the VB is fully filled
811 Table 2: The oxidation-reduction potential at pH=7 vs NHE and half-cell reactions
CO2 + 2e− → •
CO2 -1.90
•
CO2 + 2H++ 2e− → HCOOH -0.61
•
CO2 + 2H++ 2e− → CO + H2O -0.53
•
CO2+ 4H++ 4e− → HCHO + H2O -0.48
•
CO2 + 6H++ 6e− → CH3OH + H2O -0.38
•
CO2 + 8H++ 8e−→ CH4+ 2H2O -0.24
H+ + 2e− → H2 -0.42
813
815 synergistic photo-electrochemical catalysis system (Huang et al., 2017; Rezaul Karim
816 et al., 2018). The catalytic activity of a photocatalyst towards CO2 reduction depends
817 on its band gap and appropriate positioning of the conduction and valence bands
818 (Handoko and Tang, 2013). Doping semiconductors have extensively been explored in
819 the literature to enable visible-light-activated photocatalysis and the usage efficiency of
820 solar light. Several published studies have used titanium for photocatalytic applications
821 for organic compound decompositions due to its ability to oxidize organic and
822 inorganic substances in water and air through a redox reaction (Xie et al., 2018). TiO2
823 and ZnO absorb ultraviolet light of the solar spectrum due to their large bandgap’s
45
824 energy (~380 nm and ~360 nm respectively) (Liu et al., 2010). The bandgap (the
825 distance between the valence band and conduction band) of each semiconductor
826 determines are shown in Figure 14, it shows the relationship between the bandgap and
827 EMF at pH 7. Several alternative semiconductors such as Fe2O3, ZnO, WO3, SrTiO3,
828 NaTaO3, CdS, Ag3PO4, BiPO4 and g-C3N4 have been reported to have promising
829 electro-photocatalytic performance (Kubacka et al., 2011; L. White et al., 2015; Liu et
831
832 Figure 14: Band gap energy of various semiconductors (Wang et al., 2009). Copy right
833 2019
834 More recently, Nandal et al. (Nandal et al., 2022) performed a selective conversion of
835 CO2 into ethanol via a PEC method (see Figure 15). The catalytic systems involve a
836 mixture of Graphene oxide/Copper oxide and a copper-based MOF (GO/CuxO, Cu-
837 MOF). The catalytic system was solar-driven and activated in the presence of sunlight.
838 They reported an ethanol yield of 162 µM cm-1. Most researchers believe that the PEC
839 reduction technique is not only efficient, but it is also greener and sustainable. Other
46
840 PEC for CO2 reduction surveyed in this section are listed in the supplementary file
842
843 Figure 15: PEC reduction of CO2 to ethanol. Adapted with permission from reference
847 considering the large amount of CO2 generated by the industrial processes and the
848 associated environmental degradation. It facilitates the production of clean fuels and
849 hydrocarbons such as CO, CH4, and C2+ products (Fan and Tahir, 2022). The
850 hydrogenation reaction can proceed via thermal catalysis, photocatalysis, and
851 photothermal catalysis. While thermal and photocatalytic processes require the action
852 of heat and light energies respectively, the photothermal catalytic approach combines
853 both in a single reaction system. In photothermal catalytic systems, the CO2
854 transformation reaction is driven by light energy into a thermally inspired catalytic
855 reaction system or utilizing the photoinduced thermal effect of a catalyst. The light
856 energy initiates a photoinduced redox reaction and a photoinduced thermally driven
857 reaction (Oladoye et al., 2021). For instance, the exposure of a transition metal oxide-
858 based catalyst to a sufficient amount of light energy causes an excitation of photons that
859 produces electron-hole pairs in the conduction band (CB) and valence band (VB)
47
860 (Figure 16). These highly energetic electrons enter the anti-bonding orbitals of the
861 reactants and activate the reactant molecules to stimulate a redox reaction. In other
862 instances, the absorbed photon could be converted into heat energy (Liu et al., 2022).
863 At an activated temperature, the active sites on the surface of the photothermal catalytic
864 material could induce a thermally driven catalytic reaction. Several semiconducting
865 metal oxides with tunable porosity and defective active sites have been designed for the
867
48
869 Table 3: Examples of photothermal catalysts for the hydrogenation of CO2 in fuels and hydrocarbons
(mmol·g-1·h-1)
bar) methanol
to the production of
hydrogenation performance
49
120 μm- 300 W Xe Pressure (1 CO 1780 Strong light-absorption ability (Shen et
SiNCs@Co arc lamp bar), power over the entire solar spectrum al., 2022)
conversion rate
10 min
Nb2C- 300 W Xe 36-sun CO 8.5 mol·g-1·h-1 Ni NPs greatly enhance the (Wu et al.,
50
nanosheet- arc lamp illumination activation of H2 and accelerate 2021)
properties
α-MoC and β- 300 W 130 °C, 5 h CO and CH4 25.8 mL·g-1·h- dissociation of light-promoted (Zhao et
1
Mo2C mixed xenon lamp surface intermediate (formate al., 2021)
enhanced activity of
51
hydrogenation
Pd2Cu alloy UV-Vis (350 150 °C C2H5OH 4.1 Pd active sites utilize the (Elavarasan
SNAs support 300 W Xe Pressure (1 CO 0.433 Photoexcited electron-hole pairs (D. Zhang
(2.5 W
cm−2)
Co@AAO Sunlight Power CO 1666 A dual-pass AAO with a large (D. Lou et
52
performance
step.
870 SPR-Surface plasmon resonance; MOF-Metal organic frameworks; SiNCs-Silica nanocones, NPs-Nanoparticles, SNA-Silicon
53
872 3.5 Thermochemical transformation of CO2
873 The thermochemical transformation of CO2 is an interesting strategy that has been gainfully
874 explored for the reduction, splitting, or reforming of CO2 to fuels/products such as carbon
875 monoxide, methanol, methane, and dimethyl ether in the presence of suitable catalysts (Fuqiang
876 et al., 2017; Roy et al., 2018). Although, there are many high-temperature thermochemical
877 energy conversion processes for energy storage, energy conversion and water splitting reactions
878 (Ali et al., 2020; Cheng et al., 2021; Lorentzou et al., 2014; Wu et al., 2018), the solar-induced
879 thermochemical cyclic method is been emphasized as an outstanding approach for the efficient
880 conversion of CO2 (Table 4). This approach has produced intriguing results from several the
881 experimental tests/designs conducted in different solar reactor models, which has given rise to
882 more intensified research and technology development in this field (Guene Lougou et al., 2020).
883 Many studies have shown that the solar thermochemical CO2‐transformation is a promising
884 solution for solar energy harvesting and storage using materials like Fe‐Ni alloy embedded in a
885 perovskite (Hu et al., 2021), Ca- and Al-doped nSmMnO3 perovskite (Gao et al., 2022), and
886 ZrO2-supported NiFe2O4 (Shuai et al., 2021), to convert CO2 into CO with a 99% conversion rate
887 of 381 mL g−1 min−1 at 850°C, high conversion of 595.56 mmol/g, and 37.89 mL of CO yield in
888 46 min, respectively. Meanwhile, Ni-doped CaCO3 transformed 86.6% of CO2 into syngas along
889 with CH4 in a fixed bed photo-reactor (Teng et al., 2022). The solar thermochemical process
890 follows a two-step dissociation approach wherein the first step is an endothermic process that
891 involves the production of O2 by partially reducing a perovskite or metal oxide (MO) to a non-
892 stoichiometric oxide with the assistance of high-temperature solar heat (Eq. 7). The second step
893 usually occurs at low temperature via an exothermic process. At this stage, the reduced
54
894 perovskite or MO is oxidized with H2O and/or CO2 giving rise to the formation of H2 and/or CO
895 (Eq. 8-9), together with the stoichiometric oxide (Chen et al., 2022; Nair and Abanades, 2018).
899 In the same vein, the photo-thermochemical (PTC) model produces photoinduced oxygen
900 vacancies on the surface of the MO catalyst after exposure to UV-light energy (Eq. 10). Further,
901 the oxygen vacancies are used to reduce CO2 at temperatures below 873 K, thereby producing
905 By exploring the Fisher Tropsh (F-T) technology, solar-assisted MOs can also afford carbon-free
906 alternative fuels i.e. production of H2 that can be used directly as a fuel or converted into other
907 value-added chemicals (Figure 17) (G. Takalkar and Bhosale, 2019). Besides, other
908 heterogeneous catalysts that change dynamically during the thermochemical reactions as well as
909 possessing catalytic active sites and definite reaction pathways have been identified for CO2
910 transformation (Feng et al., 2021). For example, the loading of IrOx onto LaFeO3 almost double
911 the maximal CO release rate and increases the maximal O2 evolution rate 0.5 times i.e. a 5-fold
912 CO production (Jiang et al., 2016); meanwhile, the sol-gel synthesized CoxFe3−xO achieved
913 higher CO generation rate via CO2 splitting (Takalkar and Bhosale, 2019).
55
914
915 Figure 17: Solar-assisted thermochemical process for the production of hydrogen and syn gas by
916 the F-T approach. Reproduced from (Takalkar and Bhosale, 2019). Copyright 2019 Elsevier Ltd
thermochemical of the
material process
μmol/g
56
gcat min−1
2020)
Steinfeld, 2011)
2014)
CO/H2
0.4 m/s
(1.6-fold
increase)
Haeussler, 2021)
1600 °C
57
rate of 40 sccm
918
920 Non-thermal plasma (NTP) catalysis is an auspicious technology that is gaining significance in
921 C1 chemistry. It has been postulated as an effective alternative to conventional thermal catalysis
922 due to the unique interplay between the plasma and heterogeneous catalyst. Thus, this approach
923 could overcome the kinetic limitation of conventional thermal catalysis without the need for any
924 external heat sources (Parastaev et al., 2018). Typically, this process employs an electric
925 discharge of partially ionized gas made up of electrons possessing average electron energy in the
926 range of 1-10 eV (Chen et al., 2020c). These electrons have the capacity to activate stable C1
927 molecules as well as denature their bonding structures in the gaseous phase under moderate
928 conditions, thereby producing atoms, ions, and vibrationally-excited reactive species (Mehta et
929 al., 2019). The NTP catalytic approach has the potential of accelerating many energy-intensive
930 and kinetically and/or thermodynamically limited heterogeneous catalytic reactions such as CO 2
931 hydrogenation (Song et al., 2022), water–gas shift reaction (Xu et al., 2019), hydrocarbon
932 reformation (Wu et al., 2018) and CO2 splitting (Chen et al., 2019), amongst others.
933 In Chen et al. (Chen et al., 2020a) Ni NPs supported on silicalite-1 zeolites with suitable active
934 sites and pore structures were comparably investigated in CO2 hydrogenation under the thermal
935 and NTP conditions. The comparative investigation showed that the hierarchical meso-
936 microporous structures and the associated well-dispersion of Ni species were highly beneficial to
937 the catalysis under the NTP conditions, especially at relatively high voltages. Using a dielectric
938 barrier discharge (DBD) reactor at 150°C, Zeng and Tu (Zeng and Tu, 2017) showed that the
939 presence of Ni in the reactor and injection of argon (up to 60%) in the reaction promotes the
58
940 conversion of CO2 to CO and CH4 (selectively, 85%) and the energy efficiency of the plasma
941 process. In the reactor, Ar decelerates the breakdown voltage of the feed gas and accelerates
942 charge transfer through the reactor. Furthermore, the NTP-activated hydrogenation of CO2 in the
943 presence of Ru/MgAl layered double hydroxide catalysts (having low activation energy)
944 provided ∼84 % of CH4 yield at relatively low temperatures (Xu et al., 2020). In another
945 approach, a MOF-supported Ni catalyst (15Ni/UiO-66) presented 99% selectively towards the
946 conversion of CO2 to CH4 under NTP conditions and the catalyst was still stable after 20h of
947 reaction (H. Chen et al., 2020b). The experimental investigations of Nizio et al. (Nizio et al.,
948 2016) revealed that ceria and zirconia-promoted Ni-containing hydrotalcite-derived catalysts
949 could yield high CH4 under NTP conditions and temperatures < 330 °C.
951 The discovery and development of efficient technologies for the catalytic transformation of CO2
952 is an exciting area of research (Ramírez-Valencia et al., 2021) as, in parallele of lowering the
953 atmospheric CO2, this would allow the access to essential chemicals and fuels such as formic
954 acid (HCOOH), methanol (CH3OH), carbon monoxide (CO), ethanol (CH3CH2OH) and methane
955 (CH4), among others, that are presently produced from fossil resources. Figure 18 describes the
59
957
958 Figure 18: Potential products from CO2 conversion (Ramírez-Valencia et al., 2021).
959
961 Among the myriad of products accessible from CO2, formic acid is a valuable intermediate
962 chemical and also used in the domain of energy for reversible hydrogen storage (Figure 19)
963 (Moret et al., 2014). Formic acid is of great importance in textile industries, pharmaceuticals and
964 food companies dues to its acidic nature and reducing properties (Afshar, 2014). In particular,
965 the leather and tanning industries are one of the largest global consumers of formic acid (Pérez-
966 Fortes et al., 2016). Other critical applications of formic acid include preservatives and
967 antibacterial agents in animal feeds (Anadón et al., 2006). Therefore, formic acid is a highly
60
969
970 Figure 19: Reversible hydrogenation of CO2 to formic acid (Li et al., 2002)
971 Several synthetic methods for the conversion of CO2 to formic acid have been explored. This
973 catalytic reduction (Asiri and Lichtfouse, 2020; Hu et al., 2013; Lu et al., 2014; Scibioh and
974 Viswanathan, 2004; Yadav et al., 2012). Catalytic conversion of CO2 to formic acid was first
975 reported by Inoue (Inoue et al., 1976) in 1976. Since then, CO2 has become an attractive starting
976 molecule for the synthesis of formic acid. Compared to other methodologies, the catalytic
977 reduction method offers higher selectivity to targeted product, thus minimizing the risk of
978 unwanted CO as a byproduct that may poison the fuel. However, the selectivity of the preferred
979 product depends on the reduction method and the nature of the catalyst used.
980
981 Weilhard et al. reported the synthesis of formic acid from CO2 feedstock using an ionic liquid as
982 an alternative to expensive precious metal-based catalytic systems (Weilhard et al., 2021). As
61
983 shown in Scheme 3, the synthetic strategy consists of a mixture of CO2 and H2 in 1:1 molar ratio.
986 very high catalytic efficiency with turnover numbers (TON) and turnover frequency (TOF) of 8 x
988
Dioxane/water
990
991 Moret et al. (Moret et al., 2014) applied a ruthenium-based homogeneous catalytic system to
992 convert CO2 feedstock into formic acid directly. The catalytic conversion was performed
993 following the catalytic life cycle described in Figure 20. The reaction was performed at 40 °C
994 using a mixture of water and dimethyl sulphoxide (DMSO) as the solvent.
995
996 Figure 20: Catalytic cycle and hydrogenation of CO2 into formic acid over [HRu(PTA)4Cl]
997 They reported that the catalytic system is highly sustainable than the existing routes. In addition,
998 it is highly recyclable and gives a higher yield of the target product. More recently, Mubarak et
62
999 al. (Mubarak et al., 2022) published a greener approach via photocatalytic transformation of CO2
1000 into formic acid. The catalytic procedure involves the breaking down of CO2 into valuable
1001 chemicals in the presence of sunlight and Ag-doped TiO2 NPs. This catalytic system iss highly
1002 selective with HCOOH as the single product, formic acid being obtained at a yield of 193 μM
1005 In the series of products derived from CO2, carbon monoxide (CO) is one of the most critical
1006 chemicals and feedstock. It is a primary raw material along with H2 gas for syngas production
1007 via Fischer-Tropsch synthesis to produce liquid fuels (Zeng et al., 2018). It is also valuable for
1008 the metallurgical industries (Liu et al., 2017). Most of the efforts concerning CO2 reduction into
1009 CO efforts can be listed in two major categories, namely thermochemical reduction and
1010 electrochemical reduction (Porosoff et al., 2016). CO2 being a very stable molecule, it requires a
1011 very high temperature (> 1200 °C) to decompose to CO (Equation 13). Therefore, this procedure
1014 Various catalytic materials have been employed to reduce the energy cost of the catalytic
1015 conversion of CO2 to CO. For example, Au nanocatalysts have shown tremendous potential in
1016 converting CO2 to CO with a remarkable selectivity (Qi et al., 2019). Qi et al. (Qi et al., 2021)
1017 reported a successful electrochemical conversion of CO2 to CO over Au nanocatalyst. Wen et al.
1018 (Wen et al., 2018) used a nanoporous Au leave electrocatalyst prepared by a de-alloying
1019 technique to reduce CO2 to CO, achieving a very high catalytic efficiency to CO (90% FE).
1020 While gold-based catalysts are the most successful catalysts for the conversion of CO2 to CO, the
1021 high cost of Au may prevent large-scale industrial applications (Ma et al., 2016b) which requires
63
1022 to find low-cost alternative catalysts. Metallic Ag is gaining considerable attention for CO2
1023 conversion because of its significantly lower cost compared to Au and also high selectivity to
1024 CO (Ma et al., 2016b). This has been demonstrated by Ma et al. (Ma et al., 2016b) who
1026 and polycrystalline Ag. The proposed reduction mechanism is described in Scheme 3. The
1027 obtained catalytic efficiency was 80% Faradic efficiency which is comparable to the results
1029
1030
1031 Scheme 3: Mechanism of the conversion of CO2 to CO via electrochemical reduction proposed
1033 Dong et al. (Dong et al., 2021) reported the use of low-cost ZnOHF nanorods as a substitute to
1034 expensive Au and Ag metals for the reduction of CO2 to CO by electrochemical reduction. They
1035 obtained a CO Faraday efficiency of 76.4% with the CO current density of 57.53mA/Cm 2. It is
1036 worth mentioning that the catalytic efficiency of this Zn-based catalyst is not as high as Au and
1037 Ag-based ones, but the price is significantly lower which may open avenues for a large-scale
64
1040 The conversion of CO2 to MeOH has been drawing tremendous attention lately because of the
1041 vast applications of methanol (Olah, 2013). Methanol is a multipurpose chemical feedstock
1042 (Offermanns et al., 2014) for producing various value-added products and fuel. The current
1043 production rate of methanol is 98 million metric tons per year (Kajaste et al., 2018). It is also an
1044 important hydrogen carrier. The various applications and Carbon cycle of methanol are described
1046
1048 Considering the vast application of methanol (Figure 16), conversion of CO2 to MeOH is a
1049 crucial research area. The production of methanol has been dominated by the conversion of
1050 syngas (a mixture of CO and H2) and CO2 (Nieminen et al., 2019). From the mechanistic point of
1051 view, the synthetic pathway may occur through direct CO2 hydrogenation (eq 14) or via reverse
1052 water gas shift (RWGS) (eq 15) followed by CO hydrogenation (eq 16) (Gaikwad et al., 2020).
65
1053 A challenge often characterized by this synthetic procedure is that it is performed at very high
1054 temperature and pressure (Bertau et al., 2014). In addition, it also suffers from poor selectivity
1056 To improve this procedure, various homogeneous and heterogeneous catalytic hydrogenation
1058 have been explored with promising results (Navarro-Jaén et al., 2021).
1063 heterogeneous catalyst at 513-533K temperature and 50-100 bar pressure (Bart and Sneeden,
1064 1987). It is worth mentioning that Cu alone exhibits a very low activity; to enhance the activity,
1065 the Cu catalyst is usually tethered on support or promoters such as Al2O3, ZnO, and ZrO2. In
1066 most cases, the synergy between the Cu catalyst and the promoters only achieves a conversion of
1067 approximately 30% due to a fast deactivation of the Cu catalyst (Navarro-Jaén et al., 2021).
1068 Although the activity is low, Cu-based catalysts are very cheap thus justifying their
1069 predominance in methanol synthesis. However, the growing demand for methanol induced
1071 For instance, Pd-based heterogeneous catalytic systems have been used to improve the synthesis
1072 of methanol. Díez-Ramírez et al. (Díez-Ramírez et al., 2016) reported the application of Pd/ZnO
66
1073 catalyst to convert CO2 to methanol under atmospheric conditions, observing a low conversion
1074 rate due to the side reactions from coke formation and water from the (RWGS) reaction (eq 15).
1075 Various attempts have been made to suppress side reactions from the (RWGS) reaction.
1076 Fiordaliso et al. (Fiordaliso et al., 2015) reported the use of supported Pd-Ga bimetallic catalysts,
1077 specifically, GaPd2/SiO2 for converting CO2 to methanol, at low CO2 pressure, and with a MeOH
1078 production rate of 0.12 mol%. Although the catalytic activity of this GaPd2/SiO2 was better than
1079 in previous studies, the major drawback is the Pd cost, thus limiting practical applications.
1080 Ahmad et al. (Ahmad and Upadhyayula, 2019), developed an efficient, cheaper bimetallic
1081 catalyst Ga3Ni5 for CO2 conversion and compared their results with those of Fiordaliso et al.
1082 (Fiordaliso et al., 2015). The activity of the Ga3Ni5 catalyst was found to be higher under the
1083 same experimental conditions, showing a MeOH conversion rate of 38.85 µmol gcat-1 min1
1085 While Ga, Zn, Cu, Al, Pd, Ni, Pt-based catalysts have been widely tested for MeOH synthesis,
1086 To motivate a cost-effective catalyst, scientists are considering an alternative to Pt and noble
1087 metals. For instance, CeO2-based catalyst is getting attention due to its abundant oxygen and
4+ 3+
1088 reversible valence charges (Ce and Ce ) (F. Wang et al., 2016). This potential was
1089 demonstrated by Chang et al. (Chang et al., 2021), who applied a CeO2-based catalyst Cu/ZnO–
1091 Perovskites have been also identified as low-cost approach for CO2 conversion to methanol. This
1092 is described by Zhan et al. (Zhan et al., 2014), who employed a perovskite-based catalyst to
1093 produce methanol from CO2. They reported an excellent catalytic efficiency with impressive
67
1095 Apart from heterogeneous-based catalytic systems, homogeneous catalysts have been also
1096 reported for methanol synthesis from CO2. Kothandaraman et al. (Kothandaraman et al., 2016)
1097 developed a Ru-based homogeneous catalytic system that showed a very good yield (79 %) in
1098 methanol. In addition, this catalyst also shows excellent recyclability with over five runs without
1099 a significant loss in activity. Zhou et al. (Zhou et al., 2022) published the use of a ruthenium
1100 pincer complex [RuH2(Me2PCH2SiMe2)2NH(CO)] which led to a good catalytic efficiency from
1102 Application of enzymes to reduce CO2 to methanol is also getting considerable attention. Zhang
1103 et al. (Z. Zhang et al., 2021) reported the enzymatic conversion of CO2 to methanol in the
1104 presence of nicotinamide adenine dinucleotide (NADH). The enzyme was immobilized on
1105 porous a MOF (ZIF-8) to improve CO2 concentration. The prepared NADH/ZIF-8 composite
1106 was placed in the bio-electrochemical cell. The methanol conversion rate was 822 μmol g-1 h-1,
1107 and the catalytic system was found stable and environmentally friendly.
1109 Ethanol is a valuable chemical in our daily life. In contrast to methanol, ethanol has some distinct
1110 advantages, such as higher energy density, lower toxicity, and transportation safety (D. Wang et
1111 al., 2016). As a result of these advantages, it intervenes in a wide range of domestic and
1112 industrial applications, including chemical raw material for chemical synthesis, in disinfectants,
1113 pharmaceuticals and cosmetics, plasticizers, polishes and dyes (An et al., 2021). In addition to its
1114 domestic and industrial applications, ethanol is a necessary fuel and hydrogen carrier (Pang et
1115 al., 2019). The United States are the world largest ethanol producer, accounting for 59% of
1116 global production. This is followed by brazil with 28% while the rest of the world accounts for
68
1118
1119
1120 Figure 22: Ethanol production in various countries for the year 2020 (Cooper et al., 2021).
1121
1122 The high energy density and the compatibility of ethanol with existing and modern combustion
1123 engines make it a vital fuel. Ethanol is usually produced from the fermentation of starch and
1124 cellulose feedstock (Ali et al., 2022; Wang et al., 2021). With the population and increasing
1125 demand for food, the search for an alternative strategy for ethanol synthesis is highly urgent.
1126 That’s why hydrogenation of CO2 to ethanol has drawn a lot of attention.
1127 Several catalytic systems have been used for CO2 conversion to ethanol, notably Pt, Co, Fe, and
1128 Cu systems (Zheng et al., 2019). Cu-based catalysts have been used for hydrogenating CO2 to
1129 ethanol due to their success in converting CO2 to methanol; however, the selectivity to ethanol is
1130 very low (Cao et al., 2016). Recently, it has been found out that ethanol selectivity can be
1131 improved by using Cu1 instead of Cuo. Despite this success, the Cu-catalyst suffers from long-
1132 term use due to instability under the reaction conditions and operating conditions.
1133 Other catalytic systems have been tested to improve the selectivity and yield of ethanol from
1134 CO2, including Pt-based catalysts and their alloys; while good yield was obtained with these
69
1135 catalysts, the high cost of the catalytic materials is an issue that limits their practical application
1136 (Choi and Liu, 2009). Lou et al. (Y. Lou et al., 2021) successfully applied a Pd/CeO2 catalyst to
1137 convert CO2 to ethanol. They reported a space yield time of 45.6 g ethanol gPd-1 h-1. Together with
1138 an excellent selectivity of 99.2 %. Zhang et al. (F. Zhang et al., 2021) reported the application of
1139 supported sodium-modified Rh NPs in zeolite to hydrogenate CO2 to ethanol that showed a CO2
1140 conversion rate of 10 % and a selectivity of 24% to ethanol. Zheng et al. (Zheng et al., 2019)
1141 reported the conversion of CO2 to ethanol using a LaCo1-xGaxO3 composite perovskite catalyst
1142 with a very good catalytic CO2 conversion of 9.8% and a 88.1% selectivity to ethanol.
1143 Another challenge is catalyst separation from the product and recyclability; therefore, extensive
1144 research efforts are is devoted to develop effective and highly reusable catalysts in order to
1147 Methane (CH4) is a C1 compound primarily used as fuel. Moreover, it is also used as a feedstock
1148 for chemical synthesis, especially as a synthetic natural gas and is also a major source of
1149 hydrogen storage (Koytsoumpa and Karellas, 2018). Therefore, the hydrogenation of CO2 to
1150 methane is an effective strategy for CO2 capture, utilization and conversion to valuable
1151 chemicals. Catalytic hydrogenation is the most widely studied route for CO2 conversion to
1152 methane (Koytsoumpa and Karellas, 2018). This procedure is called methanation or Sabatier
1153 reaction (Lu et al., 2016) as it was first reported in 1902 by Sabatier et al. (Sabatier, 1902).
1154 According to the literature, two possible mechanisms have been proposed. The first mechanism
1155 is direct methanation (eq 17) which involves the reaction of CO2 and H2 with a nominal ratio of
1156 1:4, respectively, while the second mechanism is the reaction of CO and H2 in a ratio of 1:3
1157 respectively (equation 2) (Dias and Perez-Lopez, 2021). As shown in equations 17 and 18, the
70
1158 conversion of CO2 to methane is a highly exothermic reaction which can proceed at a relatively
1159 lower temperature. However, given the thermodynamic stability of CO2 molecules due to the
1160 presence of delocalized bonds, the conversion of CO2 to methane remains a serious challenge
1164 To overcome this challenge and permit activation of CO2 at lower temperatures, highly active
1165 catalytic materials with good selectivity to methane and low deactivation must be developed.
1166 CO2 methanation has been performed with transition metals such as Co, Ni, Fe, Cu, Ru, Rh, and
1167 Pd (Y.-T. Li et al., 2022). However, due to the cost and limited availability, these noble metals
1168 are not widely employed for methane synthesis, especially on an industrial scale. Therefore, a
1169 new generation of low-cost and efficient catalysts have been investigated for methanation
1170 reaction. Among them, Ni, Fe, and Co-based catalysts have been widely studied because of their
1171 low cost and catalytic efficiency for CO2 methanation. Ni-based catalytic materials are the most
1172 studied for CO2 methanation due to their low cost and availability (Fukuhara et al., 2017). Dias
1173 et al. (Dias and Perez-Lopez, 2021) reported the conversion of CO2 to Methanol using Ni/SiO2
1174 catalyst promoted by Fe, Co, and Zn. They obtained a very high CO2 conversion (73 %) and
1176 It is generally believed that catalytic behavior depends on several factors, such as surface
1177 properties, metal support interaction, etc. Researchers are currently developing further catalytic
1178 supports to enhance the performance of catalytically active metal species. This was recently
1179 demonstrated by Li et al. (Y.-T. Li et al., 2022) who used a metal-organic framework (MOF) as a
1180 support to enhance the catalytic activity of NiFe for the hydrogenation of CO2 to methane. MOFs
71
1181 are highly porous coordination polymers that are characterized by high surface areas. As a result,
1182 they can enhance the catalytic activity in comparison to other traditional supports.
1183 Li et al. (Y.-T. Li et al., 2022) obtained a very high CO2 conversion (72.3 %) and impressive
1184 selectivity (99.3 %) with the NiFe/MOF catalytic material. Similarly, Hu et al. (Hu et al., 2022)
1185 also established the role of metal-support interaction using Ni supported on CeO2 nanofibers.
1186 The oxygen vacancy enhanced the catalytic efficiency of Ni in CeO2. The prepared catalyst
1187 recorded 82.3 % conversion of CO2 at low temperatures (250 oC - 300 oC). In addition, they also
1188 reported impressive stability of the catalyst up to 60h. This result can be attributed to synergistic
1191 The ease of availability of CO2 has provided an opportunity for energy production and
1192 commodity products for the economy. Among other compounds which could be generated from
1193 CO2 is formaldehyde (FMD). Formaldehyde, with the chemical formula HCHO, is a substance
1194 that has gained wide usage by millions of people, directly or indirectly, across the globe due to
1195 its versatility (Rauch et al., 2019; Subasi 2020; Usanmaz et al., 2002). Generally, FMD is
1196 produced industrially via three stages. These include the production of synthesis gas (a mixture
1197 of water gas, carbon(iv) oxide and hydrocarbons) by steam reforming natural gas, synthesis of
1198 MeOH, and the generation of FMD through the partial oxidation of MeOH (Bahmanpour et al.,
1199 2015). All these methods are energy-intensive procedures (Heim et al., 2017; Nguyen et al.,
1200 2020). Moreover, during these processes, CO2 was contained in most of the products formed
1202
72
1203 In a bid to curb the emission of CO2, scientists have developed methods that could possibly
1204 convert CO2 into useful products. Of these methods is the production of FMD (Friedlingstein
1205 2015; Sümbelli et al., 2019). Although the generation of FMD from CO2 is quite challenging,
1206 frantic efforts are being made to overcome this fall-out as CO2 is resistant to chemical changes
1207 (Rauch et al., 2019). One of those efforts is the conversion of CO2 at ambient temperature
1208 yielding the formation of FMD. The CO2 reacts with triphenylsilane in the presence of a
1209 magnesium catalyst, resulting in an intermediate product formed, known as bis(silyl)acetal. This
1210 product is then converted to FMD in the presence of cesium fluoride (Eq 19) (Rauch et al.,
1211 2019).
eq 19
1212
1213 Another way that allows to convert CO2 into FMD is the hydrogenation process, which different
1215 homogeneous) involved (Dang et al., 2019; Nguyen et al., 2020; Studt et al., 2015). The most
1216 prevalent catalysts used for this reaction are heterogeneous Cu-based, such as Cu-ZnO-Al2O3
1217 (Kobl et al., 2016) or Pt-promoted Cu/SiO2 (Lee et al., 2001). This latter converts CO2 to FMD at
1218 150oC for 120 min at 600 kPa. However, Lee and co-partners indicated some issues encountered
1219 with this catalyst (Lee et al., 2001). Moreso, a mesoporous graphitic carbon nitride (g-C3N4)
1220 material with specific crystallinity and surface area, has been shown to have a significant effect
1221 on the conversion of CO2 to FMD. Since this discovery other photocatalytic mesoporous
1222 nanomaterials have been explored for the photocatalytic reduction of CO2 (Park et al., 2012; Xu
73
1224 4.7 Hydrogen
1225 Aside from FMD, a critical commodity product is hydrogen, being also a fuel, that could be
1226 generated from CO2 via a process called carbon dioxide reforming. This process, represented on
1227 equation 20, consumes CO2 and methane (CH4), two greenhouse gases. This process requires
1228 high temperature, and as such, it is highly endothermic and also favored by low pressure (Wang
1231 Fan and coworkers reported this process has been catalyzed over a Ni-Co/MgO-ZrO2
1232 nanomaterial, (Fan et al., 2011). The effect of Ni-based catalysts on the CO2 reforming of
1233 methane had earlier been investigated by Takano et al., (1994). This process produces green
1234 hydrogen a quality of hydrogen which is otherwise generated via the electrolysis of water using a
1235 renewable energy source (Acar and Dincer 2022; Luo et al., 2022). This process could solve the
1236 problem of CO2 emission during hydrogen production from classical techniques (Madadi
1237 Avargani et al., 2022). However, there is a fall-out associated with the production of green
1238 hydrogen via the method mentioned above; it involves the release of carbon(ii)oxide as a by-
1239 product. Although CO has a significant value as a fuel and is crucial in manufacturing other
1240 chemicals, it has been known over the years for its highly hazardous nature (Prockop and
1241 Chichkova 2007; Rose et al., 2017; Wu and Juurlink 2014) and toxicity. Studies should focus
1242 more on the production of hydrogen from CO2, as this process is not well explored. This will go
1245 The hydrogenation of CO2 into high molecular weight hydrocarbons such as lower olefins and
1246 gasoline has become increasingly popular due to the necessity to use non-fossil carbon resources
74
1247 as alternative to fossil oil for a sustainable production of liquid fuels and chemicals (Jiao et al.,
1248 2016; Yang et al., 2017; Zhong et al., 2016). This process requires input energy to form C-H
1249 bonds and to break C=O bonds (Dang et al., 2019) and efficient catalysts.
1250 Processes based on different catalysts and producing unique products distribution have been
1251 developed in the past decades. A commonly used process is Fischer-Tropsch synthesis (FTS) that
1253 product distribution, which restricts to C2-C4 hydrocarbon fraction of about 56.7% selectivity for
1254 desired lower olefins (C2= - C4=) with 29.2% of undesired CH4 selectivity (Scheme 4). If
1255 selective formation of lower olefins and low CH4 formation can be achieved by proper catalyst
1256 modification and optimization of reaction conditions (Chen et al., 2015; Liu et al., 2015; Lu et
1257 al., 2014; Zhong et al., 2016; Zhou et al., 2015) achieve high selectivity for lower olefins and
1258 high stability, requires to develop new catalysts that operate away from Anderson–Schulz–Flory
1259 (ASF) distribution (Numpilai et al., 2017; Visconti et al., 2017; Zhang et al., 2015).
1260
1261 Scheme 4. Typical Reaction processes related to reduction of Syngas to hydrocarbons through
1262 Fischer-Tropsch synthesis and methanol-to-Olefins Technology. (Adapted without change from
75
1264 In recent years, significant progress in the development of new catalysts that increase the
1265 selectivity of product formed has been made. For instance bifunctional catalysts such as
1266 oxide/zeolite have been shown to increase selectivity for lower olefins (C2=–C4=) with significant
1267 inhibition of CH4 formation (Jiao et al., 2016; Li et al., 2019; Liu et al., 2017). Bifunctional
1268 catalysts typically consist of oxides and zeolites, which activate CO2 and catalyze C-C coupling,
1269 respectively. In a recent work Gao et al. obtained 80% C2=–C4= selectivity with ~ 4% CH4
1270 inhibition among all hydrocarbons using In-Zr composite based oxides and SAPO-34 zeolites
1271 (Gao et al., 2018; Dang et al., 2018). In addition, CO selectivity over the bifunctional catalysts
1272 could be jacked up to 85 % during CO2 reduction to lower olefins at high temperature reversed
1274 Moreover, the CO selectivity could be reduced if In2O3 is further modified by Zr and Zn
1275 promoters (Gao et al., 2017). The reduction of CO2 to hydrocarbons over bifunctional catalysts
1276 involves hydrogenation of CO2 to CHxO species or intermediates on the oxygenated sites of the
1277 oxide component as the first step, and the migration of the intermediates to the zeolite
1278 component where they are transformed into various hydrocarbons at the acid sites of the zeolite
1279 (Cheng et al., 2017; Gao et al., 2017). The transport of the CHxO intermediates plays a crucial
1280 role in suppressing the undesired RWGS side reaction and achieving a high selectivity and it can
1281 be improved by acid treatment of the zeolite crystals. The performance of bifunctional catalysts
1282 with and without acid treatment for In2O3-ZnZrOx/SAPO-34 composite is summarized in Table
1283 5. The maximum of C2=–C4= selectivity is 85 % among all hydrocarbons with 1.6 % CH4.
1284 Moreover, the decrease of zeolite crystal size (Dang et al., 2019) and shortening of the diffusion
1285 path of the methanol intermediates should favor the formation of lower olefins over the
76
1287 Table 5: Performance of In2O3-ZnZrOx and SAPO-34 as bifunctional catalysts prior to and after
y conversion
(%)
C4 =
ZnZrOx/SAPO-
34-S-a
ZnZrOx/SAPO-
34-S
ZnZrOx/SAPO-
34-S-H-a
ZnZrOx/SAPO-
34-H
ZnZrOx/SAPO-
77
34-C-a
ZnZrOx/SAPO-
34-C
1290 Information gathered from recent literature confirms that continuous efforts are being made to
1291 reduce CO2 in the environment. However, most articles are focused on its conversion to
1292 hydrocarbons, mainly methane, and other organic compounds, such as ethanol, methanol, and
1293 methanoic acid. Only limited articles were published on CO2 reduction to either FMD (mostly as
1294 intermediates) or hydrogen. Thus, more reaction pathways that yield these products should be
1295 investigated. There are some challenges involved in CO2 conversion into useful products, which
1296 should be considered. These include but are not limited to the production of harmful gases like
1297 CO, technological feasibilities in terms of more research as regards more-efficient conversion
1298 techniques, instability and low selectivity of the conversion products, and insufficient basic
1299 understanding of the catalytic processes (Sharma et al., 2022). Having pointed out these fall-outs,
1300 it is worth mentioning that scientists and other stakeholders should consider the project of CO2
1301 conversion as germane because it technically solves the problem of global warming and, in turn,
1303 The use of nanomaterials-based technologies for the transformation of CO2 is a relatively new
1304 area that scientists have to continue to explore. Unarguably, the demonstrated suitability,
1305 selectivity and applicability of nanomaterials for producing hydrocarbons and valued-added
1306 chemicals (methanol, ethanol, formic acid, formaldehyde, etc) have made nanomaterials
78
1307 promising systems for the transformation of CO2 into various fuels. The research findings so far
1308 are quite intriguing at the laboratory scale. However, the extension of these findings to industrial
1309 scale would be highly phenomenal. It is therefore recommended and/or suggested industrially
1310 suitable, and applicable catalyst (nanomaterials) be developed to selectively and efficiently
1312
1313
1314
1315
1316
1317 Funding
1319 Acknowledgements
1320 Alli, Y.A. would like to seize this opportunity to appreciate CNRS-Laboratory of Coordination
1321 Chemistry and Prof. Karine Philippot for providing him with laboratory space and enabling
1322 environment for postdoctoral research activities. Also, Oladoye, P.O. would like to appreciate
1323 Prof. Yong Cai for providing him with laboratory space and enabling environment for his
1325
1326 References
79
1327 Abanades, S., Haeussler, A., 2021. Two-step thermochemical cycles using fibrous ceria pellets
1328 for H2 production and CO2 reduction in packed-bed solar reactors. Sustainable Materials
1330 Acar, C., Dincer, I., 2022. Selection criteria and ranking for sustainable hydrogen production
1332 https://doi.org/10.1016/j.ijhydene.2022.07.137
1333 Afolabi, T. Adeniyi, Ejeromedoghene, O., Olorunlana, G.E., Afolabi, T. Adeleke, Alli, Y.A.,
1334 2022. A selective and efficient chemosensor for the rapid detection of arsenic ions in
1336 022-04665-1
1337 Afshar, A.A.N., 2014. Chemical profile: formic acid. Webpage Summ. http//chemplan.
1339 Ahmad, K., Upadhyayula, S., 2019. Conversion of the greenhouse gas CO 2 to methanol over
1342 Alaghmandfard, A., Ghandi, K., 2022. A Comprehensive Review of Graphitic Carbon Nitride
1345 Alhaddad, M., Shawky, A., 2021. Pt-decorated ZnMn2O4 nanorods for effective photocatalytic
1346 reduction of CO2 into methanol under visible light. Ceram. Int. 47, 9763–9770.
1347 https://doi.org/10.1016/j.ceramint.2020.12.116
1348 Ali, Syed Saim, Ali, Syed Saif, Tabassum, N., 2022. A review on CO2 hydrogenation to ethanol:
1349 Reaction mechanism and experimental studies. J. Environ. Chem. Eng. 10, 106962.
80
1350 Alli, Y.A., Adewuyi, S., Bada, B.S., Thomas, S., Anuar, H., 2021. Quaternary Trimethyl
1351 Chitosan Chloride Capped Bismuth Nanoparticles with Positive Surface Charges: Catalytic
1353 Ali, N., Bilal, M., Nazir, M.S., Khan, A., Ali, F., Iqbal, H.M.N., 2020. Thermochemical and
1355 fuel via waste to energy theme. Science of the Total Environment 712.
1356 https://doi.org/10.1016/j.scitotenv.2019.136482
1357 Alli, Y.A., Ejeromedoghene, O., Oladipo, A., Adewuyi, S., Amolegbe, S.A., Anuar, H., Thomas,
1359 Silver Nanoparticles with Dual Antibacterial and Antifungal Activities. ACS Appl. Bio
1361 Álvarez, A., Borges, M., Corral‐Pérez, J.J., Olcina, J.G., Hu, L., Cornu, D., Huang, R., Stoian,
1362 D., Urakawa, A., 2017. CO2 activation over catalytic surfaces. ChemPhysChem 18, 3135–
1363 3141.
1364 An, K., Zhang, S., Wang, J., Liu, Q., Zhang, Z., Liu, Y., 2021. A highly selective catalyst of
1365 Co/La4Ga2O9 for CO2 hydrogenation to ethanol. J. Energy Chem. 56, 486–495.
1366 Anadón, A., Martínez-Larrañaga, M.R., Martínez, M.A., 2006. Probiotics for animal nutrition in
1367 the European Union. Regulation and safety assessment. Regul. Toxicol. Pharmacol. 45, 91–
1368 95.
1369 Arai, T., Sato, S., Uemura, K., Morikawa, T., Kajino, T., Motohiro, T., 2010.
81
1373 Arumugam, M., Tahir, M., Praserthdam, P., 2022. Effect of nonmetals (B, O, P, and S) doped
1374 with porous g-C3N4 for improved electron transfer towards photocatalytic CO2 reduction
1376 https://doi.org/10.1016/j.chemosphere.2021.131765
1377 Asiri, A.M., Lichtfouse, E., 2020. Conversion of Carbon Dioxide Into Hydrocarbons Vol. 2
1379 Ateka, A., Rodriguez-Vega, P., Ereña, J., Aguayo, A.T., Bilbao, J., 2022. A review on the
1380 valorization of CO2. Focusing on the thermodynamics and catalyst design studies of the
1382 https://doi.org/10.1016/j.fuproc.2022.107310
1383 Atsbha, T.A., Yoon, T., Seongho, P., Lee, C.J., 2021. A review on the catalytic conversion of
1384 CO2using H2for synthesis of CO, methanol, and hydrocarbons. Journal of CO2 Utilization.
1385 https://doi.org/10.1016/j.jcou.2020.101413
1386 Ávila-Bolívar, B., García-Cruz, L., Montiel, V., Solla-Gullón, J., 2019. Electrochemical
1389 Azofra, L.M., MacFarlane, D.R., Sun, C., 2016. An intensified π-hole in beryllium-doped boron
1390 nitride meshes: Its determinant role in CO2 conversion into hydrocarbon fuels. Chem.
1392 Bahmanpour, A.M., Hoadley, A., Tanksale, A., 2015. Formaldehyde production via
1393 hydrogenation of carbon monoxide in the aqueous phase. Green Chemistry 17, 3500-3507.
1394 10.1039/C5GC00599J
1395 Bart, J.C.J., Sneeden, R.P.A., 1987. Copper-zinc oxide-alumina methanol catalysts revisited.
82
1396 Catal. Today 2, 1–124.
1397 Bashiri, N., Royaee, S.J., Sohrabi, M., 2018. The catalytic performance of different promoted
1398 iron catalysts on combined supports Al2O3 for carbon dioxide hydrogenation. Res. Chem.
1400 Benson, E.E., Kubiak, C.P., Sathrum, A.J., Smieja, J.M., 2009. Electrocatalytic and
1401 homogeneous approaches to conversion of CO2 to liquid fuels. Chem. Soc. Rev. 38, 89–99.
1402 https://doi.org/10.1039/b804323j
1403 Bertau, M., Offermanns, H., Plass, L., Schmidt, F., Wernicke, H.-J., 2014. Methanol: the basic
1405 Boden, T., Andres, R., Marland, G. Global, regional, and national fossil-fuel co2 emissions
1406 (1751-2014)(v. 2017). Environmental System Science Data Infrastructure for a Virtual
1408 Bousquet, P., Peylin, P., Ciais, P., Le Quéré, C., Friedlingstein, P., Tans, P.P., 2000. Regional
1409 changes in carbon dioxide fluxes of land and oceans since 1980. Science 290, 1342-1346.
1410 Cao, A., Liu, G., Wang, L., Liu, J., Yue, Y., Zhang, L., Liu, Y., 2016. Growing layered double
1411 hydroxides on CNTs and their catalytic performance for higher alcohol synthesis from
1413 Cassia, R., Nocioni, M., Correa-Aragunde, N., Lamattina, L., 2018. Climate change and the
1414 impact of greenhouse gasses: CO2 and NO, friends and foes of plant oxidative stress. Front.
1416 Chang, S., Na, W., Zhang, J., Lin, L., Gao, W., 2021. Effect of the Zn/Ce ratio in Cu/ZnO–CeO
1417 2 catalysts on CO 2 hydrogenation for methanol synthesis. New J. Chem. 45, 22814–22823.
1418 Chen, H., Goodarzi, F., Mu, Y., Chansai, S., Mielby, J.J., Mao, B., Sooknoi, T., Hardacre, C.,
83
1419 Kegnæs, S., Fan, X., 2020a. Effect of metal dispersion and support structure of Ni/silicalite-
1420 1 catalysts on non-thermal plasma (NTP) activated CO2 hydrogenation. Appl. Catal. B
1422 Chen, H., Mu, Y., Shao, Y., Chansai, S., Xiang, H., Jiao, Y., Hardacre, C., Fan, X., 2020b.
1423 Nonthermal plasma (NTP) activated metal–organic frameworks (MOFs) catalyst for
1425 Chen, H., Mu, Y., Shao, Y., Chansai, S., Xu, S., Stere, C.E., Xiang, H., Zhang, R., Jiao, Y.,
1426 Hardacre, C., Fan, X., 2019. Coupling non-thermal plasma with Ni catalysts supported on
1427 BETA zeolite for catalytic CO2 methanation. Catal. Sci. Technol. 9, 4135–4145.
1428 https://doi.org/10.1039/c9cy00590k
1429 Chen, H., Mu, Y., Xu, Shanshan, Xu, Shaojun, Hardacre, C., Fan, X., 2020c. Recent advances in
1430 non-thermal plasma (NTP) catalysis towards C1 chemistry. Chinese J. Chem. Eng. 28,
1432 Chen, X., Deng, D., Pan, X., Hu, Y., Bao, X., 2015. N-doped graphene as an electron donor of
1433 iron catalysts for CO hydrogenation to light olefins. Chemical Communications 51, 217–
1435 Chen, Y., Vise, A., Klein, W.E., Cetinbas, F.C., Myers, D.J., Smith, W.A., Smith, W.A., Smith,
1436 W.A., Deutsch, T.G., Neyerlin, K.C., 2020. A Robust, Scalable Platform for the
1439 https://doi.org/10.1021/acsenergylett.0c00860
1440 Chen, Z., Jiang, Q., An, H., Zhang, J., Hao, S., Li, X., Cai, L., Yu, W., You, K., Zhu, X., Li, C.,
1441 2022. Platinum Group Metal Catalyst (RuO x , PtO x , and IrO x )-Decorated Ceria-
84
1442 Zirconia Solid Solution as High Active Oxygen Carriers for Solar Thermochemical CO2
1444 Cheng, F., Small, A.A., Colosi, L.M., 2021. The levelized cost of negative CO2 emissions from
1445 thermochemical conversion of biomass coupled with carbon capture and storage. Energy
1447 Cheng, K., Zhou, W., Kang, J., He, S., Shi, S., Zhang, Q., Pan, Y., Wen, W., Wang, Y., 2017.
1448 Bifunctional Catalysts for One-Step Conversion of Syngas into Aromatics with Excellent
1450 Choi, Y., Liu, P., 2009. Mechanism of Ethanol Synthesis from Syngas on Rh(111). J. Am. Chem.
1452 Cooper, G., McCaherty, J., Huschitt, E., Schwarck, R., Wilson, C., 2021. Ethanol industry
1454 Crake, A., Christoforidis, K. C., Kafizas, A., Zafeiratos, S., Petit, C., 2017. CO2 capture and
1457 Crake, A., Christoforidis, K.C., Godin, R., Moss, B., Kafizas, A., Zafeiratos, S., Durrant, J.R.,
1458 Petit, C., 2019. Titanium dioxide/carbon nitride nanosheet nanocomposites for gas phase
1459 CO2 photoreduction under UV-visible irradiation. Appl Catal B 242, 369–378.
1460 https://doi.org/10.1016/j.apcatb.2018.10.023
1461 Dang, S., Li, S., Yang, C., Chen, X., Li, X., Zhong, L., Gao, P., Sun, Y., 2019. Selective
1464 https://doi.org/10.1002/cssc.201900958
85
1465 Dang, S., Yang, H., Gao, P., Wang, H., Li, X., Wei, W., Sun, Y., 2019. A review of research
1466 progress on heterogeneous catalysts for methanol synthesis from carbon dioxide
1468 Dey, S., Naidu, B.S., Govindaraj, A., Rao, C.N.R., 2015. Noteworthy performance of La1-
1471 https://doi.org/10.1039/c4cp04578e
1472 Dias, Y.R., Perez-Lopez, O.W., 2021. CO2 conversion to methane using Ni/SiO2 catalysts
1474 Díez-Ramírez, J., Valverde, J.L., Sánchez, P., Dorado, F., 2016. CO2 hydrogenation to methanol
1475 at atmospheric pressure: influence of the preparation method of Pd/ZnO catalysts. Catal.
1477 Dong, W., ZHONG, D., HAO, G., LI, J., Qiang, Z., 2021. ZnOHF nanorods for efficient
1478 electrocatalytic reduction of carbon dioxide to carbon monoxide. J. Fuel Chem. Technol. 49,
1479 1379–1388.
1480 Du, J., Chen, A., Hou, S., Guan, J., 2022. CNT modified by mesoporous carbon anchored by Ni
1482 https://doi.org/10.1002/cey2.223
1483 Ejeromedoghene, O., Oderinde, O., Okoye, C.O., Oladipo, A., Alli, Y.A., 2022. Microporous
1484 metal-organic frameworks based on deep eutectic solvents for adsorption of toxic gases and
1486 https://doi.org/10.1016/j.ceja.2022.100361
1487 Ekambaram, S., Iikubo, Y., Kudo, A., 2007. Combustion synthesis and photocatalytic properties
86
1488 of transition metal-incorporated ZnO. J. Alloys Compd. 433, 237–240.
1489 https://doi.org/10.1016/j.jallcom.2006.06.045
1490 Elavarasan, M., Yang, W., Velmurugan, S., Chen, J.N., Yang, T.C.K., Yokoi, T., 2022. Highly
1491 Efficient Photothermal Reduction of CO2 on Pd2 Cu Dispersed TiO2 Photocatalyst and
1493 https://doi.org/10.3390/nano12030332
1494 Fall, B., Gaye, C., Niang, M., Adekunle, Y., Abdou, A., Diagne, K., Modou, D., 2022. Removal
1495 of Toxic Chromium Ions in Aqueous Medium Using a New Sorbent Based on rGO @ CNT
1497 Fan, M.-S., Abdullah, A.Z., Bhatia, S., 2011. Hydrogen production from carbon dioxide
1500 https://doi.org/10.1016/j.ijhydene.2011.01.064
1501 Fan, W.K., Tahir, M., 2022. Recent developments in photothermal reactors with understanding
1502 on the role of light/heat for CO2 hydrogenation to fuels: A review. Chem. Eng. J. 427,
1504 Feely, R.A., Sabine, C.L., Lee, K., Berelson, W., Kleypas, J., Fabry, V.J., Millero, F.J., 2004.
1505 Impact of anthropogenic CO2 on the CaCO3 system in the oceans. Science (80-. ). 305,
1506 362–366.
1507 Feng, K., Wang, Y., Guo, M., Zhang, J., Li, Z., Deng, T., Zhang, Z., Yan, B., 2021. In-
1508 situ/operando techniques to identify active sites for thermochemical conversion of CO2 over
1510 https://doi.org/10.1016/j.jechem.2021.03.054
87
1511 Fiordaliso, E.M., Sharafutdinov, I., Carvalho, H.W.P., Grunwaldt, J.-D., Hansen, T.W.,
1512 Chorkendorff, I., Wagner, J.B., Damsgaard, C.D., 2015. Intermetallic GaPd2 nanoparticles
1513 on SiO2 for low-pressure CO2 hydrogenation to methanol: Catalytic performance and in
1515 Fisher, J. B., Eisenberg, R., 2002. Electrocatalytic reduction of carbon dioxide by using
1516 macrocycles of nickel and cobalt. J. Am. Chem. Soc. 102, 7361–7363.
1517 https://doi.org/10.1021/ja00544a035
1518 Friedlingstein, P., 2015. Carbon cycle feedbacks and future climate change. Philosophical
1519 Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences
1521 Fu, Z.C., Moore, J.T., Liang, F., Fu, W.F., 2019. Highly efficient photocatalytic reduction of
1522 CO2 to CO using cobalt oxide-coated spherical mesoporous silica particles as catalysts.
1524 Fukuhara, C., Hayakawa, K., Suzuki, Y., Kawasaki, W., Watanabe, R., 2017. A novel nickel-
1525 based structured catalyst for CO2 methanation: A honeycomb-type Ni/CeO2 catalyst to
1526 transform greenhouse gas into useful resources. Appl. Catal. A Gen. 532, 12–18.
1527 Fuqiang, W., Ziming, C., Jianyu, T., Jiaqi, Z., Yu, L., Linhua, L., 2017. Energy storage
1528 efficiency analyses of CO2 reforming of methane in metal foam solar thermochemical
1530 https://doi.org/10.1016/j.applthermaleng.2016.10.025
1531 Furler, P., Scheffe, J., Gorbar, M., Moes, L., Vogt, U., Steinfeld, A., 2012. Solar thermochemical
1532 CO2 splitting utilizing a reticulated porous ceria redox system. Energy and Fuels 26, 7051–
88
1534 Gaikwad, R., Reymond, H., Phongprueksathat, N., Von Rohr, P.R., Urakawa, A., 2020. From
1537 Gao, K., Liu, X., Jiang, Z., Zheng, H., Song, C., Wang, X., Tian, C., Dang, C., Sun, N., Xuan,
1538 Y., 2022. Direct solar thermochemical CO2 splitting based on Ca- and Al- doped SmMnO3
1539 perovskites: Ultrahigh CO yield within small temperature swing. Renew. Energy 194, 482–
1541 Gao, P., Li, S., Bu, X., Dang, S., Liu, Z., Wang, H., Zhong, L., Qiu, M., Yang, C., Cai, J., Wei,
1542 W., Sun, Y., 2017. Direct conversion of CO2 into liquid fuels with high selectivity over a
1544 Halder, P.K., Paul, N., Beg, M.R.A., 2014. Assessment of biomass energy resources and related
1546 https://doi.org/10.1016/j.rser.2014.07.071
1547 Guene Lougou, B., Shuai, Y., Zhang, H., Ahouannou, C., Zhao, J., Kounouhewa, B.B., Tan, H.,
1548 2020. Thermochemical CO2 reduction over NiFe2O4@alumina filled reactor heated by high-
1550 Haeussler, A., Abanades, S., Julbe, A., Jouannaux, J., Drobek, M., Ayral, A., Cartoixa, B., 2020.
1552 H2O and CO2 in a novel high–temperature solar reactor. Chemical Engineering Research
1554 Handoko, A.D., Tang, J., 2013. Controllable proton and CO2 photoreduction over Cu2O with
1556 https://doi.org/10.1016/j.ijhydene.2013.03.128
89
1557 He, J., Wang, X., Jin, S., Liu, Z.Q., Zhu, M., 2022. 2D metal-free heterostructure of covalent
1558 triazine framework/g-C3N4 for enhanced photocatalytic CO2 reduction with high selectivity.
1560 Heim, L.E., Konnerth, H., Prechtl, M.H.G., 2017. Future perspectives for formaldehyde:
1561 pathways for reductive synthesis and energy storage. Green Chemistry 19, 2347-2355.
1562 10.1039/C6GC03093A
1563 Herzer, D., 2022. The impact on domestic CO2 emissions of domestic government-funded clean
1564 energy R&D and of spillovers from foreign government-funded clean energy
1566 Hsu, H.C., Shown, I., Wei, H.Y., Chang, Y.C., Du, H.Y., Lin, Y.G., Tseng, C.A., Wang, C.H.,
1567 Chen, L.C., Lin, Y.C., Chen, K.H., 2013. Graphene oxide as a promising photocatalyst for
1569 Hu, B., Guild, C., Suib, S.L., 2013. Thermal, electrochemical, and photochemical conversion of
1571 Hu, F., Ye, R., Jin, C., Liu, D., Chen, X., Li, C., Lim, K.H., Song, G., Wang, T., Feng, G., 2022.
1572 Ni nanoparticles enclosed in highly mesoporous nanofibers with oxygen vacancies for
1574 Hu, Y., Wu, J., Han, Y., Xu, W., Zhang, L., Xia, X., Huang, C., Zhu, Y., Tian, M., Su, Y., Li, L.,
1575 Hou, B., Lin, J., Liu, W., Wang, X., 2021. Intensified solar thermochemical CO2 splitting
1578 Huang, Q., Yu, J., Cao, S., Cui, C., Cheng, B., 2015. Efficient photocatalytic reduction of CO2
90
1580 https://doi.org/10.1016/j.apsusc.2015.07.082
1581 Huang, Y., Yan, C.F., Guo, C.Q., Lu, Z.X., Shi, Y., Wang, Z. Da, 2017. Synthesis of GO-
1582 modified Cu2O nanosphere and the photocatalytic mechanism of water splitting for
1584 https://doi.org/10.1016/j.ijhydene.2016.10.157
1585 Inoue, T., Fujishima, A., Konishi, S., Honda, K., 1979. Photoelectrocatalytic reduction of carbon
1587 https://doi.org/10.1038/277637a0
1588 Inoue, Y., Izumida, H., Sasaki, Y., Hashimoto, H., 1976. Catalytic fixation of carbon dioxide to
1589 formic acid by transition-metal complexes under mild conditions. Chem. Lett. 5, 863–864.
1590 Jadhav, S.G., Vaidya, P.D., Bhanage, B.M., Joshi, J.B., 2014. Catalytic carbon dioxide
1591 hydrogenation to methanol: A review of recent studies. Chem. Eng. Res. Des. 92, 2557–
1592 2567.
1593 Jansen, D., Gazzani, M., Manzolini, G., Dijk, E. van, Carbo, M., 2015. Pre-combustion CO2
1595 https://doi.org/10.1016/j.ijggc.2015.05.028
1596 Jiang, B., Guene Lougou, B., Zhang, H., Geng, B., Wu, L., Shuai, Y., 2022. Preparation and
1599 Jiao, F., Li, Jinjing, Pan, X., Xiao, J., Li, H., Ma, H., Wei, M., Pan, Y., Zhou, Z., Li, M., Miao,
1600 S., Li, Jian, Zhu, Y., Xiao, D., He, T., Yang, J., Qi, F., Fu, Q., Bao, X., 2016. Selective
1601 conversion of syngas to light olefins Downloaded from. Catalysis 351, 1065–1067.
1602 Jiao, F., Pan, X., Gong, K., Chen, Y., Li, G., Bao, X., 2018. Shape‐Selective Zeolites Promote
91
1603 Ethylene Formation from Syngas via a Ketene Intermediate. Angewandte Chemie 130,
1604 4782–4786. https://doi.org/10.1002/ange.20180139Jiang, Q., Chen, Z., Tong, J., Yang, M.,
1605 Jiang, Z., Li, C., 2016. Catalytic Function of IrOx in the Two-Step Thermochemical CO2-
1607 https://doi.org/10.1021/acscatal.5b017747
1608 Kajaste, R., Hurme, M., Oinas, P., 2018. Methanol-Managing greenhouse gas emissions in the
1609 production chain by optimizing the resource base. AIMS Energy 6, 1074–1102.
1610 Kamkeng, A.D.N., Wang, M., Hu, J., Du, W., Qian, F., 2021. Transformation technologies for
1611 CO2 utilisation: Current status, challenges and future prospects. Chemical Engineering
1613 Keller, A.A., Adeleye, A.S., Conway, J.R., Garner, K.L., Zhao, L., Cherr, G.N., Hong, J.,
1614 Gardea-Torresdey, J.L., Godwin, H.A., Hanna, S., Ji, Z., Kaweeteerawat, C., Lin, S.,
1615 Lenihan, H.S., Miller, R.J., Nel, A.E., Peralta-Videa, J.R., Walker, S.L., Taylor, A.A.,
1616 Torres-Duarte, C., Zink, J.I., Zuverza-Mena, N., 2017. Comparative environmental fate and
1618 https://doi.org/10.1016/j.impact.2017.05.003
1619 Khalilzadeh, A., Shariati, A., 2019. Fe-N-TiO2/CPO-Cu-27 nanocomposite for superior CO2
1620 photoreduction performance under visible light irradiation. Sol. Energy 186, 166–174.
1621 https://doi.org/10.1016/j.solener.2019.05.009
1622 Kobl, K., Thomas, S., Zimmermann, Y., Parkhomenko, K., Roger, A.-C., 2016. Power-law
1623 kinetics of methanol synthesis from carbon dioxide and hydrogen on copper–zinc oxide
1624 catalysts with alumina or zirconia supports. Catalysis Today 270, 31-42.
1625 https://doi.org/10.1016/j.cattod.2015.11.020
92
1626 Kothandaraman, J., Goeppert, A., Czaun, M., Olah, G.A., Prakash, G.K.S., 2016. Conversion of
1627 CO2 from air into methanol using a polyamine and a homogeneous ruthenium catalyst. J.
1629 Kovačič, Ž., Likozar, B., Huš, M., 2020. Photocatalytic CO2 Reduction: A Review of Ab Initio
1630 Mechanism, Kinetics, and Multiscale Modeling Simulations. ACS Catal. 10, 14984–15007.
1631 https://doi.org/10.1021/acscatal.0c02557
1632 Koytsoumpa, E.I., Karellas, S., 2018. Equilibrium and kinetic aspects for catalytic methanation
1633 focusing on CO2 derived Substitute Natural Gas (SNG). Renew. Sustain. Energy Rev. 94,
1634 536–550.
1635 Kuang, Y., Shang, J., Zhu, T., 2020. Photoactivated Graphene Oxide to Enhance Photocatalytic
1637 https://doi.org/10.1021/acsami.9b18899
1638 Kubacka, A., Fernández-García, M., Colón, G., 2011. Advanced Nanoarchitectures for Solar
1640 https://doi.org/10.1021/cr100454n
1641 Kumar, A., Mohammed, A.A.A., Saad, M.A.H.S., Al-Marri, M.J., 2022. Effect of nickel on
1642 combustion synthesized copper/fumed-SiO2 catalyst for selective reduction of CO2 to CO.
1644 Kumaravel, V., Bartlett, J., Pillai, S.C., 2020. Photoelectrochemical conversion of carbon dioxide
1645 (CO2) into fuels and value-added products. ACS Energy Lett. 5, 486–519.
1646 Kumari, G., Zhang, X., Devasia, D., Heo, J., Jain, P.K., 2018. Watching visible light-driven CO2
1648 https://doi.org/10.1021/acsnano.8b03617
93
1649 L. Souza, M., H. B. Lima, F., 2021. Dibenzyldithiocarbamate-Functionalized Small Gold
1650 Nanoparticles as Selective Catalysts for the Electrochemical Reduction of CO2 to CO. ACS
1652 Lee, D.-K., Kim, D.-S., Kim, S.-W., 2001. Selective formation of formaldehyde from carbon
1653 dioxide and hydrogen over PtCu/SiO2. Applied Organometallic Chemistry 15, 148-150.
1654 https://doi.org/10.1002/1099-0739(200102)15:2<148::AID-AOC104>3.0.CO;2-N
1655 Lervold, S., Lødeng, R., Yang, J., Skjelstad, J., Bingen, K., Venvik, H.J., 2021. Partial oxidation
1658 Li, B., Duan, Y., Luebke, D., Morreale, B., 2013. Advances in CO2 capture technology: A patent
1660 Li, C.W., Ciston, J., Kanan, M.W., 2014. Electroreduction of carbon monoxide to liquid fuel on
1662 https://doi.org/10.1038/nature13249
1663 Li, J., Yu, T., Miao, D., Pan, X., Bao, X., 2019. Carbon dioxide hydrogenation to light olefins
1664 over ZnO-Y2O3 and SAPO-34 bifunctional catalysts. Catal Commun 129.
1665 https://doi.org/10.1016/j.catcom.2019.105711
1666 Li, P., Jia, X., Zhang, Jinping, Li, J., Zhang, Jinqiang, Wang, L., Wang, J., Zhou, Q., Wei, W.,
1667 Zhao, X., Wang, S., Sun, H., 2022. The roles of gold and silver nanoparticles on
1670 Li, Q., Jiang, L., Lin, Q., Zhou, G., Zhan, F., Zheng, Q., Wei, X., 2002. Properties of hydrogen
1671 storage alloy Mg~ 2Ni produced by hydriding combustion synthesis. CHINESE J.
94
1672 NONFERROUS Met. 12, 914–918.
1673 Li, W., Seredych, M., Rodríguez-Castellón, E., Bandosz, T.J., 2016. Metal-free Nanoporous
1676 Li, X.X., Liu, J., Zhang, L., Dong, L.Z., Xin, Z.F., Li, S.L., Huang-Fu, X.Q., Huang, K., Lan,
1679 https://doi.org/10.1021/acsami.9b03861
1680 Li, Y., Wang, W.N., Zhan, Z., Woo, M.H., Wu, C.Y., Biswas, P., 2010. Photocatalytic reduction
1681 of CO2 with H2O on mesoporous silica supported Cu/TiO2 catalysts. Appl. Catal. B
1683 Li, Y., Zeng, Z., Zhang, Y., Chen, Y., Wang, W., Xu, X., Du, M., Li, Z., Zou, Z., 2022.
1686 Li, Y.-T., Zhou, L., Cui, W.-G., Li, Z.-F., Li, W., Hu, T.-L., 2022. Iron promoted MOF-derived
1687 carbon encapsulated NiFe alloy nanoparticles core-shell catalyst for CO2 methanation. J.
1689 Li, Z., Feng, Y., Li, Y., Chen, X., Li, N., He, W., Liu, J., 2022. Fabrication of Bi/Sn bimetallic
1692 Li, Z., Liu, J., Shi, R., Waterhouse, G.I.N., Wen, X.D., Zhang, T., 2021. Fe-Based Catalysts for
1693 the Direct Photohydrogenation of CO2 to Value-Added Hydrocarbons. Adv. Energy Mater.
95
1695 Lingampalli, S.R., Ayyub, M.M., Rao, C.N.R., 2017. Recent Progress in the Photocatalytic
1697 https://doi.org/10.1021/acsomega.7b00721
1698 Liu, H., Gao, X., Shi, D., He, D., Meng, Q., Qi, P., Zhang, Q., 2022. Recent Progress on
1699 Photothermal Heterogeneous Catalysts for CO2 Conversion Reactions. Energy Technol. 10,
1701 Liu, K., Sakurai, M., Aono, M., 2010. ZnO-based ultraviolet photodetectors. Sensors 10, 8604–
1703 Liu, S., Tao, H., Zeng, L., Liu, Qi, Xu, Z., Liu, Qingxia, Luo, J.-L., 2017. Shape-dependent
1704 electrocatalytic reduction of CO2 to CO on triangular silver nanoplates. J. Am. Chem. Soc.
1706 Liu, W., Zhai, P., Li, A., Wei, B., Si, K., Wei, Y., Wang, X., Zhu, G., Chen, Q., Gu, X., Zhang,
1707 R., Zhou, W., Gong, Y., 2022. Electrochemical CO2 reduction to ethylene by ultrathin CuO
1709 Liu, X., Wang, M., Zhou, C., Zhou, W., Cheng, K., Kang, J., Zhang, Q., Deng, W., Wang, Y.,
1710 2017. Selective transformation of carbon dioxide into lower olefins with a bifunctional
1711 catalyst composed of ZnGa2O4 and SAPO-34. Chemical Communications 54, 140–143.
1712 https://doi.org/10.1039/c7cc08642c
1713 Liu, Y., Chen, C., Valdez, J., Motta Meira, D., He, W., Wang, Y., Harnagea, C., Lu, Q., Guner,
1714 T., Wang, H., Liu, C.H., Zhang, Q., Huang, S., Yurtsever, A., Chaker, M., Ma, D., 2021.
1717 Liu, Y., Chen, J.F., Bao, J., Zhang, Y., 2015. Manganese-modified Fe3O4 microsphere catalyst
96
1718 with effective active phase of forming light olefins from syngas. ACS Catal 5, 3905–3909.
1719 https://doi.org/10.1021/acscatal.5b00492
1720 Lorentzou, S., Karagiannakis, G., Pagkoura, C., Zygogianni, A., Konstandopoulos, A.G., 2014.
1721 Thermochemical CO2 and CO2/H2O splitting over NiFe2O4 for solar fuels synthesis.
1723 Lou, D., Xu, A.B., Fang, Y., Cai, M., Lv, K., Zhang, D., Wang, X., Huang, Y., Li, C., He, L.,
1724 2021. Cobalt-Sputtered Anodic Aluminum Oxide Membrane for Efficient Photothermal
1726 https://doi.org/10.1002/cnma.202100162
1727 Lou, Y., Zhu, W., Wang, L., Yao, T., Wang, S., Yang, Bo, Yang, Bing, Zhu, Y., Liu, X., 2021.
1728 CeO2 supported Pd dimers boosting CO2 hydrogenation to ethanol. Appl. Catal. B Environ.
1730 Loutzenhiser, P.G., Steinfeld, A., 2011. Solar syngas production from CO2 and H2O in a two-
1731 step thermochemical cycle via Zn/ZnO redox reactions: Thermodynamic cycle analysis. Int
1733 Low, J., Cheng, B., Yu, J., 2017. Surface modification and enhanced photocatalytic CO 2
1734 reduction performance of TiO 2 : a review. Appl Surf Sci 392, 658–686.
1735 https://doi.org/10.1016/j.apsusc.2016.09.093
1736 Lu, B., Quan, F., Sun, Z., Jia, F., Zhang, L., 2019. Photothermal reverse-water-gas-shift over
1737 Au/CeO2 with high yield and selectivity in CO2 conversion. Catal. Commun. 129, 105724.
1738 https://doi.org/10.1016/j.catcom.2019.105724
1739 Lu, H., Yang, X., Gao, G., Wang, J., Han, C., Liang, X., Li, C., Li, Y., Zhang, W., Chen, X.,
1740 2016. Metal (Fe, Co, Ce or La) doped nickel catalyst supported on ZrO2 modified
97
1741 mesoporous clays for CO and CO2 methanation. Fuel 183, 335–344.
1742 Lu, J., Yang, L., Xu, B., Wu, Q., Zhang, D., Yuan, S., Zhai, Y., Wang, X., Fan, Y., Hu, Z., 2014.
1743 Promotion effects of nitrogen doping into carbon nanotubes on supported iron fischer-
1745 https://doi.org/10.1021/cs400931z
1746 Lu, P., Gao, D., He, H., Wang, Q., Liu, Z., Dipazir, S., Yuan, M., Zu, W., Zhang, G., 2019.
1747 Facile synthesis of a bismuth nanostructure with enhanced selectivity for electrochemical
1749 https://doi.org/10.1039/c9nr01094g
1750 Lu, X., Leung, D.Y.C., Wang, H., Leung, M.K.H., Xuan, J., 2014. Electrochemical reduction of
1752 Luo, Z., Wang, X., Wen, H., Pei, A., 2022. Hydrogen production from offshore wind power in
1754 https://doi.org/10.1016/j.ijhydene.2022.03.162
1755 Ma, M., Djanashvili, K., Smith, W.A., 2016a. Controllable Hydrocarbon Formation from the
1756 Electrochemical Reduction of CO2 over Cu Nanowire Arrays. Angew. Chemie - Int. Ed. 55,
1758 Ma, M., Trześniewski, B.J., Xie, J., Smith, W.A., 2016b. Selective and efficient reduction of
1761 Madadi Avargani, V., Zendehboudi, S., Cata Saady, N.M., Dusseault, M.B., 2022. A
1762 comprehensive review on hydrogen production and utilization in North America: Prospects
98
1763 and challenges. Energy Conversion and Management 269, 115927.
1764 https://doi.org/10.1016/j.enconman.2022.115927
1765 Maeda, K., 2019. Metal-Complex/Semiconductor Hybrid Photocatalysts and Photoelectrodes for
1767 https://doi.org/10.1002/adma.201808205
1768 Mehta, P., Barboun, P., Go, D.B., Hicks, J.C., Schneider, W.F., 2019. Catalysis Enabled by
1769 Plasma Activation of Strong Chemical Bonds: A Review. ACS Energy Lett. 4, 1115–1133.
1770 https://doi.org/10.1021/acsenergylett.9b00263
1771 Miao, Z., Hu, P., Nie, C., Xie, H., Fu, W., Li, Q., 2019. ZrO2 nanoparticles anchored on
1772 nitrogen-doped carbon nanosheets as efficient catalyst for electrochemical CO2 reduction. J.
1774 Mo, Z., Zhu, X., Jiang, Z., Song, Y., Liu, D., Li, Hongping, Yang, X., She, Y., Lei, Y., Yuan, S.,
1775 Li, Huaming, Song, L., Yan, Q., Xu, H., 2019. Porous nitrogen-rich g-C3N4 nanotubes for
1776 efficient photocatalytic CO2 reduction. Appl. Catal. B Environ. 256, 117854.
1777 https://doi.org/10.1016/j.apcatb.2019.117854
1778 Mondal, M.K., Balsora, H.K., Varshney, P., 2012. Progress and trends in CO2 capture/separation
1780 Moret, S., Dyson, P.J., Laurenczy, G., 2014. Direct synthesis of formic acid from carbon dioxide
1782 Mou, S., Wu, T., Xie, J., Zhang, Y., Ji, L., Huang, H., Wang, T., Luo, Y., Xiong, X., Tang, B.,
1783 Sun, X., 2019. Boron Phosphide Nanoparticles: A Nonmetal Catalyst for High-Selectivity
1785 https://doi.org/10.1002/adma.201903499
99
1786 Mubarak, S., Dhamodharan, D., Byun, H.-S., Pattanayak, D.K., Arya, S.B., 2022. Efficient
1788 formed on the surface of nanoporous structured Ti foil. J. Ind. Eng. Chem.
1789 Mustafa, A., Lougou, B.G., Shuai, Y., Wang, Z., Tan, H., 2020. Current technology development
1790 for CO2 utilization into solar fuels and chemicals: A review. Journal of Energy Chemistry.
1791 https://doi.org/10.1016/j.jechem.2020.01.023
1792 Mutz, B., Carvalho, H.W.P., Mangold, S., Kleist, W., Grunwaldt, J.D., 2015. Methanation of
1793 CO2: Structural response of a Ni-based catalyst under fluctuating reaction conditions
1795 https://doi.org/10.1016/j.jcat.2015.04.006
1796 Nair, M.M., Abanades, S., 2018. Experimental screening of perovskite oxides as efficient redox
1797 materials for solar thermochemical CO 2 conversion. Sustain. Energy Fuels 2, 843–854.
1798 https://doi.org/10.1039/c7se00516d
1799 Nandal, N., Prajapati, P.K., Abraham, B.M., Jain, S.L., 2022. CO2 to ethanol: A selective
1801 oxide/copper oxide and a copper-based metal-organic framework. Electrochim. Acta 404,
1802 139612.
1803 Navarro-Jaén, S., Virginie, M., Bonin, J., Robert, M., Wojcieszak, R., Khodakov, A.Y., 2021.
1804 Highlights and challenges in the selective reduction of carbon dioxide to methanol. Nat.
1806 Nguyen, C.T.K., Quang Tran, N., Seo, S., Hwang, H., Oh, S., Yu, J., Lee, J., Anh Le, T., Hwang,
1807 J., Kim, M., Lee, H., 2020. Highly efficient nanostructured metal-decorated hybrid
1808 semiconductors for solar conversion of CO2 with almost complete CO selectivity. Mater.
100
1809 Today 35, 25–33. https://doi.org/10.1016/j.mattod.2019.11.005
1810 Nguyen, T.D., Van Tran, T., Singh, S., Phuong, P.T.T., Bach, L.G., Nanda, S., Vo, D.-V.N.
1811 Conversion of Carbon Dioxide into Formaldehyde. In: Inamuddin, Asiri AM, Lichtfouse E,
1812 editors. Conversion of Carbon Dioxide into Hydrocarbons Vol. 2 Technology. Springer
1814 Nielsen, D.U., Hu, X.M., Daasbjerg, K., Skrydstrup, T., 2018. Chemically and electrochemically
1817 Nieminen, H., Laari, A., Koiranen, T., 2019. CO2 hydrogenation to methanol by a liquid-phase
1819 Nizio, M., Benrabbah, R., Krzak, M., Debek, R., Motak, M., Cavadias, S., Gálvez, M.E., Da
1820 Costa, P., 2016. Low temperature hybrid plasma-catalytic methanation over Ni-Ce-Zr
1822 https://doi.org/10.1016/j.catcom.2016.04.023
1823 Numpilai, T., Witoon, T., Chanlek, N., Limphirat, W., Bonura, G., Chareonpanich, M.,
1825 different temperatures for CO2 hydrogenation to light olefins. Appl Catal A Gen 547, 219–
1827 Offermanns, H., Plass, L., Bertau, M., 2014. From Raw Materials to Methanol, Chemicals and
1828 Fuels. Methanol Basic Chem. Energy Feed. Futur. Springer Heidelberg/Berlin, Ger. 1–7.
1829 Ola, O., Maroto-Valer, M.M., 2015. Transition metal oxide based TiO2 nanoparticles for visible
1830 light induced CO2 photoreduction. Appl. Catal. A Gen. 502, 114–121.
1831 https://doi.org/10.1016/j.apcata.2015.06.007
101
1832 Oladoye, P.O., Adegboyega, S.A., Giwa, A.-R.A., 2021. Remediation potentials of composite
1834 review of recent literatures. Environ Nanotechnol Monit Manag 16, 100568.
1835 https://doi.org/10.1016/j.enmm.2021.100568
1836 Olah, G.A., 2013. Towards oil independence through renewable methanol chemistry. Angew.
1838 Omoregbe, O., Mustapha, A.N., Steinberger-Wilckens, R., El-Kharouf, A., Onyeaka, H., 2020.
1839 Carbon capture technologies for climate change mitigation: A bibliometric analysis of the
1841 https://doi.org/10.1016/j.egyr.2020.05.003
1842 Otgonbayar, Z., Youn Cho, K., Oh, W.-C., 2020. Novel Micro and Nanostructure of a
1845 Pang, J., Zheng, M., Zhang, T., 2019. Synthesis of ethanol and its catalytic conversion, in:
1847 Parastaev, A., Hoeben, W.F.L.M., van Heesch, B.E.J.M., Kosinov, N., Hensen, E.J.M., 2018.
1850 https://doi.org/10.1016/j.apcatb.2018.08.011
1851 Paris, C., Karelovic, A., Manrique, R., Le Bras, S., Devred, F., Vykoukal, V., Styskalik, A.,
1852 Eloy, P., Debecker, D.P., 2020. CO2 Hydrogenation to Methanol with Ga- and Zn-Doped
102
1855 Park, H.-a., Choi, J.H., Choi, K.M., Lee, D.K., Kang, J.K., 2012. Highly porous gallium oxide
1856 with a high CO2 affinity for the photocatalytic conversion of carbon dioxide into methane.
1858 Pawelec, B., Guil-López, R., Mota, N., Fierro, J.L.G., Navarro Yerga, R.M., 2021. Catalysts for
1859 the conversion of co2 to low molecular weight olefins—a review. Materials.
1860 https://doi.org/10.3390/ma14226952
1861 Peiseler, L., Cabrera Serrenho, A., 2022. How can current German and EU policies be improved
1862 to enhance the reduction of CO2 emissions of road transport? Revising policies on electric
1863 vehicles informed by stakeholder and technical assessments. Energy Policy 168, 113124.
1864 https://doi.org/10.1016/j.enpol.2022.113124
1865 Pérez-Fortes, M., Schöneberger, J.C., Boulamanti, A., Harrison, G., Tzimas, E., 2016. Formic
1866 acid synthesis using CO2 as raw material: Techno-economic and environmental evaluation
1868 Petrovic, B., Gorbounov, M., Masoudi Soltani, S., 2021. Influence of surface modification on
1869 selective CO2 adsorption: A technical review on mechanisms and methods. Microporous
1871 Porosoff, M.D., Yan, B., Chen, J.G., 2016. Catalytic reduction of CO 2 by H 2 for synthesis of
1872 CO, methanol and hydrocarbons: challenges and opportunities. Energy Environ. Sci. 9, 62–
1873 73.
1874 Prockop, L.D., Chichkova, R.I., 2007. Carbon monoxide intoxication: An updated review.
1876 https://doi.org/10.1016/j.jns.2007.06.037
103
1877 Purwanto, Deshpande, R.M., Chaudhari, R.V., Delmas, H., 1996. Solubility of Hydrogen,
1878 Carbon Monoxide, and 1-Octene in Various Solvents and Solvent Mixtures. Journal of
1880 Qi, Z., Biener, J., Biener, M., 2019. Surface oxide-derived nanoporous gold catalysts for
1882 Qi, Z., Biener, M.M., Kashi, A.R., Hunegnaw, S., Leung, A., Ma, S., Huo, Z., Kuhl, K.P.,
1883 Biener, J., 2021. Electrochemical CO2 to CO reduction at high current densities using a
1885 Qiao, J., Liu, Y., Hong, F., Zhang, J., 2014. A review of catalysts for the electroreduction of
1887 https://doi.org/10.1039/c3cs60323g
1888 Rahmani, S., Meshkani, F., Rezaei, M., 2019. Preparation of Ni-M (M: La, Co, Ce, and Fe)
1889 catalysts supported on mesoporous nanocrystalline γ-Al2O3 for CO2 methanation. Environ.
1891 Ramírez-Valencia, L.D., Bailón-García, E., Carrasco-Marín, F., Pérez-Cadenas, A.F., 2021.
1892 From CO2 to value-added products: a review about carbon-based materials for electro-
1894 Rauch, M., Strater, Z., Parkin, G., 2019. Selective Conversion of Carbon Dioxide to
1896 Derived from Carbon Dioxide into Organic Molecules. Journal of the American Chemical
1898 Ren, D., Tee Wong, N., Denny Handoko, A., Huang, Y., Siang Yeo, B., 2015. Mechanistic
1899 Insights into the Enhanced Activity and Stability of Agglomerated Cu Nanocrystals for the
104
1900 Electrochemical Reduction of Carbon Dioxide to n-Propanol. J. Phys. Chem. Lett. 7, 20–24.
1901 https://doi.org/10.1021/acs.jpclett.5b02554
1902 Rezaul Karim, K.M., Ong, H.R., Abdullah, H., Yousuf, A., Cheng, C.K., Rahman Khan, M.M.,
1904 under visible light irradiation. Int. J. Hydrogen Energy 43, 18185–18193.
1905 https://doi.org/10.1016/j.ijhydene.2018.07.174
1906 Rose, J.J., Wang, L., Xu, Q., McTiernan, C.F., Shiva, S., Tejero, J., Gladwin, M.T., 2017.
1909 Rosen, B.A., Hod, I., 2018. Tunable Molecular-Scale Materials for Catalyzing the Low-
1911 https://doi.org/10.1002/adma.201706238
1912 Roy, S., Cherevotan, A., Peter, S.C., 2018. Thermochemical CO2 Hydrogenation to Single
1913 Carbon Products: Scientific and Technological Challenges. ACS Energy Lett. 3, 1938–1966.
1914 https://doi.org/10.1021/acsenergylett.8b00740
1915 Sabatier, P., 1902. New synthesis of methane. Comptes Rendus 134, 514–516.
1916 Schimel, D., Stephens, B.B., Fisher, J.B., 2015. Effect of increasing CO2 on the terrestrial
1917 carbon cycle. Proceedings of the National Academy of Sciences 112, 436-441.
1918 10.1073/pnas.1407302112
1919 Scibioh, M.A., Viswanathan, B., 2004. Electrochemical reduction of carbon dioxide: a status
1920 report, in: Proc Indian Natn Sci Acad. pp. 1–56.
1921 Sha, F., Han, Z., Tang, S., Wang, J., Li, C., 2020. Hydrogenation of Carbon Dioxide to Methanol
105
1923 https://doi.org/10.1002/cssc.202002054
1924 Sharma, D., Sharma, R., Chand, D., Chaudhary, A., 2022. Nanocatalysts as potential candidates
1925 in transforming CO2 into valuable fuels and chemicals: A review. Environmental
1927 https://doi.org/10.1016/j.enmm.2022.100671
1928 Shen, H., Peppel, T., Strunk, J., Sun, Z., 2020. Photocatalytic Reduction of CO 2 by Metal-Free-
1930 https://doi.org/10.1002/solr.201900546
1931 Shen, Q., Chen, Z., Huang, X., Liu, M., Zhao, G., 2015. High-Yield and Selective
1934 https://doi.org/10.1021/acs.est.5b00066
1935 Shen, X., Li, C., Wu, Z., Tang, R., Shen, J., Chu, M., Xu, A., Zhang, B., He, L., Zhang, X., 2022.
1936 Rationally designed nanoarray catalysts for boosted photothermal CO2 hydrogenation.
1938 Shi, H., Tian, C., Liu, X., Sun, N., Song, C., Zheng, H., Gao, K., Wang, X., Jiang, Z., Xuan, Y.,
1939 Ding, Y., 2023. Ni-phyllosilicate nanotubes coated by CeO2 for ultra-efficiency of 36.9%
1942 Shi, L., Yang, L., Zhou, W., Liu, Y., Yin, L., Hai, X., Song, H., Ye, J., 2018. Photoassisted
1945 Shuai, Y., Zhang, H., Guene Lougou, B., Jiang, B., Mustafa, A., Wang, C.H., Wang, F., Zhao, J.,
106
1946 2021. Solar-driven thermochemical redox cycles of ZrO2 supported NiFe2O4 for CO2
1948 https://doi.org/10.1016/j.energy.2021.120073
1949 Solomon, S., Daniel, J.S., Sanford, T.J., Murphy, D.M., Plattner, G.K., Knutti, R., Friedlingstein,
1950 P., 2010. Persistence of climate changes due to a range of greenhouse gases. Proc. Natl.
1952 Solomon, S., Plattner, G.K., Knutti, R., Friedlingstein, P., 2009. Irreversible climate change due
1953 to carbon dioxide emissions. Proc. Natl. Acad. Sci. U. S. A. 106, 1704–1709.
1954 https://doi.org/10.1073/pnas.0812721106
1955 Song, C., Liu, X., Xu, M., Masi, D., Wang, Y., Deng, Y., Zhang, M., Qin, X., Feng, K., Yan, J.,
1956 Leng, J., Wang, Z., Xu, Y., Yan, B., Jin, S., Xu, D., Yin, Z., Xiao, D., Ma, D., 2020.
1957 Photothermal Conversion of CO2with Tunable Selectivity Using Fe-Based Catalysts: From
1959 Song, C., Wang, J., Wang, X., AlQahtani, M.S., Knecht, S.D., Bilen, S.G., Chu, W., 2022.
1960 Synergetic Effect of Nonthermal Plasma and Supported Cobalt Catalyst in Plasma-
1962 https://doi.org/10.2139/ssrn.4070252
1963 Stolar, T., Prašnikar, A., Martinez, V., Karadeniz, B., Bjelić, A., Mali, G., Friščić, T., Likozar,
1964 B., Užarević, K., 2021. Scalable mechanochemical amorphization of bimetallic Cu-Zn
1965 MOF-74 catalyst for selective CO2reduction reaction to methanol. ACS Appl. Mater.
1967 Studt, F., Behrens, M., Kunkes, E.L., Thomas, N., Zander, S., Tarasov, A., Schumann, J., Frei,
1968 E., Varley, J.B., Abild-Pedersen, F., Nørskov, J.K., Schlögl, R., 2015. The Mechanism of
107
1969 CO and CO2 Hydrogenation to Methanol over Cu-Based Catalysts. ChemCatChem 7,
1971 Subasi, N.T., 2020. Formaldehyde Advantages and Disadvantages: Usage Areas and Harmful
1973 10.5772/intechopen.89299
1974 Sultan, S., Lee, H., Park, S., Kim, M.M., Yoon, A., Choi, H., Kong, T.H., Koe, Y.J., Oh, H.S.,
1975 Lee, Z., Kim, H., Kim, W., Kwon, Y., 2022. Interface rich CuO/Al2CuO4 surface for
1976 selective ethylene production from electrochemical CO2 conversion. Energy Environ. Sci.
1977 https://doi.org/10.1039/d1ee03861c
1978 Sümbelli, Y., Biçen Ünlüer, Ö., Ersöz, A., Say, R., 2019. Synergistic effect of binanoenzyme and
1979 cryogel column on the production of formic acid from carbondioxide. Journal of Industrial
1981 Tada, S., Shimizu, T., Kameyama, H., Haneda, T., Kikuchi, R., 2012. Ni/CeO2 catalysts with
1982 high CO2 methanation activity and high CH4 selectivity at low temperatures. Int. J.
1984 Takalkar, G., Bhosale, R.R., 2019. Solar thermocatalytic conversion of CO2 using
1986 https://doi.org/10.1016/j.fuel.2019.115624
1987 Takalkar, G.D., Bhosale, R.R., 2019. Application of cobalt incorporated Iron oxide catalytic
1988 nanoparticles for thermochemical conversion of CO2. Appl. Surf. Sci. 495, 143508.
1989 https://doi.org/10.1016/j.apsusc.2019.07.250
1990 Takano, A., Tagawa, T., Goto, S., 1994. Carbon dioxide reforming of methane on supported
108
1992 Teng, L., Xuan, Y., Da, Y., Sun, C., Liu, X., Ding, Y., 2022. Direct solar-driven reduction of
1993 greenhouse gases into hydrocarbon fuels incorporating thermochemical energy storage via
1995 https://doi.org/10.1016/j.cej.2022.135955
1996 Usanmaz, S.E., Akarsu, E.S., Vural, N., 2002. Neurotoxic effects of acute and subacute
1997 formaldehyde exposures in mice. Environmental Toxicology and Pharmacology 11, 93-
1999 Visconti, C.G., Martinelli, M., Falbo, L., Infantes-Molina, A., Lietti, L., Forzatti, P., Iaquaniello,
2000 G., Palo, E., Picutti, B., Brignoli, F., 2017. CO2 hydrogenation to lower olefins on a high
2001 surface area K-promoted bulk Fe-catalyst. Appl Catal B 200, 530–542.
2002 https://doi.org/10.1016/j.apcatb.2016.07.047
2003 Wang, L., Ma, T., Dai, S., Ren, T., Chang, Z., Fu, M., Li, X., Li, Y., 2020. Solar thermochemical
2004 CO2 splitting with doped perovskite LaCo0.7Zr0.3O3: Thermodynamic performance and
2006 Wang, D., Bi, Q., Yin, G., Zhao, W., Huang, F., Xie, X., Jiang, M., 2016. Direct synthesis of
2007 ethanol via CO 2 hydrogenation using supported gold catalysts. Chem. Commun. 52,
2008 14226–14229.
2009 Wang, D., Huang, R., Liu, W., Sun, D., Li, Z., 2014. Fe-based MOFs for photocatalytic CO2
2010 reduction: Role of coordination unsaturated sites and dual excitation pathways. ACS Catal.
2012 Wang, F., Wei, M., Evans, D.G., Duan, X., 2016. CeO 2-based heterogeneous catalysts toward
2014 Wang, J., Yu, Y., Cui, J., Li, X., Zhang, Y., Wang, C., Yu, X., Ye, J., 2022. Defective g-
109
2015 C3N4/covalent organic framework van der Waals heterojunction toward highly efficient S-
2017 https://doi.org/10.1016/j.apcatb.2021.120814
2018 Wang, S., Lu, G.Q., Millar, G.J., 1996. Carbon Dioxide Reforming of Methane To Produce
2019 Synthesis Gas over Metal-Supported Catalysts: State of the Art. Energy & Fuels 10, 896-
2021 Wang, W., Ciais, P., Nemani, R.R., Canadell, J.G., Piao, S., Sitch, S., White, M.A., Hashimoto,
2022 H., Milesi, C., Myneni, R.B., 2013. Variations in atmospheric CO2 growth rates coupled
2023 with tropical temperature. Proceedings of the National Academy of Sciences 110, 13061-
2025 Wang, X., Maeda, K., Thomas, A., Takanabe, K., Xin, G., Carlsson, J.M., Domen, K.,
2026 Antonietti, M., 2009. A metal-free polymeric photocatalyst for hydrogen production from
2028 Wang, X.X., Duan, Y.H., Zhang, J.F., Tan, Y.S., 2022. Catalytic conversion of CO2 into high
2029 value-added hydrocarbons over tandem catalyst. Ranliao Huaxue Xuebao/Journal of Fuel
2031 Wang, Y., Liu, X., Han, X., Godin, R., Chen, J., Zhou, W., Jiang, C., Thompson, J.F., Mustafa,
2032 K.B., Shevlin, S.A., Durrant, J.R., Guo, Z., Tang, J., 2020. Unique hole-accepting carbon-
2033 dots promoting selective carbon dioxide reduction nearly 100% to methanol by pure water.
2035 Wang, Y., Wang, K., Zhang, B., Peng, X., Gao, X., Yang, G., Hu, H., Wu, M., Tsubaki, N.,
110
2038 Wang, Y., Xie, Y., Deng, M., Liu, T., Yang, H., 2021. Incorporation of Polyoxometalate in
2039 Sulfonic Acid-modified MIL-101-Cr for Enhanced CO2 Photoreduction Activity. Eur. J.
2041 Wang, Y., Zhang, X., Chang, K., Zhao, Z., Huang, J., Kuang, Q., 2022. MOF Encapsulated AuPt
2044 Wang, Z., Yuan, Q., Shan, J., Jiang, Z., Xu, P., Hu, Y., Zhou, J., Wu, L., Niu, Z., Sun, J., Cheng,
2045 T., A. Goddard III, W., 2020. Highly Selective Electrocatalytic Reduction of CO2 into
2047 https://doi.org/10.1021/acs.jpclett.0c01261
2048 Wang, Zhou jun, Song, H., Pang, H., Ning, Y., Dao, T.D., Wang, Zhuan, Chen, H., Weng, Y.,
2049 Fu, Q., Nagao, T., Fang, Y., Ye, J., 2019. Photo-assisted methanol synthesis via CO2
2050 reduction under ambient pressure over plasmonic Cu/ZnO catalysts. Appl. Catal. B Environ.
2052 Wei, Y.P., Liu, Y., Guo, F., Dao, X.Y., Sun, W.Y., 2019. Different functional group modified
2053 zirconium frameworks for the photocatalytic reduction of carbon dioxide. Dalt. Trans. 48,
2055 Weilhard, A., Argent, S.P., Sans, V., 2021. Efficient carbon dioxide hydrogenation to formic
2056 acid with buffering ionic liquids. Nat. Commun. 12, 1–7.
2057 Wen, X., Chang, L., Gao, Y., Han, J., Bai, Z., Huan, Y., Li, M., Tang, Z., Yan, X., 2018. A
2058 reassembled nanoporous gold leaf electrocatalyst for efficient CO2 reduction towards CO.
2060 Woyessa, G.W., dela Cruz, J. ar B., Rameez, M., Hung, C.H., 2021. Nanocomposite catalyst of
111
2061 graphitic carbon nitride and Cu/Fe mixed metal oxide for electrochemical CO2 reduction to
2063 Wu, D., Deng, K., Hu, B., Lu, Q., Liu, G., Hong, X., 2019. Plasmon-Assisted Photothermal
2066 Wu, S., Zhou, C., Doroodchi, E., Nellore, R., Moghtaderi, B., 2018. A review on high-
2067 temperature thermochemical energy storage based on metal oxides redox cycle. Energy
2069 https://doi.org/10.1016/j.enconman.2018.05.017
2070 Wu, J., Huang, Y., Ye, W., Li, Y., 2017. CO2 Reduction: From the Electrochemical to
2072 Wu, P.E., Juurlink, D.N., 2014. Carbon monoxide poisoning. Cmaj 186, 611.
2073 10.1503/cmaj.130972
2074 Wu, Z., Hao, X., Zhou, W., Yao, S., Han, J., Tang, X., Zhang, X., 2018. N-pentane activation
2077 Wu, Z., Li, C., Li, Z., Feng, K., Cai, M., Zhang, D., Wang, S., Chu, M., Zhang, C., Shen, J.,
2078 Huang, Z., Xiao, Y., Ozin, G.A., Zhang, X., He, L., 2021. Niobium and Titanium Carbides
2079 (MXenes) as Superior Photothermal Supports for CO2Photocatalysis. ACS Nano 15, 5696–
2081 Xie, W., Li, R., Xu, Q., 2018. Enhanced photocatalytic activity of Se-doped TiO2 under visible
2083
112
2084 Xu, C., Zhang, Y., Pan, F., Huang, W., Deng, B., Liu, J., Wang, Z., Ni, M., Cen, K., 2017.
2085 Guiding effective nanostructure design for photo-thermochemical CO2 conversion: From
2087 https://doi.org/10.1016/j.nanoen.2017.09.023
2088 Xu, H., Ouyang, S., Li, P., Kako, T., Ye, J., 2013. High-Active Anatase TiO2 Nanosheets
2089 Exposed with 95% {100} Facets Toward Efficient H2 Evolution and CO2 Photoreduction.
2091 Xu, L., Jiang, Q., Xiao, Z., Li, X., Huo, J., Wang, S., Dai, L., 2016. Plasma-Engraved Co3O4
2092 Nanosheets with Oxygen Vacancies and High Surface Area for the Oxygen Evolution
2094 Xu, S., Chansai, S., Stere, C., Inceesungvorn, B., Goguet, A., Wangkawong, K., Taylor, S.F.R.,
2095 Al-Janabi, N., Hardacre, C., Martin, P.A., Fan, X., 2019. Sustaining metal–organic
2096 frameworks for water–gas shift catalysis by non-thermal plasma. Nat. Catal. 2, 142–148.
2097 https://doi.org/10.1038/s41929-018-0206-2
2098 Xu, Shanshan, Chansai, S., Shao, Y., Xu, Shaojun, Wang, Y. chi, Haigh, S., Mu, Y., Jiao, Y.,
2099 Stere, C.E., Chen, H., Fan, X., Hardacre, C., 2020. Mechanistic study of non-thermal plasma
2100 assisted CO2 hydrogenation over Ru supported on MgAl layered double hydroxide. Appl.
2102 Yadav, R.K., Baeg, J.-O., Oh, G.H., Park, N.-J., Kong, K., Kim, J., Hwang, D.W., Biswas, S.K.,
2103 2012. A photocatalyst–enzyme coupled artificial photosynthesis system for solar energy in
2104 production of formic acid from CO2. J. Am. Chem. Soc. 134, 11455–11461.
2105 Yang, J., Pan, X., Jiao, F., Li, J., Bao, X., 2017. Direct conversion of syngas to aromatics.
113
2107 Yang, P., Guo, S., Yu, X., Zhang, F., Yu, B., Zhang, H., Zhao, Y., Liu, Z., 2019. Photocatalytic
2110 https://doi.org/10.1021/acs.iecr.9b00242
2111 Yang, X., Lee, J.H., Kattel, S., Xu, B., Chen, J.G., 2022. Tuning Reaction Pathways of
2114 Yang, X., Li, K., Wang, G., Li, X., Zhou, P., Ding, S., Lyu, Z., Chang, Y., Zhou, Y., Zhu, W.,
2115 2022. 2D Catalysts for CO 2 Photoreduction: Discussing Structure Efficiency Strategies and
2116 Prospects for Scaled Production Based on Current Progress. Chem. – A Eur. J.
2117 https://doi.org/10.1002/chem.202201881
2118 Yin, Z., Palmore, G.T.R., Sun, S., 2019. Electrochemical Reduction of CO2 Catalyzed by Metal
2120 Yu, S., Wilson, A.J., Heo, J., Jain, P.K., 2018. Plasmonic Control of Multi-Electron Transfer and
2121 C-C Coupling in Visible-Light-Driven CO2 Reduction on Au Nanoparticles. Nano Lett. 18,
2123 Yuan, J., Zhi, W.Y., Liu, L., Yang, M.P., Wang, H., Lu, J.X., 2018. Electrochemical reduction of
2124 CO2 at metal-free N-functionalized graphene oxide electrodes. Electrochim. Acta 282, 694–
2126 Zeng, J., Bejtka, K., Ju, W., Castellino, M., Chiodoni, A., Sacco, A., Farkhondehfal, M.A.,
2127 Hernández, S., Rentsch, D., Battaglia, C., 2018. Advanced Cu-Sn foam for selectively
2128 converting CO2 to CO in aqueous solution. Appl. Catal. B Environ. 236, 475–482.
2129 Zeng, Y., Tu, X., 2017. Plasma-catalytic hydrogenation of CO2 for the cogeneration of CO and
114
2130 CH4 in a dielectric barrier discharge reactor: Effect of argon addition. J. Phys. D. Appl.
2132 Zeng, Z., Huang, H., Fu, Z., Lai, H., Long, B., Ali, A., Song, T., Deng, G.J., 2021. Plasmonic Cu
2133 NPs-embedded phenothiazine benzene with tunable bonding units for superior
2135 https://doi.org/10.1016/j.apsusc.2021.149361
2136 Zhan, H., Li, F., Gao, P., Zhao, N., Xiao, F., Wei, W., Sun, Y., 2014. Influence of element
2137 doping on La–Mn–Cu–O based perovskite precursors for methanol synthesis from CO 2/H
2139 Zhang, C., Yang, S., Wu, J., Liu, M., Yazdi, S., Ren, M., Sha, J., Zhong, J., Nie, K., Jalilov, A.S.,
2140 Li, Z., Li, H., Yakobson, B.I., Wu, Q., Ringe, E., Xu, H., Ajayan, P.M., Tour, J.M., 2018.
2143 Zhang, D., Lv, K., Li, C., Fang, Y., Wang, S., Chen, Z., Wu, Z., Guan, W., Lou, D., Sun, W.,
2144 Yang, D., He, L., Zhang, X., 2021. All-Earth-Abundant Photothermal Silicon Platform for
2145 CO2 Catalysis with Nearly 100% Sunlight Harvesting Ability. Sol. RRL 5, 1–7.
2146 https://doi.org/10.1002/solr.202000387
2147 Zhang, F., Zhou, W., Xiong, X., Wang, Yuhao, Cheng, K., Kang, J., Zhang, Q., Wang, Ye, 2021.
2150 Zhang, J., Lu, S., Su, X., Fan, S., Ma, Q., Zhao, T., 2015. Selective formation of light olefins
2151 from CO2 hydrogenation over Fe-Zn-K catalysts. Journal of CO2 Utilization 12, 95–100.
2152 https://doi.org/10.1016/j.jcou.2015.05.004
115
2153 Zhang, X., Zhao, X., Chen, K., Fan, Y., Wei, S., Zhang, W., Han, D., Niu, L., 2021. Palladium-
2154 modified cuprous(i) oxide with {100} facets for photocatalytic CO2reduction. Nanoscale
2156 Zhang, Z., Li, J., Ji, M., Liu, Y., Wang, N., Zhang, X., Zhang, S., Ji, X., 2021. Encapsulation of
2159 Zhao, J., Bai, Y., Liang, X., Wang, T., Wang, C., 2021. Photothermal catalytic
2162 Zheng, J., Kang, A.N., WANG, J., Jing, L.I., Yuan, L.I.U., 2019. Direct synthesis of ethanol via
2163 CO2 hydrogenation over the Co/La-Ga-O composite oxide catalyst. J. Fuel Chem. Technol.
2165 Zhong, L., Yu, F., An, Y., Zhao, Y., Sun, Y., Li, Z., Lin, T., Lin, Y., Qi, X., Dai, Y., Gu, L., Hu,
2166 J., Jin, S., Shen, Q., Wang, H., 2016. Cobalt carbide nanoprisms for direct production of
2168 Zhou, Xiangping, Ji, J., Wang, D., Duan, X., Qian, G., Chen, D., Zhou, Xinggui, 2015.
2169 Hierarchical structured α-Al2O3 supported S-promoted Fe catalysts for direct conversion of
2171 https://doi.org/10.1039/c5cc00786k
2172 Zhou, Y., Zhao, Y., Shi, X., Tang, Y., Yang, Z., Pu, M., Lei, M., 2022. A theoretical study on the
2175 Zhu, W., Zhang, Y.J., Zhang, H., Lv, H., Li, Q., Michalsky, R., Peterson, A.A., Sun, S., 2014.
116
2176 Active and selective conversion of CO2 to CO on ultrathin Au nanowires. J. Am. Chem.
2178 Zhu, W., Zhao, K., Liu, S., Liu, M., Peng, F., An, P., Qin, B., Zhou, H., Li, H., He, Z., 2019.
2181 https://doi.org/10.1016/j.jechem.2019.03.030
117