2023 Cej Co2 Red Porp Rev

Download as pdf or txt
Download as pdf or txt
You are on page 1of 32

Chemical Engineering Journal 470 (2023) 144249

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Review

A review of the development of porphyrin-based catalysts for


electrochemical CO2 reduction
Shengshen Gu a, Aleksei N. Marianov b, Tiandan Lu a, Jing Zhong a, *
a
Jiangsu Key Laboratory of Advanced Catalytic Materials and Technology, School of Petrochemical Engineering, Changzhou University, Changzhou 213164, China
b
School of Chemical and Biomolecular Engineering, The University of Sydney, Sydney, NSW 2006, Australia

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

Keywords: The reduction of greenhouse gas carbon dioxide (CO2) to carbon-based fuels provides a potential solution to the
Porphyrins sustainable and scalable energy storage challenges. To solve the issues with activity, selectivity and stability,
Electrochemistry much attention has been put on porphyrin catalysts, and more recently, the integration of porphyrins on
CO2 reduction
conductive supports that operate in heterogeneous conditions. This review summarizes the main principles and
Homogeneous catalysis
strategies explored for the application of porphyrin catalysts to electrochemical CO2 reduction, ranging from
Heterogeneous catalysis
homogeneous catalysis and heterogeneous catalysis to structure modification and immobilization techniques.
The first section discusses mechanistic study of porphyrin catalysts including the catalytic steps involved in CO2
reduction, the effects of porphyrin ligands on catalytic activities. The second section provides insight into the
kinetic study of porphyrin catalysts. The third section presents examples of porphyrins in homogeneous catalysis
with particular focus on iron porphyrins and the boosting effects brought by Lewis acid and Brønsted acid. The
next section summarizes the main techniques for the heterogenization of porphyrin catalysts on conductive
supports, including non-covalent, covalent and periodic immobilization, whereas periodic immobilization
comprises porphyrin scaffolds or frameworks in the structure. The last section gives an update of porphyrins
employed in a flow cell. This review surveys the recent advances and basic principles of porphyrins in CO2
reduction, and the findings can be instrumental for designing efficient and selective catalysts for CO2 reduction.

1. Introduction to convert CO2 to various products such as methanol, ethanol, methane,


ethane, etc [4–6]. A major drawback obstructing the wider applications
Carbon dioxide electrochemical reduction reaction (CO2ERR) pow­ of such catalysts is the lack of selectivity. [7] Although there are reports
ered by renewable energy is an attractive sustainable approach to demonstrating Ag is capable of selectively reducing CO2 to CO at the
convert CO2 to value-added fuels or chemicals [1]. The development of overpotential of less than 500 mV [8,9], the cost of Ag presents a major
CO2ERR is pivotal to the sustainability of the world as it provides a concern for large scale applications.
possible solution to the issue of excessive CO2 emissions. Owing to the In contrast, molecular catalysts are superior to solid state catalysts in
abundance of CO2 in the atmosphere and availability of electricity, terms of achieving a high selectivity at a relatively low cost. The com­
CO2ERR is deemed as an economically viable process, and hence has mon molecular catalysts include phthalocyanines, quaterpyridines,
attracted extensive attention. However, there are factors limiting the bipyridines and etc. Cobalt phthalocyanine was firstly used for CO2ERR
broader application of the technique, such as low efficiency, selectivity in 1974 after being heterogenized on graphite [10]. Then extensive
and stability. Previous studies reveal that the key point in overcoming research has been carried out and the results suggest that phthalocya­
the limitations lies in the design of electrocatalysts [2], which could nines with cobalt, nickel and iron centers are active for CO2ERR, among
alleviate the kinetic energy barriers and mediate Proton Coupled Elec­ which cobalt exhibiting the best selectivity for CO [11,12]. However,
tron Transfers (PCETs) associated with CO2ERR [3]. Significant efforts synthesis of different phthalocyanine structures is troublesome due to
were put in the synthesis and evaluation of catalysts, among which solid their relatively low solubility in solvents. The first example of qua­
state catalysts and molecular catalysts are widely investigated. terpyridine applied in CO2ERR was reported by Che in 1995 where a
Solid state catalysts such as Ag, Cu, Ni, Pd, etc. have been employed selectivity of 80% for CO was achieved on cobalt quaterpyridine [13].

* Corresponding author.
E-mail address: [email protected] (J. Zhong).

https://doi.org/10.1016/j.cej.2023.144249
Received 6 February 2023; Received in revised form 30 May 2023; Accepted 18 June 2023
Available online 21 June 2023
1385-8947/© 2023 Elsevier B.V. All rights reserved.
S. Gu et al. Chemical Engineering Journal 470 (2023) 144249

While introduction of nickel and iron into quaterpyridines does not The above observation was further evaluated by computer simula­
provide a satisfactory CO selectivity [13,14]. The relatively simple tion, which validates the role of metal centers of porphyrins in CO2ERR.
structure of quaterpyridine restricts the options of modifications that In the initial step of CO2ERR, CO2 is activated by the nucleophilic attack
could be made on quaterpyridines, limiting the room for further of an electron-rich metal centre on the electrophilic C atom. The
improvements. reduction of CO2 requires populating its C-O σ* (LUMO) and C-O π*
Porphyrins, a major group of molecular catalysts, are extensively (LUMO + 1) orbitals with electrons from the metal center [25]. To
exploited owing to their high selectivities and activities. Compared with satisfy the requirement, an M+1,0 center with a d8 configuration in a
the above molecular catalysts, porphyrins have good solubilities in most square-pyramidal ligand field is the best for binding CO2 via its filled dz2
solvents and versatile structures, which underlines the importance of a (s) and dxz/yz (p back-bonding) orbitals. As such, porphyrins with Fe, Co
thorough discussion of porphyrins. The merits of using porphyrins are as and Ni are considered as the best catalysts for CO2ERR due to their d8
follows. First, the structures of porphyrins could be modified with li­ electron configuration. The product selectivity is largely dependent on
gands via organic approaches, which endows porphyrins with ultra- the ability of CO to remain adsorbed on the metal center [26]. The firm
performance towards CO2ERR [15,16]. Second, porphyrins could be adsorption of CO on metal could lead to further reduction to hydro­
applied to homogeneous electrocatalysis via direct dissolution in sol­ carbons while weak adsorption would release CO as the final product.
vents, and additionally, heterogeneous electrocatalysis via immobiliza­ Research indicated that Fe, Co, and Ni possessed doubly occupied dz2
tion on conductive supports [17]. Third, the reduction products could be orbitals, which would repel the lone pair of electrons on CO and release
tuned via altering the metal centers of porphyrins. Fourth, the active CO as the main product [27]. Meanwhile, further reduction can take
sites of porphyrins could be easily located during CO2ERR and thus be place on Cu center, affording CH4 as the product. In addition, for In and
analyzed via in-situ spectroscopy for mechanism study. Sn with outermost σ or p electrons, electron transfer takes place at the
Although there exist several overarching reviews on porphyrin cat­ more localized, lower energy orbital, affording [CO2]* as the interme­
alysts [17,18], a coverage of more recent developments in the field is diate, which is further protonated to HCOOH in the presence of a proton
required as new insights and developments have been reported. Also, a source [28].
thorough discussion of the principles to functionalize the porphyrins,
optimize the hybrid structures, and enhance the electron delivery is 2.2. Mechanisms in homogeneous conditions
required for the development of new types of catalysts. Additionally, the
strategies of enhancing porphyrin activities could be well extended to The process of CO2ERR on porphyrins could be performed in ho­
other types of molecular catalysts, which again necessitates a review of mogeneous and heterogeneous conditions, where the former requires
porphyrins. dissolution of porphyrins in solvents and the latter requires the immo­
Owing to the merits mentioned above, it is essential to have a review bilization of porphyrins on supports. Comprehensive studies comparing
focusing on porphyrins. Most of the published reviews embody a broad identical porphyrins in homogeneous and heterogeneous environments
range of molecular catalysts, which restricts the discussion and attention invariably demonstrate an overall enhancement in the catalytic perfor­
on porphyrins [19–22]. Also, some of the state-of-the-art development mance upon immobilization onto conductive supports [29]. The obser­
such as flow cell setup, in-situ spectroscopies, ultrathin nanosheets and vation indicates that the mechanisms involved in homogeneous and
their applications in porphyrins were not mentioned in detail. The heterogeneous CO2ERR might be different and thus need further
interpretation of previous work, gaps between research and industry, investigation. In this regard, cobalt porphyrins were chosen in an
and future directions were not fully discussed. This review has included attempt to probe into the above issue [23,30]. In homogeneous condi­
the above aspects and thus distinguished itself from previous reviews. tions, CO2ERR was observed to take place after the cobalt center was
Herein, we strive to present an up-to-date account of the porphyrin reduced to Co0 in a solvent of acetonitrile as confirmed by cyclic vol­
catalysts employed in CO2ERR with the focus on the following aspects. tammetry (CV) [31]. Thus Co0 is identified as the active metal site for
First, the mechanism involved in porphyrin catalysts is discussed. Sec­ CO2ERR. The subsequent reduction reaction follows the pathway as
ond, the benchmark for the evaluation of the catalytic performance of discussed above, starting with CO2 binding to the active Co0 sites and
porphyrins is discussed. Third, the cases of porphyrins operating in ending up with the release of CO (Fig. 1a) [32]. The formation of Co0
homogeneous catalysis are summarized. Fourth, the immobilization proceeds via the sequential injection of a pair of electrons into the
strategies for heterogeneous catalysis using porphyrins are presented macrocyclic cobalt porphyrins, which would require a negative poten­
and the corresponding activities are compared. Fifth, the updated tial of − 1.5 V vs. NHE (overpotential η = 970 mV) and thus consume
development in porphyrin catalysts is presented. excessive energy. Concerning the energy input, homogeneous CO2ERR is
deemed as a less efficient process.
2. Mechanistic pathways of CO2ERR on porphyrins z2In homogeneous conditions, the performance of porphyrins can be
enhanced via introduction of proton-donating ligands to the peripheral
2.1. The role of metal centers structures of porphyrins [33,34]. Introduction of proton-donating hy­
droxyls to iron porphyrins was pioneered by Savéant and an enhancement
The reaction pathway of CO2ERR on the metal porphyrins proceeds in the CO2ERR was reported [35]. On the other hand, through-structure
with the adsorption of CO2 on the active metal sites, followed by elec­ electronic effects such as electron donating and withdrawing were also
tron and proton transfer, accompanied with bond breaking and forma­ manipulated in order to improve the activities of metal porphyrins. The
tion, and ends up with the release of products [23]. An analysis of the introduction of electron donating methoxyl substituents would increase
steps would lead to the finding that the metal centers of porphyrins play the density of electrons on the metal center and thus facilitate the rate-
an important role in the formation of products. The effects of metal limiting binding of CO2 molecules [4]. In turn, the CO2ERR process
centers on CO2ERR have been thoroughly studied with seven different would be accelerated owing to the electron donating effects. Meanwhile
metal porphyrins [24]. Fe, Co, and Ni as the metal centers will the introduction of electron-withdrawing fluorine to porphyrins has been
contribute to the generation of CO as the major product under atmo­ proved to lower the overpotential by decreasing the electron density at
spheric CO2. In particular, Ni could catalyze HCOOH production at a the metal center, thus accelerating the transfer of electrons from the
comparable rate as CO production under enhanced pressure of 20 atm support to the active center [36]. The decreased electron density, how­
CO2. Cu is uniquely identified as the only metal that could produce not ever, would undermine the ability of the metal center to bind to CO2 and
only CO, but also hydrocarbons as the products. While Mn and Mg show thus bring detriment to the activity. As a result, a delicate design of the
barely any activities for CO2ERR, with all the current efficiencies go to porphyrin structures based on the analysis of electronic effects brought by
hydrogen evolution reaction (HER) [24]. ligands is needed in achieving an ideal catalyst [4].

2
S. Gu et al. Chemical Engineering Journal 470 (2023) 144249

Fig. 1. Proposed mechanisms of CO2ERR catalyzed by porphyrins in (a) homogeneous [32] and (b) heterogeneous systems. [37] Adapted with permission from
Ref. [32],Copyright 2020, American Chemical Society and Ref. [37], Copyright 2021, American Chemical Society.

2.3. Mechanisms in heterogeneous conditions mechanism in heterogeneous conditions [41]. It could reveal the
valence states or structures of porphyrins and thus clarify the potential
In contrast to homogeneous conditions, the potential required for mechanisms involved. Here the techniques employed in porphyrin cat­
heterogeneous CO2ERR is greatly reduced upon immobilization of co­ alysts are summarized.
balt porphyrins onto conductive supports [29]. It is observed that CoI In-situ XAS measurements were employed to monitor changes in the
rather than Co0 acts as the active site for CO2ERR in heterogeneous oxidation state of the Zn and Co centers in porphyrins [42,43]. The Zn K-
conditions, which explains the reduced potential as well as energy edge X-ray absorption near edge structure (XANES) spectra underwent
consumption. The formation of CoI only requires one electron transfer at no visible changes as the potential varied (Fig. 2a), indicating that the
the potential of − 0.53 V vs. NHE, rendering the process more energy oxidation state of Zn in the Zn porphyrin remained constant [42]. The
efficient. CoI as the active site would then bind to CO2 and take part in oxidation state of Zn center was determined to be + 2 by a comparison to
the following catalytic steps. (Fig. 1b) [37]. Such a mechanism is ZnO and Zn samples, which was further validated by the derivatives of
remarkable as it circumvents the costly step of Co0 formation. Hence the XANES spectra (Fig. 2b). These results suggested that Zn center was
heterogeneous CO2ERR is attracting more and more attention nowadays redox innocent as its oxidation state remained + 2 while porphyrin
with the potential to be industrialized. ligand might act as a redox mediator. Another example of in-situ XAS
In heterogeneous conditions, the activity of metal porphyrins can be was conducted on tetra(4-pyridyl) porphyrin cobalt(II) (STPyP-Co) as
drastically affected by the ligands. In particular, the effects of electron shown in Fig. 2c [43]. Quantitative analysis based on the Co XANES
donating and withdrawing ligands are evaluated based on peripheral spectra revealed that the Co oxidation state reached 2.08 at OCV
functionalized cobalt porphyrins [38]. The performance of immobilized (Fig. 2d). When the potential of − 0.62 V vs. RHE was applied, an
cobalt porphyrins can be improved by introducing electron donating average oxidation state of 1.68 was observed, indicating that partial
ligands on the structures while weakened by electron-withdrawing reduction of Co(II) to Co(I). The results verified that reduced Co was the
groups, which is similar to the observations in homogeneous condi­ active site for CO2ERR.
tions. The improved activities are owing to the densified electrons on the In-situ UV–vis was employed to estimate the formal redox potential
cobalt center contributed by the electron donating groups [39]. Sur­ of metal centers in porphyrins [44,45]. The UV–Vis spectra of the COFs
prisingly, the trend contradicts with porphyrin-based covalent organic comprising Co porphyrins experienced changes attributable to Co(II)/Co
frameworks (COFs), in which electron-withdrawing groups are observed (I) reduction under different potentials (Fig. 2e) [44]. Using the steady-
to promote CO2ERR [40]. The inconsistency here might be due to the state UV–vis response to the reduction potentials in combination with
structures of porphyrin-based COFs are distinctively different from the Nernst equation, the redox potential (E1/2) could be estimated to be
molecular porphyrins. The electron-withdrawing groups are located on more negative than − 0.52 V vs. RHE (Fig. 2f), in agreement with the CV
the organic linkers instead of porphyrin rings. The relatively long dis­ measurements (E1/2 ≈ − 0.67 V vs. RHE). In another case, UV–vis was
tance would undermine the adverse effect of the groups on the active applied to metal organic frameworks (MOFs) comprising Co porphyrins
metal center. Instead, the electrons withdrawn by the groups could at different potentials (Fig. 2g) [45]. The difference spectra were plotted
effectively migrate along the backbones of COFs to the active metal based on the changes of Soret band and subsequently quantified to
center, thus promoting the activities. deduce the formal redox potential (E1/2) of the cobalt center. The formal
In addition, the effect of cationic ligands was investigated via the reduction potential of the Co porphyrin unit in the MOFs was deter­
immobilized cationic porphyrins bearing pyridyl or anilinium. An mined to be − 0.4 V vs. RHE (Fig. 2h), which was consistent with the
improved activity was recorded on these porphyrins, most likely due to position of the first cathodic wave in CV.
the electrostatic stabilization of the transition state by the cationic li­ In-situ Raman could provide information on the structure of por­
gands during the rate-limiting electron transfer step. A further experi­ phyrins during catalysis [45]. For example, in situ surface-enhanced
ment revealed that anionic ligands on porphyrins would have a negative Raman spectroscopy was utilized to confirm the integrity of the
effect on the catalytic activity. Thus the findings provide a robust cri­ porphyrin units in MOFs throughout the catalysis (Fig. 2i) [45]. At each
terion for screening candidate catalysts prior to experimental tests, applied potential, the primary Raman peaks attributed to the porphyrin
which can serve as the guidelines for future work. linker remained in the spectrum, confirming the preservation of the
frameworks.
2.4. In-situ techniques for mechanism study In-situ FTIR could provide insight into the intermediates formed
during CO2ERR, which is fundamental in understanding the catalytic
In-situ techniques provide an experimental way to illustrate the pathways. Such a technique was applied to a COFs material consisting of

3
S. Gu et al. Chemical Engineering Journal 470 (2023) 144249

Fig. 2. (a) XANES spectra of Zn at different potentials (V vs. NHE); (b) Derivatives of the Zn XANES spectra; (c) XANES spectra of STPyP-Co (V vs. RHE); (d)
Oxidation states analysis by edge position fitting of XANES of STPyP-Co; (e) UV–Vis recorded on COFs at potentials from 0.23 to − 0.57 V vs. RHE; (f) First derivative
of a polynomial fit of Δabsorbance vs. potential; (g) UV–vis recorded on MOFs at potentials from 0.2 to − 0.7 V vs. RHE; (h) First derivative of a polynomial fit of
Δabsorbance vs. potential; (i) The Raman spectra of MOFs at potentials from − 0.1 V to − 0.7 V vs. RHE. Adapted with permission from Ref. [42], Copyright 2017,
American Chemical Society; Ref. [43], Copyright 2019, Wiley; Ref. [44], Copyright 2015, Science; Ref. [45], Copyright 2015, American Chemical Society.

copper porphyrins at reducing potentials [46]. The peaks corresponding analysis. For example, information about the rate-determining step
to the OH deformation (1255 cm− 1), C-O stretch (1337 cm− 1), sym­ (RDS) could be provided by Tafel analysis. Tafel slope, often reported in
metric stretch (1398 cm− 1), asymmetric stretch (1567 cm− 1) of *COOH units of mV per decade of current (mV/dec), relates the catalytic current
intermediate were observed, indicating *COOH was a key intermediate. (jcat) to the applied potential (or overpotential) as indicated in Eq. (E1)
And *COH and *CH2O, as crucial intermediates for CO2 reduction to [47]. The equation can be derived via the partial differentiation of
CH4, were also observed at the position of 1035 cm− 1 and 1473 cm− 1. Butler-Volmer equation, a fundamental formula relating the applied
Thus in-situ FTIR provides a feasible way to investigate the course and potential (E) with current density (i) in electrochemical reactions [48].
the mechanism of CO2 reduction to CH4. Assuming the catalytic steps before the RDS are quasi-equilibrium and
the ambient temperature (T) is 298 K, the Tafel slope can be derived as:
3. Kinetic study of CO2ERR on porphyrins [ ]
∂(− E) 2.303RT 2.303RT 59mV/dec
mTafel = = = =
∂log(− i) Ci ,T (1 − α)F F(n + (1 − β)q) n + (1 − β)q
3.1. Kinetics derived from Tafel analysis
(E1)
Information about the decoupled involvement of electron–proton
where n is the number of electrons transferred before the RDS, q is the
transfer of CO2ERR catalyzed by porphyrins could be provided by Tafel

4
S. Gu et al. Chemical Engineering Journal 470 (2023) 144249

number of electrons transferred during the RDS, and β is the symmetry despite the fact that proton delivery constitutes an important step in
factor (usually 0.5). Tafel slope can thus be used to provide information CO2ERR as depicted in Fig. 1. The conclusion is in line with the above
about the RDS such as electron transfer numbers. Tafel slope should be Tafel analysis which assigns either electron or mass transfer as the RDS.
linear for analysis which requires little mass transport limitations and However, it is worth noting that the above studies are conducted under
low coverage of adsorbed species. As such, Tafel analysis should be pH from near neutral (pH = 6.3) to basic (pH = 13.6). In an acidic
conducted at low overpotentials as mass transport limitations become condition, the CO partial current would be drastically affected by pH as
apparent at high overpotentials that would distort Tafel slopes (Fig. 3a). the competitive HER becomes apparent.
For porphyrin catalyzed CO2ERR in heterogeneous conditions, first Most studies on CO2 order dependence have yielded an order of 1
electron transfer is the RDS as depicted in Fig. 1b. This signifies n = 0, q through a wide pressure range of 10 kPa to 100 kPa, which intuitively
= 1 in E1, and a Tafel slope of 118 mV/dec is expected. As a result, a agrees with the need of one CO2 molecule to afford CO or formate
Tafel slope of ~ 118 mV/dec implies that the first electron transfer is the [38,44,50]. In contrast to the above, a recent study revealed the CO
limiting step. Tafel slopes above 118 mV/dec typically imply that mass partial current increased linearly with CO2 pressure (an order of 1)
transport comes into play which distorts the slopes. While for homoge­ below ~ 20.2 kPa, while above the threshold, the partial current
neous porphyrins, two electron transfer comprises the RDS (Fig. 1a) and remained unchanged (an order of 0) [49]. The difference might come
thus the Tafel slope of 59 mV/dec is typical for a rate-limiting electron from a gas-fed flow cell was used here while previous studies were
transfer process. Tafel slopes deviating from 59 mV/dec denotes that conducted in an H-type cell. The structure design of the flow cell greatly
mass transport is becoming a limiting factor. promotes the access of CO2 to the catalytic centers and thus exempts CO2
from being a kinetic limiting step. As a result, CO2 was not involved in
3.2. Kinetics in heterogeneous conditions the RDS at higher partial pressure (>20.2 kPa) for a flow cell. More
importantly, the role of electrolyzer configuration was emphasized as it
The above findings have been applied to heterogeneous porphyrins might significantly impact the kinetics of CO2ERR.
for determination of the RDS successfully. For example, cobalt tetra­ The kinetic studies examining the bicarbonate order dependence
phenylporphyrin (CoTPP) heterogenized on carbon paper at a low found that CO partial current was 0th-order dependence on bicarbonate
loading of 8 × 10− 10 mol/cm2 exhibits a Tafel slope of 119 mV/dec, concentration [38,50]. When the partial pressure of CO2 was adjusted to
indicating that the first electron transfer is involved in the RDS of CO2 101 kPa and 10.1 kPa, respectively, the 0th-order dependence still
reduction [38]. While heterogenization of CoTPP on carbon nanotubes persisted, further confirming bicarbonate was not involved at the RDS as
(CNTs) at a much higher loading of 1.7 × 10− 7 mol/cm2 gives rise to a a reactant or a proton source [49]. It should be noted that the above
Tafel slope of 255 mV/dec [29]. The Tafel slope is heavily deviated from studies were conducted under a constant ionic strength by tuning the
the 118 mV/dec, suggesting that mass transport process such as CO2 sodium bicarbonate and sodium perchlorate ratio in the electrolyte. It is
delivery and product desorption exert substantial effects. A comparison acknowledged that the ionic strength should remain constant as the
of the two cases revealed that lower loadings were necessary to plot strength would impact the activities of ions and affect the catalytic
kinetically interpretable Tafel slopes as higher loadings of porphyrins performance of CO2ERR [51]. Conflicting reports on the order depen­
may prevent the mass transport through the porphyrin aggregates and dence of bicarbonate appeared as these studies were conducted in
thus limit the reaction rates. However, a similar loading study was electrolytes containing different concentrations of bicarbonate with
performed on CoTPP and the Tafel slope was found to be independent of different ionic strengths [52,53]. As such, opposite results showing an
loadings [38]. A Tafel slope of ~ 120 mV/dec was maintained through order of 1 at high bicarbonate concentrations were reported and the
the loadings of 10-10 to 10-8 mol/cm2. The inconsistencies here may be results should be re-examined at a constant ionic strength.
due to the range of loading amount is narrow, which is insufficient for
the change in Tafel slopes to take place. Further increase of the loading 3.3. Kinetics in homogeneous conditions
amount to 10− 7 mol/cm2 would induce a higher Tafel slope as discussed
before. The above understanding of kinetics is based on heterogeneous
Order-dependence studies provide another set of tools to probe conditions. To compare the activities of the homogeneous porphyrins,
which species are involved in the RDS for heterogeneous CO2ERR. The Savéant and his group developed a strategy using “catalytic Tafel plots”
dependence of CO partial current on pH, CO2 pressure, and bicarbonate for the purpose [35,54–56].
concentration provides the basis for electrokinetic understanding. Eq. (E2) exhibits a definite relationship between TOF and over­
A few studies found no pH dependence, indicating CO evolution rate potential that characterizes the catalytic properties of the target catalyst
over these Co porphyrins exhibits 0th-order kinetics on H+ [38,44,49]. [54]. TOF is defined as the amount of product produced per active
This observation implies that proton transfer is not involved in the RDS catalyst per time with units of inverse time. In homogeneous catalysis,

Fig. 3. (a) Plot of overpotential vs. log jcat for Tafel analysis on porphyrins; (b) Plot of TOF vs. overpotential η for three hypothetic catalysts C1, C2, and C3 with
different TOFmax and Ecat/2 values reducing CO2 under the same conditions. Adapted with permission from Ref. [57], Copyright 2018, American Chemical Society.

5
S. Gu et al. Chemical Engineering Journal 470 (2023) 144249

TOF is calculated based on the number of product molecules and active 4. CO2ERR in homogeneous conditions
catalyst molecules present in the reaction–diffusion layer over a certain
period of time. Thus, TOF is derived as E2, displaying as a function of From the roadmap of porphyrins employed in CO2ERR in Fig. 4a, the
overpotential (η). At large overpotentials, TOF increases alongside until use of homogeneous porphyrins can be dated back to 1979 when cobalt
it reaches the plateau value of kcatC0cat. Whereas at zero overpotential, porphyrins with ligands such as carboxyl (CoTCPP) and sulfonate
logTOF0 is defined by (E4). (CoTPP-SULF) were proved to be active for CO2 reduction via CV scans
0
[58]. Since then, the electrochemical reduction of CO2 by metal­
kcat CCO
TOF = 2
(E2) loporphyrins has been widely investigated for more than 40 years.
F 0 0 − η)]
1+ exp[RT (ECO2 /CO − Ecat Caution should be taken when defining homogeneous or heteroge­
neous catalysis. As reported by Wang et al., homogeneously dissolved
At η = maximum, TOF max = kcat C0CO2 (E3) catalyst was heterogenized on the support during electrocatalysis and
acted as the genuine active center, rendering the system as heteroge­
F ( 0 )
neous catalysis [59]. However, owing to the lack of awareness, most of
At η = 0, logTOF 0 = logTOF max − ECO2 /CO − E0cat (E4)
RTln10 previous work fails to recognize the point. To be in line with previous
Eq. (E2) on TOF-overpotential relationship is illustrated in Fig. 3b. work, homogeneous catalysis mentioned in this review refers to the
The catalytic Tafel plots show two asymptotes which match well with cases where metal porphyrins are dissolved in the solvents such as DMF,
previous discussion. TOF increases linearly at low overpotentials until it acetonitrile and etc. to perform CO2ERR. A diagram of the reaction
reaches a plateau value of kcatC0cat at elevated overpotentials. The mechanism is presented as below (Fig. 4b), which clearly displays the
intersection with the axis at zero overpotential corresponds to the TOF0 role of metal porphyrins. The metal porphyrins act as a redox mediator
value, which is an intrinsic indicator of the catalyst performance. With to pass the electrons from the electrode to the reactants at the applied
the goal of improving the TOF and lowering the overpotential of the potential. In a typical process, the porphyrins are reduced and activated
system, a good catalyst would be expected to have catalytic Tafel plots in-situ on the electrode surface, which in turn react with CO2 with cat­
shifted to the upper left as shown in Fig. 3b. As a result, it is clear that C1 alytic rate constant (kcat) depending on the intrinsic activity and the
(green line) represents the best catalyst with the highest TOF throughout extrinsic conditions such as potential, medium, and etc. In homogeneous
the entire potential range. However, it cannot be simply determined CO2ERR, only a small portion of the porphyrins that are close to the
which one of C2 (blue line) and C3 (red line) has a better performance as electrode surface would participate in the catalytic process, leading to a
their TOF functions reach an intersection at η = 0.26 V. Catalyst C3 has a low utilization of catalysts. Different metal centers are employed in the
higher E0cat value and a lower TOFmax compared with C2, which results in porphyrin structures for evaluation of the catalytic performance and
a better performance at η < 0.26 V and an inferior performance at η > selectivity. In this section, porphyrins with iron, cobalt, nickel and zinc
0.26 V. centers are discussed and the benchmark for catalyst evaluation is
presented.

Fig. 4. (a) Timeline of the development of various porphyrin catalysts for CO2ERR; Porphyrin catalyzed CO2ERR via (b) homogeneous pathways; (c) heterogeneous
pathways (M = active metal sites in porphyrins). Adapted with permission from Ref. [20], Copyright 2020, Royal Society of Chemistry.

6
S. Gu et al. Chemical Engineering Journal 470 (2023) 144249

4.1. Iron porphyrins C atom is subjected to protonation, resulting in the breaking of the Fe-C
bond instead of the C-O bond, which contributes to the formation of
The early study on the tetraphenyl porphyrin (TPP) complex FeTPP formic acid.
was carried out in the absence of proton donors in a NEt4ClO4/DMF Based on the above observations, modifications of the structure of
electrolyte, whereby the catalytic formation of CO was ascribed to FeTPP to provide local environment similar to the one brought by Lewis
[Fe0(TPP)]2- (Entry 1, Table 1) [60]. However, the catalytic process acid and Brønsted acid were investigated. In an attempt to achieve this,
under aprotic conditions is accompanied by a modest catalytic response phenolic groups were installed in the ortho and ortho′ positions of FeTPP
and rapid deactivation of the catalyst, presumably caused by carboxyl­ to afford FeTDHPP as shown in Entry 2 in Table 1 [33]. A significant
ation and hydrogenation of the ligand. Therefore, a low faradaic effi­ improvement of the catalytic performance was observed for FeTDHPP
ciency (FE) of CO formation and poor catalyst stability were observed owing to dual roles of the phenolic groups, both acting as intramolecular
for the process. hydrogen bond donors and the internal proton-transferring relays. It was
Later in the 1990 s, it was found that the addition of a mono- (Li+, demonstrated that the phenolic ligands helped to stabilize the Fe(II)–
Na+) or divalent Lewis acid (Mg2+, Ca2+, Ba2+) to the electrolyte could CO2 adduct by intramolecular hydrogen bonding, which was confirmed
enhance the activity and stability of FeTPP [61,62]. A comparison of by Density Functional Theory (DFT) calculations. As a consequence of
several Lewis acids leads to the finding that the enhancement in activity the stabilized adduct, electron transfer for the second reduction was
follows the order of Mg2+ ≈ Ca2+ > Ba2+ > Li+ > Na+. As revealed by more difficult than the first one. The second reduction took place along
long run electrolysis, CO and formate are formed as the major product with internal proton transfer and C-O bond breaking, which was fol­
and byproduct, whereby the exact product ratio depends on the type of lowed by the reprotonation of the ligand phenols [65]. The effects of
Lewis acid. phenolic ligands was further validated via a comparison of FeTDHPP
The CV studies have been employed to investigate the role of Lewis with methoxyl-substituted equivalent FeTDMPP, which did not consti­
acid in the catalysis. As shown in Fig. 5a, the scan of negative potential tute a hydrogen bond donor (Entry 3, Table 1) [33]. It turned out the
in the absence of CO2 affords three reversible redox couples, denoting turnover frequency (TOF) of FeTDHPP was 9 orders of magnitude higher
[FeII(TPP)], [FeI(TPP)]-, and [Fe0(TPP)]2-. In the presence of CO2 and a than that of FeTDMPP.
Lewis acid Mg2+, a large catalytic current appeared at the potential of The influence of electron-withdrawing ligands on the FeTPP-
the [Fe0(TPP)]2- species, indicating that catalysis was triggered by the catalyzed CO2-to-CO conversion was studied by successive replace­
last electron transfer step. Next, the metal center of the [Fe0(TPP)]2- ment of the phenyl rings by one, two, and four perfluorophenyl groups
species underwent a nucleophilic attack by the electrophilic CO2 and (Entry 4–6, Table 1) [66]. Two opposite trends were observed upon
formed an intermediate adduct Fe(II)–CO2. In the case of Mg2+ as the increasing the number of electron-withdrawing ligands at a given con­
Lewis acid, the Fe(II)–CO2 adduct reacted with another CO2 molecule centration of PhOH. First, the overpotential was reduced due to the
with the assistance of Mg2+ (Fig. 5b), which ended up in a Fe(II) positive shift of the redox potential of the Fe(I)/Fe(0) couple caused by
carbonyl complex. The formation of the ion pair between the negatively the electron withdrawing substituents, thus favouring the electro­
charged oxygen atoms in the adduct and positively charged Mg2+ catalysis. Second, the electron-withdrawing substituents decreased the
allowed the Lewis acid to pull out an electron pair which had been electron density on the Fe(0) center of the catalyst, which would lower
pushed from Fe(0) to CO2 in the previous step, thereby weakening the C- the nucleophilicity of the metal center for binding CO2 and thus restrict
O bonds of CO2 and facilitating the release of CO. As a result, the the catalytic rate.
addition of cation Mg2+ can be deemed as an efficient Lewis acid and the In order to maximize the benefit of perfluorophenyl groups while
formation of the ion pair assisted by Mg2+ is responsible for the accel­ minimizing the adverse effect, iron porphyrin with two perfluorophenyl
eration of the catalytic process. groups at the opposite positions while preserving hydroxyl groups on the
An even more remarkable enhancement on catalytic performance of other two phenyls was synthesized (Entry 7, Table 1) [36]. The afforded
FeTPP was observed by the addition of weak Brønsted acids (CF3CH2OH, catalyst FCAT thus showed an Eonset at 150 mV less than that for FeTPP
PhOH, 1-propanol, 2-pyrrolidine), which resulted in highly selective CO owing to the fluorinated phenyls and displayed an increased TOF
production without significant formation of hydrogen [56,63,64]. The number than FeTPP owing to the phenol groups. Combining the benefits
product distribution was found to be dependent on acidity, where use of from perfluorophenyls and phenols, the catalyst is highly efficient for
1-propanol in bulk electrolysis led to the generation of CO (FE ≈ 60%) CO2 reduction to CO in homogeneous conditions.
and formate (FE ≈ 35%), the use of the more acidic CF3CH2OH afforded A further improvement of the FeTPP catalyst was studied by Cost­
CO almost exclusively (FE > 96%). entin et al. using the ionic moieties to substitute the phenyl pendants.
On the basis of voltammetry analysis, the role of Brønsted acid (HA) The positively charged trimethylammonium (TMA) groups were intro­
in the electrocatalysis was elucidated. The CV performed under CO2 in duced to the para-positions of meso-phenyls to afford Fe-p-TMA as
the presence of Brønsted acid (HA) afforded an increased catalytic cur­ indicated in Entry 8 of Table 1 [67]. The catalyst exhibited a high
rent at the potential of [Fe0(TPP)]2− species, similar to that was selectivity of 98–100% in CO2-to-CO conversion in nonaqueous DMF
observed in the presence of Lewis acid (Fig. 5a). The increased current is electrolyte containing proton donors or aqueous electrolyte [67]. The
ascribed to the proposed synergistic action of Brønsted acid as illustrated good stability and high TOFCO number attained in aqueous electrolyte
in Fig. 6. The negatively charged O in the adduct Fe(II)–CO2 forms were remarkable with the competing HER being greatly inhibited.
hydrogen bonds with two HA molecules. The hydrogen bonds weaken Further study of the effect of ionic groups on CO2ERR led to the finding
the C-O bonds in the adduct by pulling out the electrons and facilitate that trimethylammonium groups installed in the ortho- rather than para-
the formation of CO product. The nature of the synergistic effect of weak positions of all four phenyl rings would induce a drastic enhancement in
Brønsted acids is therefore the same as that of Lewis acids described the catalytic performance (Entry 9, Table 1) [30]. The overpotential for
above. In contrast to Lewis acids, the use of Brønsted acids leads to water CO2ERR was significantly reduced to 220 mV while the TOFco increased
formation instead of precipitates, which would not block the electrode to 106 s− 1 in DMF electrolyte with PhOH as the proton donor. In com­
surface. The dehydration step in Fig. 6 was then studied by Savéant [56] parison with Fe-p-TMA, the catalytic current of Fe-o-TMA was three
and a PCET mechanism was put forward. times higher. It was proposed that shift of the TMA groups from the para
The use of a weaker Brønsted acid such as 1-propanol often brings to the ortho-positions contributed to the Coulombic stabilization of the
about the formation of formic acid. The phenomenon could be explained intermediate Fe-CO2 adduct (Fig. 7a, b), which made it the most efficient
by the above analysis. As shown in Fig. 6, a weaker Brønsted acid lacks electrocatalyst for homogeneous CO2-to-CO conversion reported to date.
the efficiency in pulling out the electrons from the CO2 ligand in the Other types of ionic moieties such as negatively charged sulfonate
intermediate adduct, leading to a higher basicity of the C atom. Thus the groups and their effects on CO2ERR were also investigated [30]. The

7
S. Gu et al. Chemical Engineering Journal 470 (2023) 144249

Table 1
Fe porphyrins as homogeneous catalysts for CO2ERR.
Entry Catalyst Structure Electrolyte V vs. NHE FE TOF Product Ref.

1 FeTPP NEt4ClO4/DMF N.A. N.A. N.A. CO [60]

1
2 FeTDHPP NBu4PF6/DMF − 1.16 V 94% 3162.3 s− CO [33]

1
3 FeTDMPP NBu4PF6/DMF − 1.42 V N.A. 20.0 s− CO [33]

4 FeF5TPP NBu4PF6/DMF N.A. N.A. 10-9-102.5 s− 1


CO [66]

5 FeF10TPP NBu4PF6/DMF − 1.26 V 90% 10-9-101 s− 1


CO [66]

6 FeF20TPP NBu4PF6/DMF N.A. N.A. 10-7.5-10-0.5 s− 1


CO [66]

1
7 FCAT NBu4PF6/DMF − 1.08 V 100 ± 10% 240 s− CO [36]

(continued on next page)

8
S. Gu et al. Chemical Engineering Journal 470 (2023) 144249

Table 1 (continued )
Entry Catalyst Structure Electrolyte V vs. NHE FE TOF Product Ref.

8 Fe-p-TMA 0.1 M KCl − 0.86 V 98–100% 10-4.5-104.5 s− 1


CO [67]

9 Fe-o-TMA NBu4PF6/DMF − 0.96 V 100% 102.5-106 s− 1


CO [30]

10 Fe-p-PSULF NBu4PF6/DMF N.A. N.A. N.A. CO [30]

1
11 o-Fe2DTPP 10 %H2O /DMF − 1.35 V 95% 4300 s− CO [68]

Fig. 5. (a) 100 mV/s CV scans of 1 mM FeTPP in DMF + 0.1 M NEt4ClO4 in the absence of CO2 (left) and in the presence of CO2 and Mg2+ (right). (b) Proposed
synergistic action of the iron metal center and Mg2+ on the C-O bond-breaking process. Adapted with permission from Ref. [61], Copyright 1996, American
Chemical Society.

structure of the catalyst is presented in Entry 10 of Table 1. Unlike TMA


groups, the sulfonate groups of Fe-p-PSULF exert negligible effect on the
standard potential of the catalyst (E0cat = − 1.428 V vs SHE). The un­
derlying reason is that there is no Coulombic interaction between the
sulfonates and the negative charges provided by the intermediate Fe(0)–
CO2 adduct (Fig. 7c).
With particular attention on CO dehydrogenase, efforts were made to
synthesize new electrocatalysts in an attempt to emulate the natural
systems. This has been achieved by a ligand featuring two covalently
linked co-facial porphyrin macrocycles as indicated in Entry 11, Table 1
[68]. CV performed on o-Fe2DTPP catalyst in DMF under Ar showed at
Fig. 6. Proposed synergistic action of iron metal center and Brønsted acids on least 3 quasi-reversible redox peaks, arising from overlapping reduction
the breaking of C-O bond. Adapted with permission from Ref. [57], Copyright of the o-Fe3+
2 DTPP ions to the ultimately reduced species Fe2DTPP.
0
2018, American Chemical Society.

9
S. Gu et al. Chemical Engineering Journal 470 (2023) 144249

Fig. 7. The Coulombic interaction between the ligands of (a) Fe-p-TMA (b) Fe-o-TMA (c) Fe-p-PSULF and the adsorbed CO2.

Meanwhile CV under CO2 showed a new wave after reduction to the section. For nickel porphyrins, most of the studies are conducted in
Fe2+
2 DTPP state, possibly due to pre-association of CO2. A 10 h long run heterogeneous conditions. There is a lack of reports of nickel porphyrins
electrolysis was conducted in H2O:DMF (1:9) electrolyte at − 1.35 V vs. operating as homogeneous catalysts for CO2ERR. The majority of the
NHE, which afforded FECO of 95% [68]. nickel based catalysts for homogeneous CO2ERR are Ni complexes with
cyclams, pincer ligands and clusters [69]. In addition to the common
4.2. Porphyrins with other metal centers metal centers such as iron, cobalt and nickel, zinc is one of the less
common metals that has been introduced to porphyrins for CO2ERR.
Cobalt and nickel are frequently used as the metal centers in por­ One report based on zinc porphyrin demonstrated its promising activity
phyrins besides iron. However, porphyrins with cobalt or nickel centers in H2O:DMF (1:9) for CO2 reduction [42]. However the catalyst was
are mostly applied to heterogeneous catalysis. For cobalt porphyrins, noncovalently deposited on polytetrafluoroethylene-treated carbon
there is a scarce of examples operating as homogeneous catalysts in fibre paper and thus is discussed in more detail in the following section.
solvents for CO2ERR. Instead, water soluble porphyrins such as CoTCPP Another report involved AgII and PdII octa-ethylporphyrins as the ho­
and CoTPP-SULF were synthesized and their activities for CO2ERR in mogeneous catalysts for CO2 reduction to oxalate in CH2Cl2 (Entry 3&4,
aqueous electrolyte were reported (Entry 1&2, Table 2). CV performed Table 2) [70]. Notably no CO was detected after electrocatalysis, indi­
in aqueous solution showed elevated currents upon shifting from N2 to cating the important role of metal centers on the reduction products.
CO2 atmosphere, with evidence of HCOOH formation, but the products The porphyrin catalysts discussed above utilize earth-abundant
were not quantified [58]. In comparison, heterogeneous electrocatalysis metals such as Fe, Co, Ni, Zn, thus reducing the cost greatly compared
via immobilization of cobalt porphyrins on conductive supports has with noble metal-based catalysts. Besides, iron porphyrins display high
been reported as an effective way for conducting CO2ERR. The examples selectivities of over 90% towards CO2 reduction to CO in general, which
of heterogenized cobalt porphyrins are presented in the following verifies the great potential of such catalysts in the process of

Table 2
Porphyrin with different metal centers as homogeneous catalysts for CO2ERR.
Entry Catalyst Structure Electrolyte V vs. NHE FE TOF Product Ref.

1 CoTCPP Phosphate buffer N.A. N.A. N.A. HCOOH [58]

2 CoTPP-SULF Phosphate buffer N.A. N.A. N.A. HCOOH [58]

3 AgOEP TBAF/CH2Cl2 − 1.255 V N.A. N.A. Oxalate [70]

4 PdOEP TBAF/CH2Cl2 − 1.405 V N.A. N.A. Oxalate [70]

10
S. Gu et al. Chemical Engineering Journal 470 (2023) 144249

industrialization. However, homogeneous catalysis usually requires the 70% could be achieved at the current density of 25 mA/cm2 under at­
use of organic solvents as the electrolyte, which would bring potential mospheric CO2 pressure.
pollution to the environment. With growing concern over environ­ In a step further, tetraphenylporphyrins with different metal centers
mental issues, heterogeneous catalysis provides an alternative approach (MTPP) such as Co, Fe, Ni, Zn and Cu were dip-coated on carbon black
as less-polluting aqueous electrolytes are involved. Although research based gas diffusion electrodes (GDE) for evaluation of activity towards
work on heterogeneous catalysis has just started in recent decades, it CO2ERR under atmospheric and 20 atm of CO2, respectively (Entry 3,
presents a promising future direction for the field. Table 3) [24]. For immobilized CoTPP and FeTPP that were compara­
tively active under atmospheric CO2, their current efficiencies for the
5. CO2ERR in heterogeneous conditions CO2 reduction increased up to 97.4% and 84.6%, respectively, under 20
atm CO2. For NiTPP, ZnTPP and CuTPP that showed low activity under
Compared with homogeneous catalysis where porphyrins operate in atmospheric CO2, their current efficiencies for CO2 reduction increased
the same phase with CO2, heterogeneous catalysis requires immobili­ up to 18%, 65.8% and 50.5%, respectively, under 20 atm CO2.
zation of porphyrins on solid supports which results in catalysts oper­ Among different metal porphyrins, copper porphyrin was found to
ating in a separate phase from CO2. The limitation of homogeneous catalyze CO2 reduction to hydrocarbons in addition to common product
catalysis lies in the low utilization of porphyrins owing to the fact that of CO in heterogeneous conditions. A remarkably high production of
limited porphyrins in the first diffusion layer at the electrode surface can methane and ethylene was reported by using copper porphyrin with
be electro-activated. In contrast, heterogeneous catalysis offers the hydroxyl pendants [CuTPP-(OH)8] dropcasted on carbon fiber paper
possibility of circumventing the limitation as the porphyrins are (Entry 4, Table 3) [79]. This electrocatalyst could convert CO2 to
confined on the electrode surface which allows the major part to be methane and ethylene at − 1.4 V vs. NHE with a TOF of 4.3 and 1.8 s− 1,
activated [71]. In a typical heterogeneous catalytic process, the catalyst respectively, accounting for 44% of the conversion rate in total. The
porphyrins no longer diffuse in the solution but stay on the electrode hydroxyl groups were speculated to be vital in the production of hy­
surface, rendering the electrode surface as the active site for CO2 drocarbons and stabilization of the *CHO and *CHO-CO intermediates,
reduction. (Fig. 4c). which are essential precursors for methane and ethylene products.
The immobilization of porphyrins onto supporting electrodes re­ Wang et al. studied the active sites during CO2ERR using zinc(II)
quires an appropriate selection of the supports [72,73]. The choice of 5,10,15,20-tetramesitylporphyrin (ZnPor) dropcasted on carbon fiber
supports depends on the properties of the materials such as surface paper for CO2 reduction to CO (Entry 5, Table 3) [42]. ZnPor afforded a
chemistry, conductivity and stability under reducing conditions [74]. FECO of 95% and a current density of 2.1 mA/cm2 at the potential of
There are two types of materials mainly used as the supports, metal − 1.70 V vs. NHE in the electrolyte of H2O:DMF (1:9). While the metal
oxides and carbon-based materials [75,76]. Carbon-based materials are center is usually considered to be vital in electrocatalysis, in-situ spec­
especially suitable as a cathode for immobilization of porphyrin cata­ troscopy analysis revealed that zinc was redox-innocent throughout the
lysts because of their high overpotential for the competing HER and potential range. Instead, porphyrin ligand was identified as the redox
their stability under reducing conditions. mediator by CV analysis. The findings here is inconsistent with previous
To immobilize porphyrins on the supports, there are three strategies research which emphasizes the importance of metal centers in defining
frequently employed including non-covalent, covalent and periodic activity and selectivity. It was thus proposed, similar to metal centers,
polymerization. The following sections focus on the differences between the macrocyclic porphyrin rings may also act as a reservoir to store and
the three strategies and the corresponding examples. deliver electrons for CO2 reduction. The results suggest that the
macrocyclic rings could have equally important effects on the activity
5.1. Noncovalent immobilization strategy and selectivity as the metal centers, which underlies the possibility of
designing a wider range of molecular catalytic systems for CO2ERR.
Noncovalent strategy has been employed to immobilize porphyrins Shen et al. proposed a mechanism for the heterogeneous CO2ERR
without damaging the intrinsic properties of the supports. The immo­ using cobalt protoporphyrin (CoPP) dip-coated on pyrolytic graphite
bilization utilizes the strong affinity of porphyrins to supports via π-π (Entry 6, Table 3) [80]. The immobilized CoPP could reduce CO2 to CO,
interactions and can be achieved in a convenient and straightforward CH4 and CH3OH in an aqueous electrolyte at a moderate overpotential of
way such as dropcasting or dip-coating. The electron transport rates can 500 mV. The study demonstrated the catalyst-bound CO⋅−2 radical anion
be improved due to the closer proximity of porphyrins to the electrode had a strong Brønsted-base character and its formation could be affected
and the electron conductivity can be enhanced from the in-plane π-π by the pH of the electrolyte significantly. The pH dependence of CO2ERR
stacking. The immobilized porphyrins also minimize the contact be­ product distribution was found to be different from the competing HER.
tween the supports and electrolyte, hence reducing the opportunity of Compared with H2 which was predominantly produced at pH = 1, they
HER on supporting electrodes [77]. identified CO production was catalyzed at pH = 3, and CH4 production
The common techniques employed for non-covalent immobilization was catalyzed at pH = 1. As a result, a key strategy in achieving a high FE
include dip-coating and dropcasting. The former is achieved via soaking of such a catalyst lies in lowering the potential for the formation of
supporting electrodes in the solvent of catalyst for a specific time fol­ CO⋅−2 radical anion, stabilizing the binding of the anion to the catalyst,
lowed by drying while the latter involves dropping a portion of catalyst and adjusting the pH range.
solution onto the surface of the supporting electrodes followed by Until now, the strategy of non-covalent immobilization is realized
evaporation of the solvent. Compared with dip-coating, dropcasting almost exclusively on carbonaceous supports. Carbon paper, carbon
allows for a facile determination of the catalyst loading amount. cloth, GC, pyrolytic graphite, activated carbon fibers, and etc are
CoTPP was immobilized on CNTs via dropcasting and an improve­ commonly used as the supports [17]. In a study conducted by Aoi et al.,
ment in the activity was observed compared with homogeneous situa­ supports with different surface morphologies and graphitic degrees were
tion (Entry 1, Table 3) [29]. In the heterogeneous phase, CoTPP showed compared for porphyrin immobilization [81]. It was found that there
improved activities towards CO2 reduction to CO with a FECO of 90% was a significant increase in FECO when the catalyst cobalt chlorin
and a long-term stability at a lower overpotential, whereas in the ho­ (CoCh) was dropcasted on multi-wall CNTs compared to a graphene
mogeneous phase CoTPP produced CO with a compromised FECO of oxide matrix under similar operating conditions (Entry 7, Table 3). This
50%. Derivatization of CoTPP with methoxyl groups could afford enhancement in FECO was attributed to the higher graphitic degree of
CoTMPP, which was demonstrated to be an active catalyst for CO2 CNTs, which led to elevated π-π interactions between CoCh and the
reduction to CO after immobilization on activated carbon fiber paper support CNTs. A similar phenomena was observed by Hu et al. when
(Entry 2, Table 3) [78]. At the potential of − 1.056 V vs. NHE, FECO of different supports such as GC, carbon black (CB) and CNTs were

11
S. Gu et al. Chemical Engineering Journal 470 (2023) 144249

Table 3
Summary of non-covalently immobilized porphyrins for CO2ERR.
Entry Catalyst Support Structure V vs. NHE FE TOF Product Ref.
− 1
1 CoTPP CNTs − 1.106 V 90% 0.078 s CO [29]

2 CoTMPP Carbon fiber − 1.056 V 70% N.A. CO [78]


paper

3 MTPP (M = Co, Carbon black − 0.76 to CoTPP (97.4%); FeTPP (84.6%); N.A. CO [24]
Fe, Ni, Zn, Cu) − 1.14 V NiTPP (18%); ZnTPP (65.8%);
CuTPP (50.5%)

4 CuTPP-(OH)8 Carbon fiber − 1.4 V CO (10%); CH4 (25%);C2H4 (19%) CH4 (4.3 s− 1); CO, CH4, [79]
paper C2H4 (1.8 s− 1) C2H4

1
5 ZnPor Carbon fiber − 1.70 V 95% 14.4 s− CO [42]
paper

6 CoPP Pyrolytic − 0.977 V CO (40%); CH4 (2.4%); CH3OH (N. CO (0.8 s− 1) CO, CH4, [80]
graphite A.) CH3OH

7 CoCh CNTs − 1.1 V 89% N.A. CO [81]

(continued on next page)

12
S. Gu et al. Chemical Engineering Journal 470 (2023) 144249

Table 3 (continued )
Entry Catalyst Support Structure V vs. NHE FE TOF Product Ref.

8 CoTPP Glassy carbon − 1.106 V 28% N.A. CO [29]

9 CoTPP Carbon black − 1.106 V 97% N.A. CO [29]

10 InPP Pyrolytic − 2.066 V 75% (DDAB); 60% (P4VP); 55% N.A. HCOOH [82]
graphite (PEDOT:PSS); 40% (Nafion)

1
11 CATpyr CNTs − 1.03 V 97% 0.04 s− CO [84]

1
12 CoTPyPP CNTs − 1.025 V 95% 2.10 s− CO [85]

1
13 Fe-p-TMA Graphene − 0.924 V 97% 2.9 s− CO [86]

1
14 Fe-p-TMA Graphene − 0.791 V 96.2% 0.5 s− CO [87]
hydrogel

employed for immobilization of CoTPP (Entry 1 & 8 & 9, Table 3) [29]. afforded a comparable selectivity for CO2 reduction to CO (FECO = 97%)
GC loaded CoTPP produced CO at a poor selectivity (FECO = 28%) and but a lower current density compared with CNTs, owing to the catalyst
low current densities compared with CNTs, most likely due to the low detachment from the CB scaffold during electrolysis, whereas catalyst
surface area of GC which was adverse to CoTPP immobilization. CB was more firmly connected with CNTs.

13
S. Gu et al. Chemical Engineering Journal 470 (2023) 144249

Deposition of porphyrin catalysts onto polymeric films on the surface optimization.


of supports has emerged as a viable way for non-covalent immobiliza­ Apart from π-π interactions, electrostatic interactions were also
tion. With a careful selection of the polymeric films, proton transfer for employed for non-covalent immobilization. For instance, the positively
CO2 reduction could be facilitated and CO2 gas diffusion and product charged backbone of tetraphenyltrimethylammonium (TMA) porphyrin
transport could be enhanced. In the research conducted by Birdja et al., enables the complex to be immobilized onto a support via both elec­
indium(III) protoporphyrin IX (InPP) was immobilized on pyrolytic trostatic and π-π interactions. An example reported was the immobili­
graphite via the polymer membranes formed by Nafion, didodecyldi­ zation of an Fe tetraphenyltrimethylammonium porphyrin (Fe-p-TMA)
methylammonium bromide (DDAB), poly(4-vinylpyridine) (P4VP), and on graphene oxide framework for CO2 reduction to CO (Fig. 8a) [86].
poly(3,4-ethylenedioxythiophene) polystyrenesulfonate (PEDOT:PSS), The afforded electrocatalyst (FePGF) offered an outstanding catalytic
respectively (Entry 10, Table 3) [82]. In comparison with polymer-free performance for CO production with a FECO of 97.0% at the potential of
InPP, encapsulation of InPP in the polymers of P4VP and PEDOT:PSS − 0.924 V vs. NHE with long-term stability of 24 h (Entry 13, Table 3).
afforded an enhanced selectivity and activity towards CO2 reduction to The good performance was attributed to the highly delocalized electron
HCOOH, whereas DDAB exhibited a temporary good selectivity which density and enhanced electron transfer via immobilization of the cata­
dropped quickly and Nafion exhibited the worst selectivity. The differ­ lyst onto the well-constructed graphene oxide framework. In a step
ence might be due to the presence of axial coordination to indium center further, Fe-p-TMA was dispersed onto a graphene hydrogel via electro­
by electron donation in P4VP and PEDOT:PSS, while this interaction is static interactions to form a three dimensional framework (Fig. 8b) [87].
absent in Nafion and DDAB. The immobilization technique had also The hybrid catalyst (FePGH) exhibited a good stability of over 20 h as
been applied to cobalt phthalocyanine (CoPc) via the polymeric P4VP well as a remarkable selectivity of 96.2% for CO production at an even
films formed on the surface of edge-plane graphite (EPG) [83]. The smaller potential of − 0.791 V vs. NHE (Entry 14, Table 3). The activity
polymer encapsulated CoPc displayed a FECO of 90% and a TOFCO of 4.8 improvement at an even lower overpotential was ascribed to the 3D
s− 1 at − 1.01 vs. NHE, which were drastically improved over the FECO structures of graphene hydrogels which afforded conductive pathway,
(36%) and TOFCO (0.6 s− 1) by the EPG adsorbed CoPc alone. efficient gas diffusion and good electrolyte accessibility.
Derivatization of porphyrins with an anchoring group that could As summarized above, non-covalent immobilization could be ach­
offer strong π-π interactions with the supports has been attempted for ieved via π-π stacking, polymer encapsulation, electrostatic interaction,
porphyrin immobilization. Among various anchoring groups, the pyrene and etc. Based on the comparison, it could be inferred the porphyrin
moiety has been particularly popular due to the strong π-π interactions structures would have an effect on the performance. Also, a porous and
with carbonaceous supports. A pyrene-appended iron porphyrin (CAT conductive support is necessary for a better performance as indicated in
pyr) was efficiently immobilized on CNTs via non-covalent interactions Entries 13, 14 in Table 3. The above conclusion is further confirmed by
and its catalytic performance was evaluated (Entry 11, Table 3) [84]. variation of the supports in Entries 1, 8, 9 in Table 3. Among different
CATpyr catalyzed CO2 reduction to CO with a FECO of 97% and TONCO of supports, CNTs stand out due to good conductivity and high surface
813 at the current density of 0.2 mA/cm2 for 12 h. Importantly, CATpyr area. However, CNTs are powder-based and inappropriate to be used as
showed a higher selectivity and activity towards CO2ERR compared to electrode directly. Thus, finding new materials that could inherit the
iron porphyrin (CAT) with a similar structure but lacking the pyrene merits of CNTs while maintain mechanical strengths is the next target.
units. The effect of pyrene moiety on porphyrin activity was further Based on Entries 11, 12 in Table 3, strong interactions such as π-π in­
studied by Wang group (Entry 12, Table 3) [85]. The peripheral pyrene teractions between porphyrins and supports were found to enhance the
groups on 5,10,15,20-tetrakis(4-(pyren-1-yl)phenyl)porphyrin cobalt electron transfer as well as activity. Realizing the importance of a firm
(CoTPyPP) was found to enhance π-conjugation and activity. After being interaction, investigations have been made to reinforce the interactions,
immobilized on CNTs, CoTPyPP could catalyze CO2 reduction to CO among which covalent immobilization has attracted increasing interest.
with a FECO of 95% and a TOFCO of 2.1 s− 1 at − 1.025 V vs. NHE. The Although different techniques were explored for non-covalent
observations indicate that the strong interaction between pyrene and immobilization, the outlined examples invariably suffer from catalyst
π-conjugated carbon supports is important for performance leaching, sluggish electron transfer and poor stability. The common

Fig. 8. Schematic illustration of the synthesis of (a) FePGF (Fe-p-TMA/graphene framework) and (b) FePGH (Fe-p-TMA/graphene hydrogel). Adapted with
permission from Ref. [86], Copyright 2018, Wiley and Ref. [87], Copyright 2019, Royal Society of Chemistry.

14
S. Gu et al. Chemical Engineering Journal 470 (2023) 144249

issues are intrinsic of non-covalent immobilization as the non-covalent can be introduced to graphitic edges via the Hassner-type addition of IN3
bonds provide relatively weak interactions between porphyrins and which is generated in-situ and the re-aromatization of the surface
supports. Although interactions along with stability could be enhanced (Fig. 9d) [99].
via introduction of π-conjugated supports, another issue of catalyst There is a great potential in utilizing the classical chemical/elec­
stacking still exists. The stacking issue would limit the catalyst loading trochemical reactions for covalent immobilization as they enable the
and weaken the performance. In a bid to avoid the weaknesses, covalent formation of covalent bonds between metal porphyrins and supports.
bonds were proposed as an alternative method to immobilize The applications of the following reactions such as C-N coupling re­
porphyrins. actions (Buchwald–Hartwig, Ullmann, etc.) [100,101], click chemistry
(copper(I) catalyzed alkyne-azide cycloaddition reaction, etc.) [102],
reduction of diazonium salts [103], and etc. [50,104], to porphyrin
5.2. Covalent immobilization
immobilization have been demonstrated in recent research.
Owing to the difficulty in the synthetic routes, the study of covalent
Covalent immobilization of porphyrins is typically realized by a
immobilization of porphyrin appeared more recently as indicated in
direct covalent bond between porphyrins and supports [88]. The bond
Fig. 4a. The pioneering work was achieved via axial coordination of the
formation is comprised of below steps - functionalization of the support
cobalt center of CoTPP to pyridine groups on the carbonaceous supports.
surface and porphyrins followed by a chemical reaction between sup­
In a typical procedure, 4-aminopyridine was anchored on carboxylated
ports and porphyrins. Compared with non-covalent immobilization
GC via amide bond, followed by the coordination of CoTPP to the pyr­
which mainly uses relatively weak π-π interactions, covalent immobili­
idine functionalized GC (Entry 1, Table 4) [105]. The afforded electrode
zation relies on the more robust covalent bond, and thus could overcome
demonstrated a FECO of 50% and a TONCO of 105 after 4 h electrolysis
the drawbacks such as catalyst leaching, stacking, poor stability, insuf­
with no apparent decrease in current density at − 0.956 V vs. NHE. The
ficient electron transfer and kinetically slow catalytic process which are
improvement in catalytic activity was attributed to the increased elec­
frequently encountered in non-covalent immobilization [89,90].
tron density on the cobalt center after axial coordination to the electron
The techniques utilized for covalent immobilization include chemi­
donating pyridine units, which helps to stabilize the binding of CO2 to
cal and electrochemical reactions. In both cases generating functional
CoTPP. With optimizations of the operating conditions, the above
groups on carbon supports is required. The common methods for
electrode offered a high catalytic activity for CO2 reduction to CO at the
introducing functionals on carbon supports are summarized below
potential of − 0.856 V vs. NHE with a FECO of 92% and TONCO of 107
(Fig. 9). For example, carboxylic groups can be introduced on the carbon
(Entry 2, Table 4) [106]. In the following work, an oxidation reaction
supports under oxidizing conditions (Fig. 9a) [91–93]. Functional
was employed for modification of GC surface through which the amino
groups (R) such as amines, carboxylic acids, azides, or alkynes can be
pyridyl compounds were chemically bonded to GC via the formation of
introduced onto GC, carbon fiber papers, or highly oriented pyrolytic
amine cation radicals. Then CoTPP was coordinated to the pyridine
graphite via reduction of diazonium compounds of aryls bearing
groups on GC in a similar manner (Entry 3, Table 4) [107]. The modified
different R groups (Fig. 9b) [94,95]. Simultaneous introduction of wo
electrode catalyzed CO production at the current efficiency of 63% at
different functionals (R1 and R2) onto CNTs, carbon nanohorns, and
− 0.856 V vs. NHE in phosphate buffer electrolyte at pH = 6.3.
ordered mesoporous carbon surfaces can be achieved via 1,3-dipolar
Via the formation of amide bond, Fe(III) tetraphenylporphyrin with 6
cycloaddition of azomethine ylides (Fig. 9c) [96–98]. Azide groups
pendant hydroxyls and 1 carboxyl (CATCO2H) was covalently attached to
amino modified CNTs (Entry 4, Table 4) [101]. The covalent immobi­
lization afforded a surface concentration of 6.4 × 10− 9 mol/cm2 of
CATCO2H which exhibited a FECO of 95% and a TOF of 178 h− 1 at the
potential of − 1.06 V vs. NHE in aqueous 0.5 M NaHCO3. The high
selectivity and activity of the supported CATCO2H underline the potential
of grafting methods for CO2ERR in neutral water at low overpotentials.
Similarly, cobalt tetra(4-amino)phenyl porphyrin (CoTAP) was
covalently immobilized on carboxyl functionalized CNTs via the amide
bond (Entry 5, Table 4) [39]. The covalently grafted CoTAP exhibited a
FECO of ~ 100% and a TOF of 6.0 s− 1 at the overpotential of 550 mV in
0.5 M KHCO3. In comparison with non-covalently immobilized coun­
terpart which only presented a FECO of 85% and a TOF of 2.3 s− 1 at the
same conditions, a considerable improvement in the activity was ach­
ieved in covalent mode. Tafel analysis combined with microscopy im­
ages confirmed that formation of covalent bond mitigated the
aggregation of CoTAP and enhanced the intermolecular electron transfer
at the same time.
Click chemistry was utilized for covalent immobilization of cobalt
tetra(4-ethynyl)phenyl porphyrin (CoPoralkyne) on azide-terminated
diamond surfaces (Entry 6, Table 4) [102]. The copper (I) catalyzed
alkyne-azide cycloaddition afforded a monolayer of CoPoralkyne on dia­
mond, serving as a “smart” electrode for CO2 reduction to CO at a TOF of
0.8 s− 1 under the overpotential of 1490 mV in anhydrous acetonitrile.
Click chemistry may be further explored for facile attachment of tetra­
pyrrole macrocycle complexes to carbon supports.
Another strategy of covalent immobilization employing substitution
reaction at the metal center was reported by Zhu et al (Entry 7, Table 4)
Fig. 9. Common methods for introducing functional groups on carbon supports
(a) chemical oxidation of surface; (b) reduction of in-situ generated diazonium [50]. The attachment was realized by refluxing protoporphyrin IX cobalt
salts; (c) 1,3-dipolar cycloaddition of azomethine ylides; (d) Hassner-type chloride (CoPPCl) with hydroxyl-functionalized CNTs in ethanol which
addition of azide groups. Adapted with permission from Ref. [19], Copyright resulted in the substitution of Co-bonded Cl by CNTs-bonded O and
2019, American Chemical Society. subsequently the formation of a covalent bond between Co center and O

15
S. Gu et al. Chemical Engineering Journal 470 (2023) 144249

Table 4
Summary of covalently immobilized porphyrins for CO2ERR.
Entry Catalyst Support Structure V vs. NHE FE TOF Product Ref.
− 1
1 CoTPP Glassy carbon − 0.956 V 50% 6.9 s CO [105]

2 CoTPP Glassy carbon − 0.856 V 92% N.A. CO [106]

3 CoTPP Glassy carbon − 0.856 V 63% N.A. CO [107]

1
4 CATCO2H CNTs − 1.06 V 95% 0.05 s− CO [101]

1
5 CoTAP CNTs − 1.09 V ~100% 6.0 s− CO [39]

6 CoPoralkyne Diamond − 1.6 V N.A. 0.8 s− 1


CO [102]

(continued on next page)

16
S. Gu et al. Chemical Engineering Journal 470 (2023) 144249

Table 4 (continued )
Entry Catalyst Support Structure V vs. NHE FE TOF Product Ref.
1
7 CoPPCl CNTs − 1.02 V 98.3% 1.37 s− CO [50]

1
8 Fe2DTPFPP-PO3H2 FTO − 1.2 V 93% 245 s− CO [104]

1
9 CoTPP Carbon cloth − 1.05 V 67% 8.3 s− CO [103]

atom (Fig. 10a). Transmission electron microscope (TEM) image showed FECO of 98.3% and a stable current density of 25.1 mA/cm2 for more
the smooth surface of the hybrid catalyst CoPP@CNTs, implying the than 12 h at the overpotential of 490 mV. Compared with physically
absence of aggregation (Fig. 10b) The corresponding energy dispersive mixed counterpart, CoPP@CNTs led to a tripled TOF, demonstrating a
X-ray spectroscopy (EDX) maps showed that the distributions of C, N, strengthened catalyst-support interaction and improved electron
and Co overlapped and matched the nanotube structures (Fig. 10c). transfer.
CoPP@CNTs turned out to be highly efficient towards CO2ERR with a Apart from carbon-based supports, metal oxides were exploited for

Fig. 10. (a) Preparation of chemically grafted CoPPCl on CNTs (b) TEM images of CoPP@CNTs; (c) STEM image of CoPP@CNTs and the corresponding EDX maps of
C, N, and Co. Adapted with permission from Ref. [50], Copyright 2019, Wiley.

17
S. Gu et al. Chemical Engineering Journal 470 (2023) 144249

covalent immobilization. For example, phosphonic acid was used as a polymerization through CV processes. The potential range for CV should
covalent bond linker for attaching iron porphyrin dimer (Fe2DTPFPP- be determined with care since the polymerized films would be degraded
PO3H2) to the fluorine-doped tin oxide (FTO) surface (Entry 8, Table 4) at excessive potentials which would result in negative effects on the
[104]. The grafted catalyst exhibited a high activity towards CO2 catalytic performance.
reduction to CO with a TOF of 245 s− 1 and a current density of 0.25 mA/ The polymerization technique was attempted via potential scans on
cm2 at − 1.2 V vs. NHE. A larger current density of 3.0 and 2.5 mA/cm2 amino appended ZnTAP, FeTAP and CoTAP for the formation of a
could be achieved by loading a layer of mesoporous SnO2 or TiO2 conductive film on the surface of indium-tin-oxide (ITO) glass (Entry
nanoparticles on FTO, which enabled a higher loading amount of 1–3, Table 5) [110,111]. The catalytic performances of the three poly­
Fe2DTPFPP-PO3H2. In a control experiment, a phosphonic acid-free iron merized porphyrins were evaluated in an ionic liquid of BMImBF4
porphyrin dimer (Fe2DTPFPP) was immobilized on FTO support under saturated with CO2 and the product CO was detected in all cases. While
the same conditions. The Fe2DTPFPP was completely detached from the polymerized CoTPP and FeTPP could catalyze CO2 reduction with a
FTO surface after ethanol washing, demonstrating the importance of decent FECO of 67 and 78% at − 0.6 V vs. NHE, the Zn catalyst only
phosphonic acid as the anchoring group. exhibited a FECO of 15%.
In addition to the above chemical reactions, electrochemical reaction Introducing vinyl ligand to porphyrins has been proved to be effec­
was also utilized for covalent immobilization. One example is the tive in forming polymeric films on supports for CO2ERR. The polymer­
reduction of corresponding diazonium salts at reducing potentials for ization was initiated by the radical anion of vinyl ligands through an
immobilization of CoTPP on carbon cloth (Entry 9, Table 4) [103]. The electro-oxidation process. For example, cobalt protoporphyrin bearing
resulting phenylene group acted as a “molecular wire” for bonding vinyl groups was polymerized on GC support with a concentration of 14
CoTPP and carbon cloth. The covalently immobilized CoTPP catalyzed nmol/cm2 (Entry 4, Table 5) [112]. The resulting film could reduce CO2
CO2 to CO with a TOF of 8.3 s− 1 and FECO of 67% at − 1.05 V vs. NHE, to CO with a FECO of (84 ± 2)% at the potential of − 1.005 V vs NHE.
which was much higher compared with physically adsorbed counterpart A similar electro-oxidation process was performed on carbazole
with a TOF of 4.5 s− 1 and FECO of 44%. The observation validates that a appended iron porphyrin derivative (FeTCbzPP) for polymerization on
conductive link plays a vital role in the electron transfer from supports to GC and ITO supports (Entry 5, Table 5) [113]. Owing to the microporous
porphyrins. structure and good catalytic activity of the FeTCbzPP units, the afforded
Based on a comparison between Table 3 and 4, the covalently electrode possessed a high uptake capacity of CO2. However, the sta­
immobilized porphyrins exhibit an elevated FE at a less negative po­ bility of the electrode was limited since roughly 30% activity was lost
tential in general, confirming covalent immobilization is a better strat­ after 4 CV scans. Characterizations indicated that partial carboxylation
egy. As for the support, carbonaceous materials are superior to metal of the carbon atoms on the porphyrin scaffold might lead to the degra­
oxides as indicated in Entry 8 of Table 4. The carbonaceous materials, dation of the catalyst.
such as CNTs, have a high overpotential for HER, which in turn The polymerization technique was also carried out on CoTPP bearing
contribute to the selectivities of immobilized porphyrins. Another thiophene groups (CoTPP-Th). Polythiophene was formed via CV scans
favourable point brought by covalent immobilization is the uniform which afforded a polymerized porphyrin film on the surface of carbon
distribution of active sites as indicated in Fig. 10c. Covalent immobili­ fiber paper (Entry 6, Table 5) [114]. The CoTPP-based polymer cata­
zation could ease catalyst aggregation issues faced by non-covalent lyzed CO2 reduction to CO with a FECO of 66% and a TOF of 1.6 s− 1 and a
immobilization and facilitate the access of CO2 to the active sites. In a TON of 5.7 × 103 after 1 h at − 1.09 V vs. NHE. Later a 6-h long run
word, covalent immobilization could overcome the limitations of non- electrolysis demonstrated the satisfactory stability of the polymerized
covalent immobilization and maximize the activities of porphyrins to film at a constant current density of − 0.936 mA/cm2 over the course.
achieve efficient CO2ERR. In addition to electrochemical polymerization, chemical pathways
Although there are increasing interests in the use of covalent were employed for porphyrin polymerization on supports. One example
immobilization, the examples of grafted porphyrins for CO2ERR remain involves the polymerization of ethynyl appended cobalt porphyrin (Co-
limited owing to the fact that few reactions are suitable here. The co­ b) with bromo appended cobalt porphyrin (Co-a) via the Sonogashira
valent immobilization requires an initial step of functionalization of the coupling reaction (Entry 7, Table 5) [115]. A polymeric film (CoCoPCP)
porphyrins and derivatization of the sp2 carbon surface, followed by comprising 2D extended structure of porphyrins was formed on the
chemical reactions between the functionalities and carbon surface, thus surface of CNTs (Fig. 11a). The resulting CoCoPCP/CNTs exhibited a
greatly narrowing down the options for grafting. To solve the issue, high activity for CO2 reduction to CO with a FECO of 94% and TOF of 2.4
more studies are needed to explore possible chemical routes for covalent s− 1 at the overpotential of 440 mV. Calculations revealed that the rate-
anchorage of porphyrins in a facile manner. limiting step for the CO2ERR on CoCoPCP/CNTs was the formation of
*COOH intermediates with rather high free energies (Fig. 11b). How­
5.3. Periodic immobilization ever, Gibbs free energies for forming *COOH intermediates involving
dual Co-a and Co-b sites was more than 5 kcal/mol lower than that
Periodic immobilization usually involves the formation of a periodic involving single Co-a or Co-b site. These results indicated that CO2ERR
porphyrin structure on the surface of supports [108]. The periodic was much advantageous when the two types of Co centers were simul­
structure offers a high porosity, a strong connection with the supports taneously employed.
and dense active sites for CO2ERR, thereby bearing excellent catalytic In general, polymerized thin films of porphyrins could increase the
activities and attracting increasing attention nowadays [109]. This kind areal concentration of porphyrin catalysts on working electrodes for
of structure can be derived from polymerization of porphyrins or pro­ CO2ERR. Polymerization also has the merit of affording uniform layers
vided by MOFs and COFs. of porphyrins compared with non-covalent immobilization. In most
cases, polymerization is achieved via electrochemical scanning of the
5.3.1. Polymerization monomeric porphyrin at an appropriate potential range. One factor
Polymerization via electrochemical reactions provides a viable way limiting its wide application is that the optimal potential window for
to immobilize porphyrins on supports via formation of a film of periodic crosslinking the porphyrin units should be determined with care. It was
porphyrins. Since the early trial of polymerized porphyrin for CO2ERR found that porphyrins may undergo degradation at high overpotentials
(Fig. 4a), [143] efforts were made to explore versatile approaches to which would induce negative consequences on the performance of the
polymerize porphyrins in a convenient way. Functional groups such as afforded polymer. Another limiting factor is that reactive functionalities
amino, vinyl, carbazole, pyrrole, thiophene, alkynyl etc. are frequently are necessary for porphyrins to polymerize. However, functionalities
introduced to the peripheral rings of porphyrins for the purpose of suitable for the purpose remain largely unexplored. Efforts should be

18
S. Gu et al. Chemical Engineering Journal 470 (2023) 144249

Table 5
Summary of polymerized porphyrins for CO2ERR.
Entry Catalyst Support Structure V vs. NHE FE TOF Product Ref.
− 1
1 ZnTAP ITO − 0.6 V 15% 0.91 h CO [111]
2 FeTAP ITO − 0.6 V 78% 3.06 h− 1 CO [111]
3 CoTAP ITO − 0.6 V 67% 17.25 h− 1 CO [110]

1
4 Cobalt protoporphyrinIX GC − 1.005 V 84 ± 2% 0.04 s− CO [112]

5 FeTCbzPP GC/ITO N.A. N.A. N.A. CO [113]

1
6 CoTPP-Th Carbon paper − 1.09 V 66% 1.6 s− CO [114]

1
7 CoCoPCP CNTs − 0.975 V 94% 2.4 s− CO [115]

19
S. Gu et al. Chemical Engineering Journal 470 (2023) 144249

Fig. 11. (a) Structure of CoCoPCP; (b) the free energy diagram for the CO2ERR on CoCoPCP/CNTs. Adapted with permission from Ref. [115], Copyright
2021, Elsevier.

made to introduce new functionalities to porphyrins for polymerization. HCO–3, which would then chelate to the cobalt center for reduction
reaction.
5.3.2. MOFs and COFs Carboxyl functionalized cobalt tetraphenylporphyrin (CoTCPP) was
Porphyrin based MOFs and COFs have drawn extensive attention grown on FTO glass deposited with Al thin coating to generate the MOFs
owing to the unique features such as periodic porous structure, well- (Al2(OH)2TCPP-Co) in-situ (Entry 4, Table 6) [45]. The direct growth of
defined topology, controllable structure at the molecular level, MOFs onto support benefits the stability and conductivity during
increased concentration of metalloporphyrins on the supports, and etc CO2ERR while avoiding the use of binders. The MOFs catalyzed CO2
[116]. Compared to polymerized porphyrins, porphyrin based MOFs reduction to CO at a current density of 1 mA/cm2 under − 1.125 V vs.
and COFs exhibit an ordered porous crystalline structure, which allows NHE with a FECO of 76% and TON of 1400 per site (corresponding to a
for the free permeation of ions and CO2 into the interior layer of the TOF of 0.056 s− 1 per site) in 0.5 M KHCO3 electrolyte. Tafel analysis was
catalyst structure. Owing to the above mentioned advantages, the po­ conducted and the calculated slope of 165 mV/dec indicated a mass
tential use of porphyrin based MOFs and COFs has been reported in transport limited process, which might be incurred by the stacked layers
electrochemical energy conversion [117]. And their activities in of MOFs on the support. In-situ spectroelectrochemical spectra
CO2ERR have been widely explored since the pioneering report in 2015 confirmed that the Co(II) was reduced to Co(I), serving as the catalytic
(Fig. 4a). center during catalysis.
By introducing organic ligands onto metalloporphyrin structures, Active species can be integrated with MOFs to form MOF composites
porphyrins can be employed as secondary building units to coordinate to for CO2ERR. A typical example involves the integration of active Ag
the inorganic nodes for the formation of MOFs. For example, carboxyl nanocrystals into the structure of Al-PMOF ([Al2(OH)2(TCPP)]) for the
functionalized iron tetraphenylporphyrin (FeTCPP) linked with hexa- formation of Ag@Al-PMOF, whereas Al-PMOF is composed of Al node
zirconium(IV) nodes to afford MOF-525 which was later immobilized on and tetra(4-carboxyphenyl)porphyrin (TCPP) ligand (Entry 5, Table 6)
FTO supporting electrode for CO2ERR application (Entry 1, Table 6) [120]. The as-prepared Ag@Al-PMOF exhibited a two-time increase in
[16]. The reduction was achieved at − 1.3 V vs. NHE with a current the selectivity towards CO production and a drastic decrease of HER
density of 5.9 mA/cm2 for 3.5 h in 1 M trifluoroethanol solution in DMF. compared to the bare Ag nanocrystals. Another example is based on the
CO and H2 were formed as the products in a similar rate, corresponding implant of metallocene (CoCp2) to the MOFs structure of MOF-545-Co to
to a FECO of 40% and TONCO of 1500 per site (TOFCO of 0.13 s− 1 per afford CoCp2@MOF-545-Co, whereas MOF-545-Co is composed of Zr
site). The access of the CO2 to the catalytic sites was facilitated by the node and cobalt tetra(4-carboxyphenyl)porphyrin (CoTCPP) ligand
well-defined nanoscale porosity of the MOFs structure. (Entry 6, Table 6) [121]. CoCp2@MOF-545-Co presented an excellent
The above strategy was also utilized for the synthesis of PCN-222 performance towards CO2 electroreduction to CO with a FECO of 97% at
(Fe), which consists of Zr6 cluster linked by square planar FeTCPP li­ − 1.125 V vs. NHE (Fig. 12a). The good performance is ascribed to the
gands (Entry 2, Table 6) [118]. The MOFs was dip coated on carbon strong binding-interaction between CoCp2 and CoTCPP that can largely
black support with a mass ration of 1:2 for PCN-222(Fe)/C. The com­ reduce the adsorption energy of CO2 from − 0.24 eV to − 0.92 eV as
posite catalyst catalyzed CO2 reduction to CO with a FECO of 91% and a indicated by DFT calculations (Fig. 12b).
TOFCO of 0.336 s− 1 per site at the current density of 1.2 mA/cm2 under In addition to MOFs, metal–organic layers (MOLs) and porous
494 mV overpotential in 0.5 M KHCO3 electrolyte. The study suggests organic cage were developed for CO2ERR. A TPY-MOL was synthesized
that a combination of the fine activity of the porphyrin ligands, high by assembling terpyridine-based tricarboxylate ligands (TPY) and
adsorption of the porous structure and good conductivity of the carbon [Hf6(µ3-O)4(µ3-OH)4(HCO2)6] nodes, which was later functionalized
black plays a key role in enhancing the performance towards CO2ERR. with cobalt-protoporphyrin (CoPP) via carboxylate exchange with
Porphyrin based TIPP ([5,10,15,20-tetra(4-imidazol-1-yl)phenyl] formate capping groups (Entry 7, Table 6) [122]. The CoPP center and
porphyrin) linkers combined with CoII ion nodes were used to construct the pyridine moiety on the TPY-MOL layer cooperatively adsorbed and
stable MOFs of NNU-15 with two OH– ions (Entry 3, Table 6) [119]. activated the CO2 molecule by forming the [pyH+–O2C-CoPP] adducts.
NNU-15 was then dropcasted on carbon paper for electrochemical CO2 As a result, TPY-MOL-CoPP demonstrated an enhanced electrocatalytic
reduction to CO. The immobilized NNU-15 exhibited a maximum FECO activity towards CO2 reduction to CO, with a high CO/H2 selectivity of
of 99.2% and at − 1.025 V vs. NHE and an overall FECO over 96% in the 11.8, a TOFCO of 0.4 s− 1 and a current density of 1314 mA/mg Co at
potential range from − 1.025 to − 1.325 V in 0.5 M KHCO3. The inter­ − 1.261 V vs. NHE. The work highlights the potential of MOLs as a
mediate NNU-15-CO2 was observed, demonstrating the cobalt center of substrate for immobilizing catalytic sites in the aim to improve activities
NNU-15 could cooperate with OH– to activate the CO2 molecules to form for CO2ERR.

20
S. Gu et al. Chemical Engineering Journal 470 (2023) 144249

Table 6
Summary of porphyrin-based MOFs for CO2ERR.
Entry MOFs Structure Support V vs. NHE FE TOF Product Ref.
− 1
1 MOF-525 FTO − 1.3 V 40% 0.13 s CO [16]

1
2 PCN-222 Carbon black − 1.037 V 91% 0.336 s− CO [118]

1
3 NNU-15 Carbon paper − 1.025 V 99.2% 0.0104 s− CO [119]

1
4 Al2(OH)2TCPP-Co Carbon disk − 1.125 V 76% 0.056 s− CO [45]

1
5 Ag@Al-PMOF Glassy carbon − 1.507 V 56% 0.0136 s− CO [120]

1
6 CoCp2@MOF-545-Co Carbon paper − 1.125 V 97% 0.216 s− CO [121]

1
7 TPY-MOL Carbon cloth − 1.261 V 92% 0.4 s− CO [122]

(continued on next page)

21
S. Gu et al. Chemical Engineering Journal 470 (2023) 144249

Table 6 (continued )
Entry MOFs Structure Support V vs. NHE FE TOF Product Ref.
1
8 POC Glassy carbon − 1.061 V 85% 1.74 s− CO [123]

Fig. 12. (a) FE for CO and H2 on different catalysts; (b) The free energy diagrams of CO2 reduction to CO for CoTCPP (red line) and CoCp2 decorated CoTCPP (blue
line). Adapted with permission from Ref. [121], Copyright 2020, Elsevier. (For interpretation of the references to colour in this figure legend, the reader is referred to
the web version of this article.)

A porous organic cage (POC) composed of six FeTPP was formed by Following the above pioneering research, a series of COF-366-Co
condensing six tetraformylphenylporphyrins and eight triamine linkers derivatives with different electron-withdrawing groups on their
followed by metalation with iron (Entry 8, Table 6) [123]. The supra­ respective struts was synthesized via introducing fluorine and methoxyl
molecular catalyst was loaded onto CNTs-coated GC support for the to the subunit BDA (Entry 3–5, Table 7) [40]. The derivatives were
purpose of CO2 reduction to CO. The catalyst exhibited a FECO of 85% termed as COF-366-Co, COF-366-(OMe)2-Co, COF-366-(F)4-Co, COF-
and TON of 55,250 at − 1.061 V vs. NHE after 24 h electrocatalysis, 366-F-Co and the electron-withdrawing effect of the groups on the co­
which were greatly enhanced compared with the parent FeTPP. The balt site was found to increase in the same order. The current densities of
elevated activity was ascribed to the porous structure of the supramo­ the COFs derivatives towards CO2 reduction to CO at − 1.095 V vs. NHE
lecular network, which allowed an easy access of CO2 to the active increased from 45 mA/mg (COF-366-Co) to 46 mA/mg [COF-366-
center. (OMe)2-Co] and to 65 mA/mg (COF-366–F-Co), which confirmed the
Apart from MOFs, COFs are also frequently employed for CO2ERR improvement of the catalytic behaviour provided by electron-
application owing to the following benefits such as diverse and tunable withdrawing effect. Despite being the second most electron-
structures, high porosity, accessible active sites, and excellent CO2 withdrawing catalyst, COF-366-(F)4-Co does not fit in the trend. The
adsorption capacity. Compared to MOFs, COFs have better stabilities reason is due to the higher hydrophobicity of the framework brought by
under extreme electrocatalytic conditions since robust covalent bonds the four fluorines which resulted in confined access of CO2 to the cobalt
are formed in the structures. sites.
An initial attempt on porphyrin based COFs for CO2ERR was re­ Based on COF-366-Co, tetrathiafulvalene (TTF) strut was integrated
ported by Yaghi and Chang via imine condensation of 5,10,15,20-tetra­ into the structure to afford TTF-Por(Co)-COF which could enhance the
kis(4-aminophenyl)porphinate cobalt (CoTAP) with 1,4- electron transfer from the TTF to the porphyrin ring (Entry 6, Table 7)
benzenedicarboxaldehyde (BDA) (Entry 1, Table 7) [44]. As CoTAP [124]. TTF-Por(Co)-COF could catalyze CO2 reduction to CO with a FECO
was embedded into the skeleton, effective π-π stacking could facilitate of 95% at − 1.148 V vs. NHE, which is comparatively higher than COF-
charge carrier mobility around the catalyst. The as-prepared COF-366- 366-Co. DFT calculations revealed that the energy required for the
Co could catalyze CO2 reduction to CO with a FECO of 90% and a TONCO limiting step in CO2 reduction for TTF-Por(Co)-COF is lower with respect
of ~ 34000 over 24 h at − 1.095 V vs. NHE. The activity could be further to that of COF-366-Co.
enhanced by constructing a more porous structure of COFs which would More recently, the imine condensation between the amino group of
allow the feedstock CO2 to access the catalytic sites more easily. To the porphyrin and aldehyde group of the organic linkers was employed
achieve the goal, biphenyl-4,4′-dicarboxaldehyde (BPDA) was used as for constructing FeDhaTph-COF which comprised 5,10,15,20-tetra-(4-
the strut in place of BDA for synthesis of COF-367-Co (Entry 2, Table 7). aminophenyl)porphyrin Fe(III) chloride (FeTAPPCl) and 2,5-dihydroxy­
The resulting catalyst could catalyze CO2 reduction to CO with an terephthalaldehyde (Dha) (Entry 7, Table 7) [125]. The synthesis of
improved FECO of 91% and a TON of 48,000 over 24 h at − 1.095 V vs. FeDhaTph-COF was conducted on carbon cloth in-situ and the formation
NHE. of imine bonds was confirmed by XPS. The catalyst could reduce CO2 to

22
S. Gu et al. Chemical Engineering Journal 470 (2023) 144249

Table 7
Summary of porphyrin-based COFs for CO2ERR.
Entry COFs Structure Support V vs. NHE FE TOF Product Ref.
− 1
1 COF-366-Co Carbon fabric − 1.095 V 90% 0.39 s CO [44]

1
2 COF-367-Co Carbon fabric − 1.095 V 91% 0.56 s− CO [44]

3 COF-366-(OMe)2-Co Carbon fabric − 1.095 V N.A. N.A. CO [40]


4 COF-366-F-Co N.A. N.A. CO
5 COF-366–(F)4-Co N.A. N.A. CO

1
6 TTF-Por(Co)-COF Carbon fiber paper − 1.148 V 95% 0.12 s− CO [124]

1
7 FeDhaTph-COF Carbon cloth − 2.0 V 80% 0.17 s− CO [125]

1
8 TPPDA-CoPor-COFs Carbon fiber paper − 1.025 to − 1.325 V 87–90% 1.4–2.0 s− CO [126]
1
9 TPPDA-NiPor-COFs Carbon fiber paper − 1.125 to − 1.325 V 60–76% 0.2–0.4 s− CO

1
10 MWCNT-Por-COF-Co CNTs − 1.025 V 99.3% 15 s− CO [127]
11 MWCNT-Por-COF-Cu CNTs − 1.826 V 71.2% N.A. CH4

CO with a FECO of 80% and TOF of over 600 h− 1 per mole of electro­ (Entry 8, Table 7) [126]. The as-prepared catalyst showed a good ac­
active Fe sites at − 2.0 V vs. NHE over 3 h in MeCN with 0.5 M tri­ tivity towards CO2 reduction to CO with a FECO of 87–90% at − 1.025 to
fluoroethanol. XPS and ICP-OES confirmed that the FeTAPPCl molecules − 1.325 V vs. NHE and a high current density of –22.2 mA/cm2 at
in the COFs structure were retained after electrocatalysis. − 1.425 V vs. NHE. While for TPPDA-NiPor-COFs (Entry 9, Table 7)
Another material TPPDA-CoPor-COFs was obtained by the imine where the cobalt center was replaced by nickel, the activity was
condensation of tetraphenyl-p-phenylenediamine (TPPDA) and significantly lower with FECO of 60–76% at − 1.125 to − 1.325 V vs. NHE
5,10,15,20-tetrakis(4-formylphenyl)-porphyrinate cobalt (CoPor) and a current density of − 3.9 mA/cm2 at − 1.425 V vs. NHE. The

23
S. Gu et al. Chemical Engineering Journal 470 (2023) 144249

superior activity of Co site than Ni site for CO2ERR was explained by spectra of Cu center showed the typical peak of Cu0 at 567.8 eV, instead
DFT calculations, which confirmed the desorption of CO from TPPDA- the expected Cu+ peak was not visible after CO2ERR (Fig. 13e). The
NiPor-COF was endothermic, while this step for TPPDA-CoPor-COF combined results indicated that Cu0 species (i.e. Cu nanoclusters) were
was exothermic. generated during electrocatalysis and acted as the active sites for CO2
The above examples immobilize COFs on the support via non- reduction to CH4.
covalent bonds. Covalent bonds were employed for anchoring COFs on The utilization of MOFs and COFs can significantly increase the
CNTs via in-situ synthesis of benzaldehydes appended COFs on –NH2 catalytically active sites and hence the current density. Compared with
modified CNTs. The in-situ synthesis utilizes condensation between densely packed polymerized films, MOFs and COFs exhibit an ordered
5,10,15,20-Tetrakis(4-benzaldhyde)porphyrin (Por-CHO) and p-phe­ and porous structure, which enables the free diffuse of electrolyte
nylenediamine for the preparation of MWCNT-Por-COF-M (M = Co, Ni, counter ions and feedstock CO2 into the interior of the frameworks for
Fe, Cu) (Entry 10–11, Table 7) [127]. Compared with Fe or Ni center, access of the active sites. A further enhancement can be seen in Entry 10
MWCNT-Por-COF-Co catalyzed CO2 reduction to CO with the highest of Table 7, where the FE of COFs was promoted to 99.3% after covalent
FECO of 99.3% at − 1.025 V vs. NHE, as well as a high current density of immobilization. Hence the importance of a close and robust interaction
18.77 mA/cm2 and a TOFCO of 70.6 s− 1 at − 1.425 V vs. NHE in an H-cell between COFs and supports was again emphasized. Due to the trou­
filled with 0.5 M KHCO3 (Fig. 13a). While MWCNT-Por-COF-Cu (Entry blesome and limited synthetic options, most studies still immobilize
11, Table 7) exhibited the highest activity towards CO2 reduction to CH4 COFs/MOFs via non-covalent bonds. However, the issue of catalyst
with a FECH4 of 71.2% at − 1.826 V vs. NHE in a flow cell filled with 1.0 stacking would take place as discussed before. A decrease in the catalytic
M KOH (Fig. 13b). Mechanism study revealed that owing to the covalent performance can be envisioned at higher loadings due to the mass
bond between MWCNT and Por-COF-M (Co, Ni, Fe), Por-COF-M was transport limitation from the electrolyte to MOFs or COFs as the film
vertically integrated on MWCNT and electrons could continuously thickness increases. To overcome the limitations, the structure of MOFs
transfer from MWCNT to porphyrin plane of Por-COF-M (Fig. 13c). and COFs could be tuned by modifications of the periodic structure with
Subsequently, the electrons could be transmitted to active metal sites for pendant functional groups. The functional groups may have effects on
CO2 reduction via the conjugated structure in Por-COF-M. The covalent the porosity of the frameworks and thus improve mass transport along
combination mode thus led to enhanced performance. While for the backbones of MOFs and COFs. In addition, COFs and MOFs can be
MWCNT-Por-COF-Cu, HRTEM revealed that ultra-small particles (ca. 2 designed in the form of a two-dimensional (2D) thin nanosheet. It would
nm) with the lattice fringe of 0.21 nm, corresponding to the (1 1 1) be more feasible compared to the above functionalization which would
planes of Cu0, were formed after CO2ERR (Fig. 13d). In addition, Auger require troublesome chemical synthesis. The thin films would allow an

Fig. 13. (a) The current density of CO from − 0.5 to − 1.0 V; (b) Product distribution of MWCNT-Por-COF-Cu from − 0.7 to − 1.2 V; (c) Proposed mechanism for the
CO2ERR on MWCNT-Por-COF-M (Co, Ni, Fe); (d) HRTEM of MWCNT-Por-COF-Cu after CO2ERR; (e) Auger spectrum of MWCNT-Por-COF-Cu after CO2ERR. Adapted
with permission from Ref. [127], Copyright 2022, Elsevier.

24
S. Gu et al. Chemical Engineering Journal 470 (2023) 144249

easy access of CO2 molecules to the active centers and thus facilitate nanosheets (iCONs). For example, the amine groups in nanosheets
mass transport. CoTAP-CONs were methylated to obtain CoTAP-iCONs with quaternary
ammonium groups, which offered steric hindrance and electrostatic
5.3.3. Ultrathin nanosheets repulsion that could effectively reduce the thickness of CoTAP-iCONs
To overcome the limitations of COFs/MOFs with stacked layers, ul­ [128]. The performance reached a high selectivity of 96.74% for CO
trathin 2D nanosheets provides an effective strategy to facilitate the and a large current density of 11.92 mA/cm2 at − 1.08 V vs. NHE (Entry
access of CO2 feedstock to the active centers. The ultrathin structure 1, Table 8). Tafel slope of CoTAP-iCONs (122 mV/dec) was close to the
endows porphyrins with attractive properties, including high surface theoretical value of an electron-transfer-rate-determining process as
area, exposed electrocatalytic active sites, rapid mass transport, discussed in section 3.1, which indicated the mass transfer process was
enhanced electron transfer and thus excellent performance. rapid.
One of the synthetic methods for ultrathin nanosheets is to incor­ Via axial coordination between the assembly molecules tetra(4-
porate charged units into the skeleton of ionic-covalent organic pyridyl) porphyrin cobalt (II), ultrathin nanosheets STPyP-Co were

Table 8
Summary of porphyrin-based ultrathin nanosheets for CO2ERR.
Entry Catalyst Structure Support V vs. NHE FE TOF Product Ref.
1
1 CoTAP-iCONs Carbon fiber − 1.08 V 96.74% 0.7 s− CO [128]
paper

1
2 STPyP-Co Carbon fiber − 1.05 V 96% 4.21 s− CO [43]
paper

3 PML-Cu Carbon fiber − 1.11 V HCOOH (80.86%); CH4 N.A. HCOOH, CH4 [129]
paper (11.51%)
4 PML-Au Carbon fiber − 1.21 V HCOOH (40.90%); CO N.A. HCOOH, CO
paper (34.40%)

1
5 COF-366-(OMe)2- CNTs − 1.105 V 93.6% 3.30 s− CO [130]
Co@CNT
− 1
6 COF-366-(OH)2- CNTs 90% 2.78 s CO
Co@CNT
− 1
7 COF-366-(F)4- CNTs 91% 2.29 s CO
Co@CNT

8 Cu-Tph-COF-Dct Carbon fier − 1.73 V 80% N.A. CH4 [46]


paper

9 Cu2(CuTCPP) FTO − 1.55 V vs. HCOOH (68.4%); 0.566 s− 1; HCOOH, [131]


Ag/Ag+ CH3COOH (16.8%) 0.041 s− 1 CH3COOH

25
S. Gu et al. Chemical Engineering Journal 470 (2023) 144249

formed (Fig. 14d) [43]. The ultrathin 2D shape of STPyP-Co was evi­ catalyst exhibited a superior selectivity of 80% towards CH4 and a high
denced by Fig. 14a, b as a quadrilateral. The elements of Co, C, N were current density of 220 mA/cm2 at − 1.73 V vs. NHE. The enhanced ac­
uniformly distributed on the whole nanosheets as displayed in Fig. 14c. tivity was ascribed to the functionalizing exfoliation agent, which could
The activity of the nanosheets was significantly promoted via elevating afford ultrathin nanosheets with exposed active sites, strengthen the
the energy level of Co d2z orbitals, exhibiting a high selectivity of 96% for absorption/activation of CO2, enrich the CO concentration around Cu
CO and a TOF of 4.21 s− 1 (Entry 2, Table 8). Co center with raised d2z active sites, and thus facilitate the formation of CH4.
orbital was more easily to be reduced by fulfilling the d2z orbital, The design of ultrathin nanosheets has also been made on MOFs. A
resulting in an enhanced process of electron transfer from Co to CO2. 2D ultrathin MOFs termed as Cu2(CuTCPP) was constructed via the
Using a facile solvothermal method, porphyrin-based monoatomic connection of Cu porphyrin ligands through Cu2(COO)4 paddlewheels
layer (PML) was easily obtained with Cu or Au as the metal center [131]. The thickness of the nanosheet was measured to be ~3.7 nm,
(denoted as PML-Cu and PML-Au) [129]. PML-Cu performed a high FE equalling to 8 layers. The performance of Cu2(CuTCPP) was evaluated in
of 80.86% for HCOOH and 11.51% for CH4 at − 1.11 V vs. NHE, while CH3CN with 1 M H2O and 0.5 M ionic liquid. At the potential of -1.55
PML-Au generated HCOOH and CO at the FE of 40.90 % and 34.40 % at V vs. Ag/Ag+, the FE for formate and acetate was 68.4% and 16.8%, and
− 1.21 V vs. NHE (Entry 3&4, Table 8). The above results suggested that the corresponding TOF was 0.566 s− 1 and 0.041 s− 1, respectively. The
selectivity could be easily tuned when the metal centers were accurately high activity was ascribed to the cathodized reconstruction of
tailored. Compared with MOF-Cu catalyst which exhibited a typical Cu2(CuTCPP) as well as the ultrathin feature of 2D MOF nanosheets.
nanocube structure with a size/thickness of about 500 nm, the thickness The periodic structure such as MOFs and COFs, once designed as an
of PML-Cu was measured to be 0.28 nm. The monoatomic layered ultrathin nanosheet, could greatly reduce the mass transfer resistance.
structure minimizes the mass transfer resistance and boosts the elec­ The ease of mass transfer resistance was indicated by lowered Tafel
trochemical activity. slopes in the above reports [128,43,130]. For example, decreased Tafel
In-situ synthesis provides an alternative approach to afford ultrathin slopes were observed for CoTAP based materials when the thickness
nanosheets. COF-366-Co derivatives with fluorine, methoxyl and hy­ reduced, and a Tafel slope close to the theoretical value of 118 mV/dec
droxyl groups were generated in-situ on amino functionalized CNTs was achieved on ultrathin nanosheets, indicating a facilitated mass
(Entry 5–7, Table 8) [130]. The thickness of the nanosheet was only ca. transfer through the ultrathin layer [128]. As such, thick layers would
0.9 nm, corresponding to the three layers of the 2D COFs. The catalyst have a high Tafel slope due to poor mass transfer across the layers.
containing methoxyl functional groups (COF-366-(OMe)2-Co@CNT) Indeed, the above conclusion was validated and supported by Tafel
exhibited the highest activity towards CO2 reduction to CO with a FECO slopes reported for COFs and MOFs. Owing to the considerable thickness
of 93.6% and a current density of 40 mA/cm2 at the potential of − 1.105 of MOFs and COFs, high Tafel values are often encountered [44]. The
V vs. NHE. The Tafel slope displayed 190 mV/dec, which was much findings from the two types of porphyrins are correlated, highlighting
smaller than the previous report using a similar catalyst of COF-366-Co the importance of reducing the layer thickness.
with multilayers (470 mV/dec) [44]. COF-366-(OH)2-Co@CNT had a Concluded from the above-mentioned non-covalent, covalent, peri­
decreased activity owing to the weaker electron-withdrawing effect odic immobilization and etc., ultrathin nanosheets present a great po­
provided by hydroxyls. While fluorine functionalized COF-366-(F)4- tential in CO2ERR. The ultrathin nanosheets not only provide abundant
Co@CNT exhibited the lowest TOFCO due to the hydrophobicity of active sites, but also reduce the mass transport resistance. More dedi­
fluorines which greatly restricted the access of electrolyte and CO2 to the cated work is needed to explore the suitable materials for ultrathin
active sites. nanosheets at this stage. If the ultrathin nanosheets could be immobi­
Functionalizing exfoliation agent that could simultaneously modify lized on a gas diffusion electrode, preferrable via a covalent bond, the
and exfoliate bulk COFs has been attempted to construct functional assembly electrode is expected to show excellent performance in a flow
nanosheets. The functionalizing exfoliation agent 2,4-diamino-6-cholo- cell setup (to be discussed in the next part), which would bring a more
1,3,5-triazine (Dct) connected with CuTAP in the bulk COFs via C-O promising future for the commercialization of CO2ERR.
covalent bonds, affording the exfoliated nanosheet Cu-Tph-COF-Dct
[46]. The thickness of Cu-Tph-COF-Dct was only ~3.8 nm. The

Fig. 14. (a) AFM (b) TEM (c) High-angle annular dark-field (HAADF) and EDX mapping images of STPyP-Co; (d) Scheme of the axial interaction between pyridine
and central Co. Adapted with permission from Ref. [43], Copyright 2019, Wiley.

26
S. Gu et al. Chemical Engineering Journal 470 (2023) 144249

6. CO2ERR in flow cells (Entry 2, Table 9), which was more than doubled compared with that in
an H-type cell.
The above discussion of CO2ERR, be it homogeneous or heteroge­ An even more pronounced activity towards CO2 reduction to CH4
neous catalysis, is based on batch conditions where an H-type cell is used was achieved on ultrathin nanosheets comprising diaminotriazine
and CO2 is pre-purged in the cell. Compared with an H-type cell, a flow functionalized Cu porphyrin (CuTDPP) [136]. Via a facile solvothermal
cell can overcome the inherent mass-transport limitations by circulating method, a porous crystalline material was formed through hydrogen
CO2 to and away from the electrodes, leading to a higher concentration bonding between CuTDPP, which was later exfoliated to afford an ul­
of CO2 to reach the catalyst surface and larger geometric current den­ trathin (≈5nm) nanosheet (CuTDPP-NS). The activity of CuTDPP-NS
sities. More importantly, a flow cell satisfies the desire of continuous reached a FECH4 of 70% at − 2.01 V vs. NHE in a flow cell (Entry 3,
production in industry. As illustrated in Fig. 15a, a flow cell consists of a Table 9). The performance of CuTDPP-NS was superior to most of the
sandwich of mounted flow frames, gas diffusion electrodes (GDE), gas­ work on electrochemical CO2-to-CH4 reduction [136]. The reduced mass
kets and an ion exchange membrane [132]. CO2 is allowed to flow transfer resistance provided by ultrathin nanosheets, together with the
through the GDE, while the catholyte solution is circulated in between hydrogen bonding networks in the structure which could stabilize the
the GDE and the ion exchange membrane. Within the GDE, CO2ERR intermediate and transfer the proton, resulted in a facilitated binding of
takes place. H+ and *CO to produce CH4.
The report of porphyrin utilized in a GDE setup for CO2ERR appeared In addition to drop-casting, electropolymerization was used for
in the 90 s [24]. The GDE is made of carbon black, consisting of a hy­ immobilizing Co protoporphyrin (Co-PPIX) on GDE [134]. The forma­
drophilic reaction region and a hydrophobic gas diffusion region, and tion of the catalyst polymer on GDE was illustrated in Fig. 15d via the
had a very large reaction area. The performance of Co porphyrin loaded electrochemical oxidation of the two vinyl side groups to form saturated
GDE achieved a FE of nearly 100% for CO production under 20 atm. links with other PPIX molecules or carbon fiber paper support. An in­
More research has been conducted on GDE setup recently. For crease in the value of Electrochemical Active Surface Area on GDE was
example, Fe porphyrin bearing trimethyl ammonium groups (FeP) and observed after polymerization compared with spray-coating. The
carbon black were drop-casted on carbon fiber paper to manufacture modified GDE exhibited an elevated FE of 94% at − 1.80 V vs. NHE for
GDE [133]. The diffusion layer on carbon fiber paper could be clearly CO production (Entry 4, Table 9), with a much higher current density of
observed in Fig. 15b, which indicated the formation of GDE. The catalyst 33 mA/cm2.
FeP was evenly distributed on the surface of GDE as confirmed by EDX To compare the performance in a flow cell and H-type cell, Fe-
mapping in Fig. 15c. When implemented in a flow cell, this catalytic TPPNH2 was integrated on GDE and glassy carbon for each purpose
material exhibited a selectivity for CO of 98.1% and a current density of [137]. A TOF of 14 s− 1 and a FE of 87% for CO production were attained
155 mA/cm2 at the overpotential of 470 mV (Entry 1, Table 9). The at − 1.83 V vs. NHE with an extraordinary current density of 119 mA/
exceptional activity even surpassed that of silver-based nanomaterials. cm2 (Entry 5, Table 9). This remarkable current density is 12 times
Amino functionalized Cu porphyrin (CuTAP), which could reduce higher than that of the H-cell system (10 mA/cm2) at the same over­
CO2 to hydrocarbons, was also employed in a flow cell [135]. The ink of potential, indicating the product CO was generated at a fast rate.
CuTAP was drop-casted on carbon fiber paper for the preparation of The better performances validate the feasibility of flow cells for in­
GDE. At − 2.46 V vs. NHE, FE of 54.8% for CH4 production was achieved dustrial applications. The feasibility of flow cells has been further
and a remarkable partial current density of 290.5 mA/cm2 was recorded confirmed by a recent study which discovers the syngas production with

Fig. 15. (a) Graphic illustration of a typical flow cell; (b) Profile image of the carbon fiber paper and the diffusion layer of GDE; (c) SEM image and EDX elemental
mapping of the top view of GDE; (d) Schematic illustration of electropolymerization on the electrode surface. Adapted with permission from Ref. [132], Copyright
2020, American Chemical Society; Ref. [133], Copyright 2020, Wiley; Ref. [134], Copyright 2020, Royal Society of Chemistry.

27
S. Gu et al. Chemical Engineering Journal 470 (2023) 144249

Table 9
Summary of porphyrins in flow cells for CO2ERR.
Entry Catalyst Structure Support V vs. NHE FE TOF Product Ref.

1 FeP GDE − 1.42 V 98.1% N.A. CO [133]

1
2 CuTAP GDE − 2.46 V 54.8% 0.28 s− CH4 [135]

3 CuTDPP-NS GDE − 2.01 V 70% N.A. CH4 [136]

1
4 Co-PPIX GDE − 1.80 V 94% 1.01 s− CO [134]

1
5 Fe-TPPNH2 GDE − 1.83 V 87% 14 s− CO [137]

6 CoTPP GDE − 1.01 V CO (45%); H2 (55%) CO (0.01 s− 1) CO, H2 [138]


7 − 1.41 V 98% 0.13 s− 1 CO

a tunable H2/CO ratio could be achieved on flow cells [138]. The study technoeconomic potential of the CO2ERR. More attention is needed to
revealed that GDE comprising CoTPP/CNTs in a flow cell could afford a fill the gap in the design of electrolyzers, so as to improve the energy
ratio of 1.2 for H2/CO at − 1.01 V vs. NHE under 40 ◦ C, which was efficiencies of the electrochemical CO2-recycling systems.
essential for the synthesis of oxo-alcohols. While at − 1.41 V vs. NHE ­
under 20 ◦ C, the composite showed a high selectivity of 98% toward CO 7. Conclusions and outlook
(Entry 6&7, Table 9). This finding is of great importance for industry
where different H2/CO ratios are required to produce various chemicals As an important class of electrocatalysts, porphyrins with various
and fuels. metal centers and ligands have demonstrated excellent performances in
Currently, selective, robust, and fast electrochemical reduction of catalyzing CO2 to reduced forms of carbon such as CO in particular. The
CO2 remains a great challenge in the scientific field as well as in the study of porphyrins with first row transition metals for CO2ERR has
industrial environment. Flow cell setup provides a new approach to cope experienced rapid development throughout the years. With a focus on
with the challenge as it could operate at industrial scales while fulfilling structure-performance relationship, we have provided a thorough re­
customer needs and requirements. Based on the above reports, we can view of porphyrin catalysts employed in CO2ERR, covering mechanism/
conclude that molecular catalysts such as porphyrins can be used in flow kinetic study, homogeneous catalysis and heterogeneous catalysis.
cell devices and are candidates for being included in large scale CO2 With well-defined structure and active metal center, porphyrins are
electrolyzers. However, the configuration of the electrolyzer has not inherently suitable for mechanistic study by in-situ spectroscopy as
been given as much attention as the catalysts at present. Most of the discussed in the review. The review also provides insights into the
research focused on the design and synthesis of new catalysts or func­ dependence of product distribution on the type of metalloporphyrins as
tionalization of the structures, to some extent underestimating the well as the inductive effects of peripheral ligands on the activities of
important role of electrolyzer on the overall performance and metalloporphyrins. For porphyrins operating in homogeneous

28
S. Gu et al. Chemical Engineering Journal 470 (2023) 144249

conditions, the review has presented examples of iron porphyrins and attention has been put on the design of catalysts to boost the activity and
discussed the effects of Lewis acid and Brønsted acid. The studies reveal minimize the overpotential at present. However, the configuration of
that porphyrins could be suitable for achieving a selective CO2ERR electrolyzer, which would greatly impact the overpotentials of both
under appropriate conditions. Compared with homogeneous catalysis, CO2ERR on the cathode and OER on the anode, has not received equal
heterogeneous catalysis permits recycling and in-depth mechanism attention as the design of catalysts has. A significant reduction in the
study by a series of techniques and overcomes the diffusional electron overpotential of the CO2RR and OER in an electrolyzer is critical to the
transfer restrictions by an improved kinetic pathway. The selection of commercialization of CO2ERR, which underlines the importance of the
immobilization techniques is critical to achieving the desired properties optimization work on electrolyzer design. Thus, more dedicated work
and performances in the resulting hybrid catalyst. Herein, the common focusing on the issues with applications is needed for scale up trials.
techniques for immobilization such as non-covalent and covalent are In terms of future development, there are still great challenges
presented. In comparison with non-covalent immobilization, the for­ associated with the performances of porphyrin catalysts and its appli­
mation of covalent bonds affords catalysts with better activity and sta­ cation in industry. The most urgent one is that the durability of por­
bility owing to the stronger connection between porphyrins and phyrins under negative potentials does not meet the requirements of
supports. In addition to the above-mentioned immobilization tech­ industrial applications. An in-depth mechanistic study on the stability of
niques, alternative approaches for porphyrin confinement have been porphyrins is needed to unravel the key factors that have influences on
discussed, including the encapsulation into polymeric matrices, the stability. Another issue needs to be answered is catalyst poisoning.
porphyrin-based COFs and MOFs, polymerized porphyrins, as well as The major product catalyzed by porphyrins is CO, which is recognized as
ultrathin nanosheets. Compared to the molecular porphyrins, incorpo­ a principal contributor to catalyst poisoning in a range of industrial
rating porphyrins into the frameworks or scaffolds could afford periodic processes [142]. The interactions between CO and the active metal
structures with more active sites for accessing the CO2 feedstocks. But centers need to be studied over time to reveal the possibility of catalyst
the issue of catalyst stacking may arise due to the multilayers of the poisoning. In addition, although porphyrins have a high selectivity to­
frameworks. An exquisite design of the frameworks to afford ultrathin wards CO, its concentration in the gas phase is far from the standards of
nanosheets would reduce the mass transfer resistance between the layers a profitable industrial product. Thus, product separation and purifica­
of frameworks and hence increase the activity further. tion need to be developed to couple with CO2ERR technique for a
The above strategies employed in porphyrin catalyst design could be complete industrial process. Last but not least, the selectivity should be
applied to other types of molecular catalysts such as phthalocyanines. finely tuned via a careful screen of the metal center and an elaborate
Owing to the similar structures between porphyrins and phthalocya­ design of the porphyrin structure in order to afford products with high
nines, introduction of pendant groups on porphyrins for tuning their added value. We believe that future cooperation in experimental in­
stabilities and activities is found to be effective on phthalocyanines in a vestigations and scale-up tests will provide a logical approach to
similar way. For example, methoxyl groups were found to improve the advance the development of the technique further.
stability of CoTPP as the bulky donating functional groups disfavour the To conclude, the above summaries based on porphyrin catalysts are
formation of dianionic species and thus protect the vulnerable meso- instrumental to the rational design and development of new efficient
position of CoTPP [37]. The improvement in stability was also observed catalysts, which could further enhance the activity and stability of the
for nickel phthalocyanines after introducing methoxyl groups [139]. In catalytic process. Despite the great challenges in the commercialization
addition, covalent immobilization was found to improve the activities of of CO2ERR, the potential to scale up the technique in the coming decades
both porphyrins and phthalocyanines as compared to physically mixed could be envisioned.
counterparts [103,140]. Thus, an in depth understanding of the strate­
gies for enhancing porphyrin activities could be well extended to other
Declaration of Competing Interest
types of molecular catalysts, which again necessitates an overall review
of porphyrin catalysts.
The authors declare that they have no known competing financial
On the other hand, limitations of porphyrins are clear. As indicated
interests or personal relationships that could have appeared to influence
in Tables 1–7, the majority of porphyrin catalysts afford CO as the
the work reported in this paper.
product, while further reduction to hydrocarbons and alcohols are rare.
This is due to the adsorption of radical CO species on the metal center
Data availability
such as cobalt, nickel or iron is relatively weak, which would lead to the
release of CO as the final product. Although CO could be utilized as the
No data was used for the research described in the article.
feedstock to produce fuels and diesels via Fisher-Tropsch synthesis, it is
not a ready-to-use chemical as methane or methanol due to the toxicity.
Acknowledgements
Thus, reduction of CO2 to hydrocarbons or alcohols is a more desirable
process, while this is rarely realized on porphyrin catalysts. Recently,
The authors are grateful for the support by the International Joint
copper was found to be a viable platform for reducing CO2 to hydro­
Lab of Jiangsu Education Department.
carbons. With deeper insights into the structure–activity-selectivity
relationship, porphyrins with copper center might be a candidate for
achieving the above goals in the future. References
In addition, the progress and applications of CO2ERR are still
[1] H. Shin, K.U. Hansen, F. Jiao, Techno-economic assessment of low-temperature
confined to lab scale despite extensive research has been conducted on carbon dioxide electrolysis, Nat. Sustainability 4 (2021) 911–919.
CO2ERR recently. The issues such as catalyst stability, electrolyzer [2] J. Qiao, Y. Liu, F. Hong, J. Zhang, A review of catalysts for the electroreduction of
carbon dioxide to produce low-carbon fuels, Chem. Soc. Rev. 43 (2014) 631–675.
design, product purification have not been solved and current devel­
[3] R. Lin, J. Guo, X. Li, P. Patel, A. Seifitokaldani, Electrochemical reactors for CO2
opment fails to meet the requirements of industrial applications. For conversion, Catalysts 10 (2020) 473–506.
industrial applications, a current density of higher than 300 mA/cm2, a [4] M. Abdinejad, C. Dao, B. Deng, F. Dinic, O. Voznyy, X.-A. Zhang, H.-B. Kraatz,
FE of above 80%, a stability of over 80000 h, and a potential of less than Electrocatalytic reduction of CO2 to CH4 and CO in aqueous solution using
pyridine-porphyrins immobilized onto carbon nanotubes, ACS Sustain. Chem.
1.8 V are required for CO2ERR to be economically viable [141]. The Eng. 8 (2020) 9549–9557.
above specifications surpass the current densities of tens of mA/cm2 and [5] K.P. Kuhl, E.R. Cave, D.N. Abram, T.F. Jaramillo, New insights into the
stabilities of thousands of hours achieved in the reported cases [22]. A electrochemical reduction of carbon dioxide on metallic copper surfaces, Energ.
Environ. Sci. 5 (2012) 7050–7059.
high overpotential is detrimental to the energy efficiency as well as [6] N.J. Firet, T. Burdyny, N.T. Nesbitt, S. Chandrashekar, A. Longo, W.A. Smith,
application of CO2ERR due to the burdensome input energy costs. Much Copper and silver gas diffusion electrodes performing CO2 reduction studied

29
S. Gu et al. Chemical Engineering Journal 470 (2023) 144249

through operando X-ray absorption spectroscopy, Cat. Sci. Technol. 10 (2020) [34] M. Abdinejad, A. Seifitokaldani, C. Dao, E.H. Sargent, X.-A. Zhang, H.B. Kraatz,
5870–5885. Enhanced electrochemical reduction of CO2 catalyzed by cobalt and iron amino
[7] Z. Tang, E. Nishiwaki, K.E. Fritz, T. Hanrath, J. Suntivich, Cu(I) reducibility porphyrin complexes, ACS Appl. Energy Mater. 2 (2019) 1330–1335.
controls ethylene vs ethanol selectivity on (100)-textured copper during pulsed [35] C. Costentin, M. Robert, J.-M. Savéant, Current issues in molecular catalysis
CO2 reduction, ACS Appl. Mater. Interfaces 13 (2021) 14050–14055. illustrated by iron porphyrins as catalysts of the CO2-to-CO electrochemical
[8] Q. Lu, J. Rosen, Y. Zhou, G.S. Hutchings, Y.C. Kimmel, J.G. Chen, F. Jiao, conversion, Acc. Chem. Res. 48 (2015) 2996–3006.
A selective and efficient electrocatalyst for carbon dioxide reduction, Nat. [36] C. Costentin, G. Passard, M. Robert, J.-M. Savéant, Ultraefficient homogeneous
Commun. 5 (2014) 3242–3247. catalyst for the CO2-to-CO electrochemical conversion, Proc. Natl. Acad. Sci. 111
[9] W.-H. Cheng, M.H. Richter, I. Sullivan, D.M. Larson, C. Xiang, B.S. Brunschwig, (2014) 14990–14994.
H.A. Atwater, CO2 Reduction to CO with 19% efficiency in a solar-driven gas [37] A.N. Marianov, A.S. Kochubei, T. Roman, O.J. Conquest, C. Stampfl, Y. Jiang,
diffusion electrode flow cell under outdoor solar illumination, ACS Energy Lett. 5 Resolving deactivation pathways of Co porphyrin-based electrocatalysts for CO2
(2020) 470–476. reduction in aqueous medium, ACS Catal. 11 (2021) 3715–3729.
[10] S. Meshitsuka, M. Ichikawa, K. Tamaru, Electrocatalysis by metal [38] M. Zhu, D.-T. Yang, R. Ye, J. Zeng, N. Corbin, K. Manthiram, Inductive and
phthalocyanines in the reduction of carbon dioxide, J. Chem. Soc. Chem. electrostatic effects on cobalt porphyrins for heterogeneous electrocatalytic
Commun. (1974) 158–159. carbon dioxide reduction, Cat. Sci. Technol. 9 (2019) 974–980.
[11] Q. Feng, Y. Sun, X. Gu, Z. Dong, Advances of cobalt phthalocyanine in [39] S. Gu, A.N. Marianov, Y. Jiang, Covalent grafting of cobalt aminoporphyrin-based
electrocatalytic CO2 reduction to CO: a mini review, Electrocatalysis 13 (2022) electrocatalyst onto carbon nanotubes for excellent activity in CO2 reduction,
675–690. Appl Catal B 300 (2022), 120750.
[12] E. Boutin, M. Wang, J.C. Lin, M. Mesnage, D. Mendoza, B. Lassalle-Kaiser, [40] C.S. Diercks, S. Lin, N. Kornienko, E.A. Kapustin, E.M. Nichols, C. Zhu, Y. Zhao, C.
C. Hahn, T.F. Jaramillo, M. Robert, Aqueous electrochemical reduction of carbon J. Chang, O.M. Yaghi, Reticular electronic tuning of porphyrin active sites in
dioxide and carbon monoxide into methanol with cobalt phthalocyanine, Angew. covalent organic frameworks for electrocatalytic carbon dioxide reduction, J. Am.
Chem. Int. Ed. 58 (2019) 16172–16176. Chem. Soc. 140 (2018) 1116–1122.
[13] K.-M. Lam, K.-Y. Wong, S.-M. Yang, C.-M. Che, Cobalt and nickel complexes of [41] A.D. Handoko, F. Wei, Jenndy, B.S. Yeo, Z.W. Seh, Understanding heterogeneous
2,2′ : 6′,2″ : 6″,2‴-quaterpyridine as catalysts for electrochemical reduction of electrocatalytic carbon dioxide reduction through operando techniques. Nat. Cat.
carbon dioxide, J. Chem. Soc. Dalton Trans. (7) (1995) 1103–1107. 1 (2018) 922–934.
[14] Z. Guo, S. Cheng, C. Cometto, E. Anxolabéhère-Mallart, S.-M. Ng, C.-C. Ko, G. Liu, [42] Y. Wu, J. Jiang, Z. Weng, M. Wang, D.L.J. Broere, Y. Zhong, G.W. Brudvig,
L. Chen, M. Robert, T.-C. Lau, Highly efficient and selective photocatalytic CO2 Z. Feng, H. Wang, Electroreduction of CO2 catalyzed by a heterogenized Zn-
reduction by iron and cobalt quaterpyridine complexes, J. Am. Chem. Soc. 138 porphyrin complex with a redox-innocent metal center, ACS Cent. Sci. 3 (2017)
(2016) 9413–9416. 847–852.
[15] R.B. Ambre, Q. Daniel, T. Fan, H. Chen, B. Zhang, L. Wang, M.S.G. Ahlquist, [43] J. Han, P. An, S. Liu, X. Zhang, D. Wang, Y. Yuan, J. Guo, X. Qiu, K. Hou, L. Shi,
L. Duan, L. Sun, Molecular engineering for efficient and selective iron porphyrin Y. Zhang, S. Zhao, C. Long, Z. Tang, Reordering d-orbital energies of single-site
catalysts for electrochemical reduction of CO2 to CO, Chem. Commun. 52 (2016) catalysts for CO2 electroreduction, Angew. Chem. Int. Ed. 58 (2019)
14478–14481. 12711–12716.
[16] I. Hod, M.D. Sampson, P. Deria, C.P. Kubiak, O.K. Farha, J.T. Hupp, Fe-porphyrin- [44] S. Lin, C.S. Diercks, Y.-B. Zhang, N. Kornienko, E.M. Nichols, Y. Zhao, A.R. Paris,
based metal-organic framework films as high-surface concentration, D. Kim, P. Yang, O.M. Yaghi, C.J. Chang, Covalent organic frameworks
heterogeneous catalysts for electrochemical reduction of CO2, ACS Catal. 5 comprising cobalt porphyrins for catalytic CO2 reduction in water, Science 349
(2015) 6302–6309. (2015) 1208–1213.
[17] P. Gotico, Z. Halime, A. Aukauloo, Recent advances in metalloporphyrin-based [45] N. Kornienko, Y. Zhao, C.S. Kley, C. Zhu, D. Kim, S. Lin, C.J. Chang, O.M. Yaghi,
catalyst design towards carbon dioxide reduction: from bio-inspired second P. Yang, Metal-organic frameworks for electrocatalytic reduction of carbon
coordination sphere modifications to hierarchical architectures, Dalton Trans. 49 dioxide, J. Am. Chem. Soc. 137 (2015) 14129–14135.
(2020) 2381–2396. [46] Y.-R. Wang, H.-M. Ding, X.-Y. Ma, M. Liu, Y.-L. Yang, Y. Chen, S.-L. Li, Y.-Q. Lan,
[18] G.F. Manbeck, E. Fujita, A review of iron and cobalt porphyrins, phthalocyanines Imparting CO2 electroreduction auxiliary for integrated morphology tuning and
and related complexes for electrochemical and photochemical reduction of performance boosting in a porphyrin-based covalent organic framework, Angew.
carbon dioxide, J. Porphyrins Phthalocyanines 19 (2015) 45–64. Chem. Int. Ed. 61 (2022) e202114648.
[19] K.E. Dalle, J. Warnan, J.J. Leung, B. Reuillard, I.S. Karmel, E. Reisner, Electro- [47] N. Corbin, J. Zeng, K. Williams, K. Manthiram, Heterogeneous molecular catalysts
and solar-driven fuel synthesis with first row transition metal complexes, Chem. for electrocatalytic CO2 reduction, Nano Res. 12 (2019) 2093–2125.
Rev. 119 (2019) 2752–2875. [48] A.J. Bard, L.R. Faulkner, Electrochemical methods: fundamentals and
[20] F. Franco, C. Rettenmaier, H.S. Jeon, B. Roldan Cuenya, Transition metal-based applications, 2nd ed., John Wiley & Sons Inc, New York, 2001.
catalysts for the electrochemical CO2 reduction: from atoms and molecules to [49] X. Lu, B. Dereli, T. Shinagawa, M. Eddaoudi, L. Cavallo, K. Takanabe, High
nanostructured materials, Chem. Soc. Rev. 49 (2020) 6884–6946. current density microkinetic and electronic structure analysis of CO2 reduction
[21] L. Sun, V. Reddu, A.C. Fisher, X. Wang, Electrocatalytic reduction of carbon using Co and Fe complexes on gas diffusion electrode, Chem Catalysis 2 (2022)
dioxide: opportunities with heterogeneous molecular catalysts, Energ. Environ. 1143–1162.
Sci. 13 (2020) 374–403. [50] M. Zhu, J. Chen, L. Huang, R. Ye, J. Xu, Y.-F. Han, Covalently grafting cobalt
[22] S. Jin, Z. Hao, K. Zhang, Z. Yan, J. Chen, Advances and challenges for the porphyrin onto carbon nanotubes for efficient CO2 electroreduction, Angew.
electrochemical reduction of CO2 to CO: from fundamentals to industrialization, Chem. Int. Ed. 58 (2019) 6595–6599.
Angew. Chem. Int. Ed. 133 (2021) 20795–20816. [51] A.S. Malkani, J. Li, N.J. Oliveira, M. He, X. Chang, B. Xu, Q. Lu, Understanding
[23] J.-M. Savéant, Molecular catalysis of electrochemical reactions. mechanistic the electric and nonelectric field components of the cation effect on the
aspects, Chem. Rev. 108 (7) (2008) 2348–2378. electrochemical CO reduction reaction. Science, Advances 6 (2020) eabd2569.
[24] N. Sonoyama, M. Kirii, T. Sakata, Electrochemical reduction of CO2 at metal- [52] M. Zhu, R. Ye, K. Jin, N. Lazouski, K. Manthiram, Elucidating the reactivity and
porphyrin supported gas diffusion electrodes under high pressure CO2, mechanism of CO2 electroreduction at highly dispersed cobalt phthalocyanine,
Electrochem. Commun. 1 (1999) 213–216. ACS Energy Lett. 3 (2018) 1381–1386.
[25] P. Saha, S. Amanullah, A. Dey, Selectivity in electrochemical CO2 reduction, Acc. [53] J. Chen, M. Zhu, J. Li, J. Xu, Y.-F. Han, Structure-activity relationship of the
Chem. Res. 55 (2022) 134–144. polymerized cobalt phthalocyanines for electrocatalytic carbon dioxide
[26] L. Feng, K.-Y. Wang, E. Joseph, H.-C. Zhou, Catalytic porphyrin framework reduction, J. Phys. Chem. C 124 (2020) 16501–16507.
compounds, Trends Chem. 2 (2020) 555–568. [54] C. Costentin, S. Drouet, M. Robert, J.-M. Savéant, Turnover numbers, turnover
[27] N. Furuya, S. Koide, Electroreduction of carbon dioxide by metal frequencies, and overpotential in molecular catalysis of electrochemical
phthalocyanines, Electrochim. Acta 36 (8) (1991) 1309–1313. reactions. cyclic voltammetry and preparative-scale electrolysis, J. Am. Chem.
[28] W. Luo, W. Xie, M. Li, J. Zhang, A. Züttel, 3D hierarchical porous indium catalyst Soc. 134 (2012) 11235–11242.
for highly efficient electroreduction of CO2, J. Mater. Chem. A 7 (2019) [55] C. Costentin, J.-M. Savéant, Multielectron, multistep molecular catalysis of
4505–4515. electrochemical reactions: benchmarking of homogeneous catalysts,
[29] X.-M. Hu, M.H. Rønne, S.U. Pedersen, T. Skrydstrup, K. Daasbjerg, Enhanced ChemElectroChem 1 (7) (2014) 1226–1236.
catalytic activity of cobalt porphyrin in CO2 electroreduction upon [56] C. Costentin, S. Drouet, G. Passard, M. Robert, J.-M. Savéant, Proton-coupled
immobilization on carbon materials, Angew. Chem. Int. Ed. 56 (2017) electron transfer cleavage of heavy-atom bonds in electrocatalytic processes.
6468–6472. cleavage of a C-O bond in the catalyzed electrochemical reduction of CO2, J. Am.
[30] I. Azcarate, C. Costentin, M. Robert, J.-M. Savéant, Through-space charge Chem. Soc. 135 (2013) 9023–9031.
interaction substituent effects in molecular catalysis leading to the design of the [57] R. Francke, B. Schille, M. Roemelt, Homogeneously catalyzed electroreduction of
most efficient catalyst of CO2-to-CO electrochemical conversion, J. Am. Chem. carbon dioxide-methods, mechanisms, and catalysts, Chem. Rev. 118 (2018)
Soc. 138 (2016) 16639–16644. 4631–4701.
[31] D. Behar, T. Dhanasekaran, P. Neta, C.M. Hosten, D. Ejeh, P. Hambright, E. Fujita, [58] T. Katsuhiro, H. Kazuya, S. Hideo, T. Shinobu, Electrocatalytic behavior of metal
Cobalt porphyrin catalyzed reduction of CO2. radiation chemical, photochemical, porphyrins in the reduction of carbon dioxide, Chem. Lett. 8 (1979) 305–308.
and electrochemical studies, Chem. A Eur. J. 102 (1998) 2870–2877. [59] Y. Wu, G. Hu, C.L. Rooney, G.W. Brudvig, H. Wang, Heterogeneous nature of
[32] X. Chen, X.-M. Hu, K. Daasbjerg, M.S.G. Ahlquist, Understanding the enhanced electrocatalytic CO/CO2 reduction by cobalt phthalocyanines, ChemSusChem 13
catalytic CO2 reduction upon adhering cobalt porphyrin to carbon nanotubes and (2020) 6296–6299.
the inverse loading effect, Organometallics 39 (2020) 1634–1641. [60] M. Hammouche, D. Lexa, J.M. Savéant, M. Momenteau, Catalysis of the
[33] C. Costentin, S. Drouet, M. Robert, J.-M. Savéant, A local proton source enhances electrochemical reduction of carbon dioxide by iron(“0”) porphyrins,
CO2 electroreduction to CO by a molecular Fe catalyst, Science 338 (2012) 90–94. J. Electroanal. Chem. Interfacial Electrochem. 249 (1988) 347–351.

30
S. Gu et al. Chemical Engineering Journal 470 (2023) 144249

[61] I. Bhugun, D. Lexa, J.-M. Savéant, Catalysis of the electrochemical reduction of reduction of CO2 to CO using a three-dimensional porphyrin/graphene hydrogel,
carbon dioxide by iron(0) porphyrins. synergistic effect of lewis acid cations, Energ. Environ. Sci. 12 (2019) 747–755.
J. Phys. Chem. 100 (1996) 19981–19985. [88] M. Abdinejad, K. Tang, C. Dao, S. Saedy, T. Burdyny, Immobilization strategies
[62] M. Hammouche, D. Lexa, M. Momenteau, J.M. Saveant, Chemical catalysis of for porphyrin-based molecular catalysts for the electroreduction of CO2, J. Mater.
electrochemical reactions. homogeneous catalysis of the electrochemical Chem. A 10 (2022) 7626–7636.
reduction of carbon dioxide by iron(“0”) porphyrins. role of the addition of [89] F. Greenwell, G. Neri, V. Piercy, A.J. Cowan, Noncovalent immobilization of a
magnesium cations, J. Am. Chem. Soc. 113 (1991) 8455–8466. nickel cyclam catalyst on carbon electrodes for CO2 reduction using aqueous
[63] I. Bhugun, D. Lexa, J.-M. Saveant, Ultraefficient selective homogeneous catalysis electrolyte, Electrochim. Acta 392 (2021) 139015–139020.
of the electrochemical reduction of carbon dioxide by an iron(0) porphyrin [90] P. Kang, S. Zhang, T.J. Meyer, M. Brookhart, Rapid selective electrocatalytic
associated with a weak broensted acid cocatalyst, J. Am. Chem. Soc. 116 (1994) reduction of carbon dioxide to formate by an iridium pincer catalyst immobilized
5015–5016. on carbon nanotube electrodes, Angew. Chem. Int. Ed. 53 (2014) 8709–8713.
[64] I. Bhugun, D. Lexa, J.-M. Savéant, Catalysis of the electrochemical reduction of [91] M.L. Bowers, B.A. Yenser, Electrochemical behavior of glassy carbon electrodes
carbon dioxide by iron(0) porphyrins: synergystic effect of weak brönsted acids, modified by electrochemical oxidation, Anal. Chim. Acta 243 (1991) 43–53.
J. Am. Chem. Soc. 118 (1996) 1769–1776. [92] A.D. Jannakoudakis, P.D. Jannakoudakis, E. Theodoridou, J.O. Besenhard,
[65] C. Costentin, G. Passard, M. Robert, J.-M. Savéant, Pendant acid-base groups in Electrochemical oxidation of carbon fibres in aqueous solutions and analysis of
molecular catalysts: H-bond promoters or proton relays? mechanisms of the the surface oxides, J. Appl. Electrochem. 20 (1990) 619–624.
conversion of CO2 to CO by electrogenerated iron(0) porphyrins bearing [93] K.J. Ziegler, Z. Gu, H. Peng, E.L. Flor, R.H. Hauge, R.E. Smalley, Controlled
prepositioned phenol functionalities, J. Am. Chem. Soc. 136 (2014) oxidative cutting of single-walled carbon nanotubes, J. Am. Chem. Soc. 127
11821–11829. (2005) 1541–1547.
[66] I. Azcarate, C. Costentin, M. Robert, J.-M. Savéant, Dissection of electronic [94] M. Delamar, R. Hitmi, J. Pinson, J.M. Saveant, Covalent modification of carbon
substituent effects in multielectron-multistep molecular catalysis. electrochemical surfaces by grafting of functionalized aryl radicals produced from electrochemical
CO2-to-CO conversion catalyzed by iron porphyrins, J. Phys. Chem. C 120 (2016) reduction of diazonium salts, J. Am. Chem. Soc. 114 (1992) 5883–5884.
28951–28960. [95] P. Allongue, M. Delamar, B. Desbat, O. Fagebaume, R. Hitmi, J. Pinson, J.-
[67] C. Costentin, M. Robert, J.-M. Savéant, A. Tatin, Efficient and selective molecular M. Savéant, Covalent modification of carbon surfaces by aryl radicals generated
catalyst for the CO2-to-CO electrochemical conversion in water, Proc. Natl. Acad. from the electrochemical reduction of diazonium salts, J. Am. Chem. Soc. 119
Sci. 112 (2015) 6882–6886. (1997) 201–207.
[68] E.A. Mohamed, Z.N. Zahran, Y. Naruta, Efficient electrocatalytic CO2 reduction [96] V. Georgakilas, K. Kordatos, M. Prato, D.M. Guldi, M. Holzinger, A. Hirsch,
with a molecular cofacial iron porphyrin dimer, Chem. Commun. 51 (2015) Organic functionalization of carbon nanotubes, J. Am. Chem. Soc. 124 (2002)
16900–16903. 760–761.
[69] E. Simón-Manso, C.P. Kubiak, Dinuclear nickel complexes as catalysts for [97] N. Tagmatarchis, M. Prato, Functionalization of carbon nanotubes via 1,3-dipolar
electrochemical reduction of carbon dioxide, Organometallics 24 (1) (2005) cycloadditions, J. Mater. Chem. 14 (2004) 437–439.
96–102. [98] C. Cioffi, S. Campidelli, F.G. Brunetti, M. Meneghetti, M. Prato, Functionalisation
[70] J.Y. Becker, B. Vainas, R. Eger, L. Kaufman, Electrocatalytic reduction of CO2 to of carbon nanohorns, Chem. Commun. (2006) 2129–2131.
oxalate by Ag and Pd porphyrins, J. Chem. Soc. Chem. Commun. (1985) [99] C.E. Halbig, P. Rietsch, S. Eigler, Towards the synthesis of graphene azide from
1471–1472. graphene oxide, Molecules 20 (2015) 21050–21057.
[71] C.M. Lieber, N.S. Lewis, Catalytic reduction of carbon dioxide at carbon [100] H.D. Abruna, T.J. Meyer, R.W. Murray, Chemical and electrochemical properties
electrodes modified with cobalt phthalocyanine, J. Am. Chem. Soc. 106 (1984) of 2, 2’-bipyridyl complexes of ruthenium covalently bound to platinum oxide
5033–5034. electrodes, Inorg. Chem. 18 (1979) 3233–3240.
[72] N. Coutard, N. Kaeffer, V. Artero, Molecular engineered nanomaterials for [101] A. Maurin, M. Robert, Catalytic CO2-to-CO conversion in water by covalently
catalytic hydrogen evolution and oxidation, Chem. Commun. 52 (2016) functionalized carbon nanotubes with a molecular iron catalyst, Chem. Commun.
13728–13748. 52 (81) (2016) 12084–12087.
[73] X. Liu, S. Inagaki, J. Gong, Heterogeneous molecular systems for photocatalytic [102] S.A. Yao, R.E. Ruther, L. Zhang, R.A. Franking, R.J. Hamers, J.F. Berry, Covalent
CO2 reduction with water oxidation, Angew. Chem. Int. Ed. 55 (2016) attachment of catalyst molecules to conductive diamond: CO2 reduction using
14924–14950. “smart” electrodes, J. Am. Chem. Soc. 134 (2012) 15632–15635.
[74] R.M. Bullock, A.K. Das, A.M. Appel, Surface immobilization of molecular [103] A.N. Marianov, Y. Jiang, Covalent ligation of Co molecular catalyst to carbon
electrocatalysts for energy conversion, Chem. A Eur. J. 23 (2017) 7626–7641. cloth for efficient electroreduction of CO2 in water, Appl Catal B 244 (2019)
[75] L. Zhang, J.M. Cole, Anchoring groups for dye-sensitized solar cells, ACS Appl. 881–888.
Mater. Interfaces 7 (6) (2015) 3427–3455. [104] E.A. Mohamed, Z.N. Zahran, Y. Naruta, Efficient heterogeneous CO2 to CO
[76] A. Ambrosi, C.K. Chua, A. Bonanni, M. Pumera, Electrochemistry of graphene and conversion with a phosphonic acid fabricated cofacial iron porphyrin dimer,
related materials, Chem. Rev. 114 (2014) 7150–7188. Chem. Mater. 29 (2017) 7140–7150.
[77] M.N. Hossain, P. Prslja, C. Flox, N. Muthuswamy, J. Sainio, A.M. Kannan, [105] T. Atoguchi, A. Aramata, A. Kazusaka, M. Enyo, Cobalt(II)-tetraphenylporphyrin-
M. Suominen, N. Lopez, T. Kallio, Temperature dependent product distribution of pyridine complex fixed on a glassy carbon electrode and its prominent catalytic
electrochemical CO2 reduction on CoTPP/MWCNT composite, Appl Catal B 304 activity for reduction of carbon dioxide, J. Chem. Soc. Chem. Commun. (1991)
(2022) 120863–120873. 156–157.
[78] T.V. Magdesieva, T. Yamamoto, D.A. Tryk, A. Fujishima, Electrochemical [106] T. Atoguchi, A. Aramata, A. Kazusaka, M. Enyo, Electrocatalytic activity of CoII
reduction of CO2 with transition metal phthalocyanine and porphyrin complexes TPP-pyridine complex modified carbon electrode for CO2 reduction,
supported on activated carbon fibers, J. Electrochem. Soc. 149 (2002) D89–D95. J. Electroanal. Chem. Interfacial Electrochem. 318 (1991) 309–320.
[79] Z. Weng, J. Jiang, Y. Wu, Z. Wu, X. Guo, K.L. Materna, W. Liu, V.S. Batista, G. [107] H. Tanaka, A. Aramata, Aminopyridyl cation radical method for bridging between
W. Brudvig, H. Wang, Electrochemical CO2 reduction to hydrocarbons on a metal complex and glassy carbon: cobalt(II) tetraphenylporphyrin bonded on
heterogeneous molecular Cu catalyst in aqueous solution, J. Am. Chem. Soc. 138 glassy carbon for enhancement of CO2 electroreduction, J. Electroanal. Chem.
(2016) 8076–8079. 437 (1-2) (1997) 29–35.
[80] J. Shen, R. Kortlever, R. Kas, Y.Y. Birdja, O. Diaz-Morales, Y. Kwon, I. Ledezma- [108] S. De, T. Devic, A. Fateeva, Porphyrin and phthalocyanine-based metal organic
Yanez, K.J.P. Schouten, G. Mul, M.T.M. Koper, Electrocatalytic reduction of frameworks beyond metal-carboxylates, Dalton Trans. 50 (2021) 1166–1188.
carbon dioxide to carbon monoxide and methane at an immobilized cobalt [109] F.N. Al-Rowaili, A. Jamal, M.S. Ba Shammakh, A. Rana, A review on recent
protoporphyrin, Nat. Commun. 6 (2015) 8177–8184. advances for electrochemical reduction of carbon dioxide to methanol using
[81] S. Aoi, K. Mase, K. Ohkubo, S. Fukuzumi, Selective electrochemical reduction of metal-organic framework (MOF) and non-MOF catalysts: challenges and future
CO2 to CO with a cobalt chlorin complex adsorbed on multi-walled carbon prospects, ACS Sustain. Chem. Eng. 6 (2018) 15895–15914.
nanotubes in water, Chem. Commun. 51 (2015) 10226–10228. [110] D. Quezada, J. Honores, M. García, F. Armijo, M. Isaacs, Electrocatalytic
[82] Y.Y. Birdja, R.E. Vos, T.A. Wezendonk, L. Jiang, F. Kapteijn, M.T.M. Koper, Effects reduction of carbon dioxide on a cobalt tetrakis(4-aminophenyl)porphyrin
of substrate and polymer encapsulation on CO2 electroreduction by immobilized modified electrode in BMImBF4, New J. Chem. 38 (2014) 3606–3612.
indium(III) protoporphyrin, ACS Catal. 8 (2018) 4420–4428. [111] D. Quezada, J. Honores, M.J. Aguirre, M. Isaacs, Electrocatalytic reduction of
[83] Y. Liu, C.C.L. McCrory, Modulating the mechanism of electrocatalytic CO2 carbon dioxide on conducting glass electrode modified with polymeric porphyrin
reduction by cobalt phthalocyanine through polymer coordination and films containing transition metals in ionic liquid medium, J. Coord. Chem. 67
encapsulation, Nat. Commun. 10 (2019) 1683–1692. (2014) 4090–4100.
[84] A. Maurin, M. Robert, Noncovalent immobilization of a molecular iron-based [112] J.E. Pander III, A. Fogg, A.B. Bocarsly, Utilization of electropolymerized films of
electrocatalyst on carbon electrodes for selective, efficient CO2-to-CO conversion cobalt porphyrin for the reduction of carbon dioxide in aqueous media,
in water, J. Am. Chem. Soc. 138 (8) (2016) 2492–2495. ChemCatChem 8 (2016) 3536–3545.
[85] S. Dou, L. Sun, S. Xi, X. Li, T. Su, H.J. Fan, X. Wang, Enlarging the π-conjugation [113] X.-M. Hu, Z. Salmi, M. Lillethorup, E.B. Pedersen, M. Robert, S.U. Pedersen,
of cobalt porphyrin for highly active and selective CO2 electroreduction, T. Skrydstrup, K. Daasbjerg, Controlled electropolymerisation of a carbazole-
ChemSusChem 14 (2021) 2126–2132. functionalised iron porphyrin electrocatalyst for CO2 reduction, Chem. Commun.
[86] J. Choi, P. Wagner, R. Jalili, J. Kim, D.R. MacFarlane, G.G. Wallace, D.L. Officer, 52 (2016) 5864–5867.
A porphyrin/graphene framework: a highly efficient and robust electrocatalyst [114] S. Watpathomsub, J. Luangchaiyaporn, N.S. Sariciftci, P. Thamyongkit, Efficient
for carbon dioxide reduction, Adv. Energy Mater. 8 (26) (2018) 1801280. heterogeneous catalysis by pendant metalloporphyrin-functionalized
[87] J. Choi, J. Kim, P. Wagner, S. Gambhir, R. Jalili, S. Byun, S. Sayyar, Y.M. Lee, D. polythiophenes for the electrochemical reduction of carbon dioxide, New J.
R. MacFarlane, G.G. Wallace, D.L. Officer, Energy efficient electrochemical Chem. 44 (2020) 12486–12495.

31
S. Gu et al. Chemical Engineering Journal 470 (2023) 144249

[115] T. Wang, L. Xu, Z. Chen, L. Guo, Y. Zhang, R. Li, T. Peng, Central site regulation of [130] Y. Lu, J. Zhang, W. Wei, D.-D. Ma, X.-T. Wu, Q.-L. Zhu, Efficient carbon dioxide
cobalt porphyrin conjugated polymer to give highly active and selective CO2 electroreduction over ultrathin covalent organic framework nanolayers with
reduction to CO in aqueous solution, Appl Catal B 291 (2021), 120128. isolated cobalt porphyrin units, ACS Appl. Mater. Interfaces 12 (2020)
[116] H.C. Zhou, J.R. Long, O.M. Yaghi, Introduction to Metal-Organic Frameworks, 37986–37992.
Chem. Rev. 112 (2012) 673–674. [131] J.-X. Wu, S.-Z. Hou, X.-D. Zhang, M. Xu, H.-F. Yang, P.-S. Cao, Z.-Y. Gu,
[117] H.B. Wu, X.W. Lou, Metal-organic frameworks and their derived materials for Cathodized copper porphyrin metal-organic framework nanosheets for selective
electrochemical energy storage and conversion: Promises and challenges. Sci. formate and acetate production from CO2 electroreduction, Chem. Sci. 10 (2019)
Adv. 3 (2017) eaap9252. 2199–2205.
[118] B.-X. Dong, S.-L. Qian, F.-Y. Bu, Y.-C. Wu, L.-G. Feng, Y.-L. Teng, W.-L. Liu, Z.- [132] K. Torbensen, D. Joulié, S. Ren, M. Wang, D. Salvatore, C.P. Berlinguette,
W. Li, Electrochemical reduction of CO2 to CO by a heterogeneous catalyst of Fe- M. Robert, Molecular catalysts boost the rate of electrolytic CO2 reduction, ACS
porphyrin-based metal-organic framework, ACS Appl. Energy Mater. 1 (2018) Energy Lett. 5 (2020) 1512–1518.
4662–4669. [133] K. Torbensen, C. Han, B. Boudy, N. von Wolff, C. Bertail, W. Braun, M. Robert,
[119] Q. Huang, Q. Li, J. Liu, Y.R. Wang, R. Wang, L.Z. Dong, Y.H. Xia, J.L. Wang, Y.- Iron porphyrin allows fast and selective electrocatalytic conversion of CO2 to CO
Q. Lan, Disclosing CO2 activation mechanism by hydroxyl-induced crystalline in a flow cell, Chem. A Eur. J. 26 (2020) 3034–3038.
structure transformation in electrocatalytic process, Matter 1 (2019) 1656–1668. [134] N.G. Yasri, T.A. Al-Attas, J. Hu, M.G. Kibria, Electropolymerized metal-
[120] Y.T. Guntern, J.R. Pankhurst, J. Vávra, M. Mensi, V. Mantella, P. Schouwink, protoporphyrin electrodes for selective electrochemical reduction of CO2, Cat. Sci.
R. Buonsanti, Nanocrystal/metal-organic framework hybrids as electrocatalytic Technol. 11 (2021) 1580–1589.
platforms for CO2 conversion, Angew. Chem. Int. Ed. 58 (2019) 12632–12639. [135] P. Yu, X. Lv, Q. Wang, H. Huang, W. Weng, C. Peng, L. Zhang, G. Zheng,
[121] Z. Xin, Y.-R. Wang, Y. Chen, W.-L. Li, L.-Z. Dong, Y.-Q. Lan, Metallocene Promoting electrocatalytic CO2 reduction to CH4 by copper porphyrin with donor-
implanted metalloporphyrin organic framework for highly selective CO2 acceptor structures, Small 19 (4) (2023) 2205730.
electroreduction, Nano Energy 67 (2020), 104233. [136] Y.-R. Wang, M. Liu, G.-K. Gao, Y.-L. Yang, R.-X. Yang, H.-M. Ding, Y. Chen, S.-
[122] Y. Guo, W. Shi, H. Yang, Q. He, Z. Zeng, J.-Y. Ye, X. He, R. Huang, C. Wang, L. Li, Y.-Q. Lan, Implanting numerous hydrogen-bonding networks in a Cu-
W. Lin, Cooperative stabilization of the [pyridinium-CO2-Co] adduct on a metal- porphyrin-based nanosheet to boost CH4 selectivity in neutral-media CO2
organic layer enhances electrocatalytic CO2 reduction, J. Am. Chem. Soc. 141 electroreduction, Angew. Chem. Int. Ed. 60 (2021) 21952–21958.
(2019) 17875–17883. [137] M. Abdinejad, C. Dao, X.-A. Zhang, H.B. Kraatz, Enhanced electrocatalytic activity
[123] P.T. Smith, B.P. Benke, Z. Cao, Y. Kim, E.M. Nichols, K. Kim, C.J. Chang, Iron of iron amino porphyrins using a flow cell for reduction of CO2 to CO, J. Energy
porphyrins embedded into a supramolecular porous organic cage for Chem. 58 (2021) 162–169.
electrochemical CO2 reduction in water, Angew. Chem. Int. Ed. 57 (2018) [138] M.N. Hossain, R. Khakpour, M. Busch, M. Suominen, K. Laasonen, T. Kallio,
9684–9688. Temperature-controlled syngas production via electrochemical CO2 reduction on
[124] Q. Wu, R.-K. Xie, M.-J. Mao, G.-L. Chai, J.-D. Yi, S.-S. Zhao, Y.-B. Huang, R. Cao, a CoTPP/MWCNT composite in a flow cell, ACS Appl. Energy Mater. 6 (2023)
Integration of strong electron transporter tetrathiafulvalene into 267–277.
metalloporphyrin-based covalent organic framework for highly efficient [139] X. Zhang, Y. Wang, M. Gu, M. Wang, Z. Zhang, W. Pan, Z. Jiang, H. Zheng,
electroreduction of CO2, ACS Energy Lett. 5 (2020) 1005–1012. M. Lucero, H. Wang, G.E. Sterbinsky, Q. Ma, Y.-G. Wang, Z. Feng, J. Li, H. Dai,
[125] P.L. Cheung, S.K. Lee, C.P. Kubiak, Facile solvent-free synthesis of thin iron Y. Liang, Molecular engineering of dispersed nickel phthalocyanines on carbon
porphyrin COFs on carbon cloth electrodes for CO2 reduction, Chem. Mater. 31 nanotubes for selective CO2 reduction, Nat. Energy 5 (2020) 684–692.
(2019) 1908–1919. [140] J. Su, J.-J. Zhang, J. Chen, Y. Song, L. Huang, M. Zhu, B.I. Yakobson, B.Z. Tang,
[126] L. Gong, B. Chen, Y. Gao, B. Yu, Y. Wang, B. Han, C. Lin, Y. Bian, D. Qi, J. Jiang, R. Ye, Building a stable cationic molecule/electrode interface for highly efficient
Covalent organic frameworks based on tetraphenyl-p-phenylenediamine and and durable CO2 reduction at an industrially relevant current, Energ. Environ. Sci.
metalloporphyrin for electrochemical conversion of CO2 to CO, Inorg. Chem. 14 (2021) 483–492.
Front. 9 (2022) 3217–3223. [141] F.-Y. Gao, R.-C. Bao, M.-R. Gao, S.-H. Yu, Electrochemical CO2-to-CO conversion:
[127] H. Dong, M. Lu, Y. Wang, H.-L. Tang, D. Wu, X. Sun, F.-M. Zhang, Covalently electrocatalysts, electrolytes, and electrolyzers, J. Mater. Chem. A 8 (2020)
anchoring covalent organic framework on carbon nanotubes for highly efficient 15458–15478.
electrocatalytic CO2 reduction, Appl Catal B 303 (2022) 120897–120905. [142] W. Chen, J. Cao, W. Fu, J. Zhang, G. Qian, J. Yang, D. Chen, X. Zhou, W. Yuan,
[128] Y. Song, J.-J. Zhang, Y. Dou, Z. Zhu, J. Su, L. Huang, W. Guo, X. Cao, L.e. Cheng, X. Duan, Molecular-level insights into the notorious CO poisoning of platinum
Z. Zhu, Z. Zhang, X. Zhong, D. Yang, Z. Wang, B.Z. Tang, B.I. Yakobson, R. Ye, catalyst, Angew. Chem. Int. Ed. 61 (2022) e202200190.
Atomically thin, ionic-covalent organic nanosheets for stable, high-performance [143] M.A. Riquelme, M. Isaacs, M. Lucero, E. Trollund, M.J. Aguirre, Electrocatalytic
carbon dioxide electroreduction, Adv. Mater. 34 (42) (2022) 2110496. reduction of carbon dioxide at polymeric cobalt tetra (3-amino (phenyl)
[129] D. Yang, S. Zuo, H. Yang, Y. Zhou, X. Wang, Freestanding millimeter-scale porphyrin glassy carbon-modified electrodes, J. Chil. Chem. Soc. 48 (2003)
porphyrin-based monoatomic layers with 0.28 nm thickness for CO2 89–92.
electrocatalysis, Angew. Chem. Int. Ed. 59 (2020) 18954–18959.

32

You might also like