0% found this document useful (0 votes)
12 views31 pages

Dissertation Last Final

This dissertation addresses the challenge of rising atmospheric CO₂ levels by exploring electrochemical CO₂ reduction (CO₂RR) using molecular catalysts based on Earth-abundant metals like iron, manganese, zinc, and nickel. It highlights the structural and electronic features of these catalysts, their effectiveness in converting CO₂ into valuable products, and discusses strategies to overcome limitations such as stability and low performance. The research emphasizes the importance of coordination chemistry and materials innovation in advancing scalable CO₂ conversion technologies.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
12 views31 pages

Dissertation Last Final

This dissertation addresses the challenge of rising atmospheric CO₂ levels by exploring electrochemical CO₂ reduction (CO₂RR) using molecular catalysts based on Earth-abundant metals like iron, manganese, zinc, and nickel. It highlights the structural and electronic features of these catalysts, their effectiveness in converting CO₂ into valuable products, and discusses strategies to overcome limitations such as stability and low performance. The research emphasizes the importance of coordination chemistry and materials innovation in advancing scalable CO₂ conversion technologies.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 31

Abstract

The rising atmospheric CO₂ levels, largely due to fossil fuel use and industrial activity,
pose a significant climate challenge. Electrochemical CO₂ reduction (CO₂RR) offers a
sustainable solution by converting CO₂ into valuable products like CO, formate,
methane, and alcohols. Central to this process is the development of efficient, selective,
and low-cost molecular catalysts capable of driving multi-electron, multi-proton
transformations under mild conditions.

This dissertation focuses on molecular catalysts based on Earth-abundant metals—iron


(Fe), manganese (Mn), zinc (Zn), and nickel (Ni)—highlighting their structural,
electronic, and mechanistic features. It begins with an overview of CO₂ coordination
chemistry, discussing how its linear geometry and electrophilic carbon affect activation
pathways. Various coordination modes (η¹, η², O,O-bound) are analyzed to understand
how metal centers and ligands influence reactivity.

Iron porphyrin complexes, especially TPP derivatives, show high selectivity and
efficiency for CO or formate production, enhanced by ligand tuning and second-sphere
effects. Zinc, though redox-inactive, activates CO₂ via innovative ligand designs,
including ZIF-8 and Zn–porphyrin hybrids. Manganese(I) tricarbonyl complexes mimic
rhenium catalysts and perform well in both homogeneous and MOF-supported systems,
with mechanisms involving hydride intermediates and photochemical steps.
Nickel-based systems, particularly with terpyridine and cyclam ligands, demonstrate
strong photocatalytic performance, aided by immobilization on quantum dots or porous
frameworks.

The work concludes by addressing limitations such as stability, photodegradation, and


low aqueous performance, and proposes strategies like catalyst immobilization and
ligand design to overcome them. This research underscores the importance of
coordination chemistry and materials innovation in advancing scalable CO₂ conversion
technologies.

1
Introduction

The persistent increase in atmospheric carbon dioxide (CO₂) concentrations, coupled


with the depletion of fossil fuel resources, continues to exacerbate global environmental
challenges. To address these concerns, two complementary strategies are critical:
rigorous monitoring of atmospheric CO₂ levels and the development of renewable
technologies capable of converting CO₂ into value-added products¹–³. The
transformation of CO₂ into alternative fuels or chemical feedstocks presents a promising
approach to mitigating carbon emissions while simultaneously reducing fossil fuel
dependence, thus tackling the root causes of climate change¹.

Electrocatalytic CO₂ reduction has emerged as a particularly attractive method due to its
operation under ambient conditions and its ability to integrate renewable electricity².
Among various pathways, the selective photocatalytic reduction of CO₂ to carbon
monoxide (CO) is of considerable interest, given CO’s utility as a precursor for
synthesizing liquid hydrocarbons and methanol⁴. A range of materials—including noble
metals (e.g., Cu, Au, Ag, Zn)⁵–⁸, metal oxide semiconductors (ZnO, SnO₂)⁹,¹⁰, layered
dichalcogenides (MoS₂, WSe₂)¹¹,¹², and transition-metal complexes (Mn, Fe, Co, Ni, Cu,
Ru, Re)¹³–¹⁵—have been extensively investigated for catalyzing CO₂ reduction to liquid
fuels.

Fundamentally, photocatalytic CO₂ conversion involves three sequential processes: light


absorption, charge carrier separation and migration, and surface catalytic reactions.
Although substantial advancements have been achieved in optimizing light harvesting
and charge dynamics—owing in part to insights from solar water splitting
research¹⁶–²²—the surface reactions specific to CO₂ photoreduction remain less
well-developed²³,²⁴. This distinction is particularly critical, as the surface reactions
govern the ultimate catalytic efficiency and selectivity.

Current strategies for CO₂ reduction catalysis predominantly explore two classes of
materials: homogeneous molecular complexes (comprising coordination and

2
organometallic systems) and heterogeneous solid-state catalysts. In this review,
emphasis is placed on homogeneous molecular systems, which offer fine-tunable
reactivity through ligand design and metal center modulation. Effective molecular
catalysts must mediate multi-electron and multi-proton transformations, either by
stabilizing reduced metal centers via appropriate ligand fields or by enabling
ligand-centered redox processes with metals facilitating electron relay²⁵–²⁸.

Designing robust molecular electrocatalysts poses inherent challenges. An ideal


catalyst must exhibit suitable electronic properties, including an appropriately positioned
conduction band edge, chemical stability of the reduced form, and the ability to sustain
fast catalytic turnovers under practical electrolysis conditions²⁶,²⁷. In metal-organic
frameworks, such parameters are highly tunable through metal selection and ligand
engineering, allowing researchers to tailor catalysts for specific CO₂ reduction pathways.
Over the past decades, significant progress has been made in enhancing both the
catalytic activity and the mechanistic understanding of molecular systems for CO₂
electroreduction, establishing a strong foundation for future advancements in
sustainable carbon management technologies.

3
2. Background of Molecular Catalysts

2.1 Fundamental Concepts

Electrochemical processes typically occur in tandem, where an oxidation reaction at the


anode is simultaneously coupled with a reduction process at the cathode within a
photoelectrochemical cell (PEC). These paired electrochemical transformations can
individually yield significant products or intermediates crucial for synthetic chemistry and
represent a comprehensive and sustainable strategy for chemical production. Previous
research has documented notable cases of paired electro-organic reactions. Among
these, the chlor-alkali process stands as the most widely implemented. This co-reaction
strategy offers enhanced energy utilization and atom economy when compared to
separately carried-out processes. Notably, it allows for theoretical Faradaic efficiencies
up to 200%, since each electron can contribute to two separate reactions.

In the context of CO₂ electroreduction, the standard industrial configuration typically


involves coupling with the oxygen evolution reaction (OER) as the anodic process.
Electrocatalysts, broadly defined, act as facilitators of electron transfer and ideally
operate near the thermodynamic potentials required for substrate conversion. The
overpotential, defined as the difference between the externally applied voltage and the
thermodynamic redox potential (E⁰), is a key limitation in CO₂ electroreduction, as most
electrode materials necessitate significant excess voltage, reducing the overall
conversion efficiency. Both thermodynamic feasibility and reaction kinetics are critical in
this scenario.

The main role of a catalyst is to lower the overpotential. Thus, the design of effective
catalysts involves several criteria: (a) possessing redox potentials [E⁰(Catⁿ⁺/⁰)] closely
aligned with the relevant substrate/product redox potentials, (b) demonstrating high
kinetic rate constants for substrate-to-product conversion, and (c) exhibiting high
heterogeneous electron transfer rates at the electrode interface. A representative
schematic of an electrocatalytic process is shown in Figure 1.

​ ​ ​ ​ 4
Figure 1. Diagrammatic illustration of a paired electrolysis relating cathodic CO2-to-CO
conversion catalyzed with a Re complex and anodic synthesis of a 2-substituted
benzimidazole mediated by ceric ammonium nitrate

2.2 Carbon Dioxide Coordination Chemistry

2.2.1 Electronic Structure of CO₂

​The electrocatalytic reduction of carbon dioxide is initiated through specific interactions


between the reactive catalyst species and the CO₂ molecule, leading to its activation.
Carbon dioxide, a 16-electron molecule, belongs to the D∞h point group and adopts a
linear geometry in its ground or gaseous state. Despite its nonpolar character and high
thermodynamic stability, the molecule features two polar C=O bonds and two orthogonal
π-symmetry orbital systems. Historically considered a weak ligand, recent investigations
have revealed that CO₂ can coordinate in diverse binding modes and through multiple
atomic sites in its metal complexes.​

The carbon atom, associated with the LUMO of the molecule, exhibits Lewis acidic
behavior and functions as an electrophilic center. Conversely, the oxygen atoms
contribute HOMO orbitals and act as weak Lewis bases, thus serving as nucleophilic
sites. Importantly, many catalytic CO₂ transformations proceed via cooperative
acid–base activation mechanisms, where both the carbon center and an oxygen atom
are involved in bonding interactions with a metal site.

The two C=O double bonds contain π-electrons capable of engaging with d-orbitals of
transition metals in line with the Dewar–Chatt–Duncanson model of bonding (Figure 2).

​ ​ ​ ​ 5
Upon population of the CO₂ LUMO via electron transfer, the molecule adopts a bent
geometry at its lowest energy conformation. This structural change is accompanied by
elongation of the C–O bonds, a reduction in bond angle, and accumulation of negative
charge—characteristics that reflect the extent of adsorption and degree of activation.

Figure 2. Schematic representation of (a) the Dewar–Chatt–Duncanson model and


(b)the reversed Dewar–Chatt–Duncanson model41 Reprinted with permission from
Goedecke, et al., The Dewar–Chatt–Duncanson model reversed—Bonding analysis of
group-10 complexes [(PMe3)2M–EX3] (M = Ni, Pd, Pt; E = B, Al, Ga, In, Tl; X = H, F,
Cl, Br, I), Can. J. Chem. 87 (10)

2.2.2 Binding Modes of Carbon Dioxide

Carbon dioxide can coordinate to transition metals through four principal binding
strategies involving carbon and oxygen atoms, as illustrated in Figure 3 [ref. 39]. These
coordination modes rely on significant electron transfer from metal orbitals into the
antibonding molecular orbitals of CO₂. This interaction is often stabilized by
supplementary weak coordination between one or both oxygen atoms of CO₂ and an
adjacent Lewis acidic site within the metal’s coordination sphere. Such interactions are
more favorable with electron-rich (nucleophilic) metal centers.

The η¹-CO₂ complexes, where the molecule binds via a single atom (usually carbon),
tend to be unstable and typically require inert-atmosphere techniques such as Schlenk
lines or gloveboxes to isolate them successfully. As documented by Herskovitz and

​ ​ ​ ​ 6
co-workers, formation of η¹-CO₂ complexes (especially with rhodium and iridium) often
necessitates elevated CO₂ pressure. Interestingly, the coordination does not involve
ligand substitution but occurs through direct binding.

In the (C,O) coordination pattern, bonding consists of σ-donation from a CO₂ orbital to a
vacant metal orbital, combined with π-back-donation from a filled metal d-orbital
(typically dₓᵧ) into an empty π* orbital of CO₂. Aresta and Nobile reported an η²-complex,
[Ni(CO₂)(PCy₃)₂], formed by the reaction of Ni(PCy₃)₃ or [Ni(PCy₃)₃]N₂ with atmospheric
CO₂ in toluene, yielding a solvate.

The η¹-(O) coordination mode—where CO₂ binds through a single oxygen—is more
typical for electron-deficient metal centers, and the geometry of the CO₂ may remain
linear or slightly bent. The (O,O) binding motif resembles a carboxylate, exhibiting ionic
character, and is frequently observed with alkali or alkaline-earth metals, as well as on
metal surfaces during CO₂ adsorption.
Additionally, Karsch synthesized an η²-CO₂ complex via reaction of Fe(PMe₃)₄ with CO₂
in pentane, which also yielded a carbonate-containing species, Fe(PMe₃)₃(CO)(CO₃).
More recently, Komiya et al. provided a comprehensive structural analysis of
Fe(CO₂)(depe)₂, reinforcing the characterization of the earlier η² complex (Karsch).

Figure 3. CO2 coordination modes to a single metal center 39-42

​ ​ ​ ​ 7
2.3 Recent Advances in CO₂ Reduction Using Molecular Catalysts

2.3.1 Iron-Based Molecular Catalysts

As highlighted earlier, the electrochemical conversion of CO₂ into hydrocarbon fuels is a


pivotal technology for addressing climate challenges and enhancing energy
sustainability [48]. For example, the Fischer–Tropsch process converts syngas (CO and
H₂) into liquid fuels. Molecular catalysts—both in homogeneous and heterogeneous
forms—are widely utilized under electrochemical and photochemical conditions for CO₂
transformation [49].
In recent years, iron-based porphyrins and metal-organic frameworks have drawn
considerable interest for their effectiveness in catalyzing electrochemical and
photoelectrochemical CO₂ reduction [50]. Advances in catalyst design, including
synthetic modifications to improve activity and selectivity, have made homogeneous
electrocatalysis a prominent strategy. For instance, research by Taheri et al. [51]
demonstrated that an iron-based molecular electrocatalyst could selectively reduce CO₂
to format in aqueous media, operating at low overpotential with high current density and
achieving a Faradaic efficiency of 96%.
Mechanistic insights, derived from cyclic voltammetry, also allowed identification and
structural characterization of a key intermediate—an iron-hydride species formed during
the reduction process. Combined electrochemical studies in MeCN/H₂O (95:5) and in
fully aqueous conditions suggest a catalytic pathway in which the iron complex is first
reduced from the 1⁻ state to the 1²⁻ state, followed by protonation to generate the
hydride intermediate (H–1)⁻. Subsequent selective interaction between (H–1)⁻ and CO₂
results in C–H bond formation, ultimately yielding formate as the final product (see Fig.
4).

​ ​ ​ ​ 8
Figure 4. Projected Mechanism for Reduction of CO2to Format by 1-in the Existence of
Protons51. Reprinted with permission from Taheri et al., An iron electrocatalyst for
selective reduction of CO2 to form in water: including thermochemical insights, ACS
Catalysis 5 (12) (2015) 7140-7151. ACS Catal. 2015, 5, 7140−7151. Copyright (2015)
American Chemical Society

Ambre et al. [52] investigated the rational design of iron porphyrin catalysts for
enhanced and selective electrochemical CO₂ reduction to CO. In their study, they
modified the porphyrin framework by introducing ester functional groups at para, meta,
and ortho positions of the aryl rings, resulting in three distinct catalysts: Fe-pE, Fe-mE,
and Fe-oE. These molecularly engineered variants were evaluated for their catalytic
performance in CO₂-saturated electrolyte solutions supplemented with 2 mol·L⁻¹ water.
Notably, Fe-mE and Fe-oE exhibited high Faradaic efficiencies for CO production—65%
and an impressive 98%, respectively—after 2 hours of bulk electrolysis. Particularly, the
meta-substituted Fe-mE was highly selective for CO formation, with negligible hydrogen
evolution. In contrast, the para-substituted Fe-pE predominantly produced H₂ (84%
Faradaic efficiency), underscoring how positional effects of substituents can steer
product selectivity.

Cyclic voltammetry (CV) analysis in DMF revealed distinct redox signatures for each
porphyrin derivative. Among them, Fe-oE exhibited a significantly more negative
Fe(I)/Fe(0) potential than Fe-mE and Fe-pE, likely due to a stronger molecular dipole

​ ​ ​ ​ 9
effect. Consequently, Fe⁰ generated from Fe-oE is proposed to be the most potent
reductant among the series. These findings are in agreement with the work of Savéant
and co-workers [53–55], who have shown that the Fe⁰ species facilitates CO₂ reduction
at the Fe(I)/Fe(0) couple under CO₂ atmosphere.

Additionally, Rao et al. [56] explored light-driven CO₂ reduction to methane using a
functionalized iron tetraphenylporphyrin complex bearing trimethylammonio groups.
This catalyst, already known for its high performance in converting CO₂ to CO [57–59],
demonstrated the ability to catalyze a full eight-electron reduction to methane under
visible light irradiation at ambient conditions. The reaction was carried out in an
acetonitrile medium containing a photosensitizer and sacrificial electron donor,
maintaining stability over several operational cycles.

Direct photoreduction yielded CO as the main product; however, when a two-step


process was employed—first converting CO₂ to CO, followed by CO reduction to
CH₄—the system achieved methane selectivity of ~82% and a quantum yield of 0.18.
Mechanistic insights (Fig. 5) suggest the involvement of a formyl intermediate, which
may be stabilized through electrostatic interactions between the positively charged
trimethylammonium groups and the partial negative charge on the metal-bound CHO
species [60, 61].

Figure 5. Illustration of the mechanism for CO2 reduction to CH456. Reprinted with
permission from Rao et al., Visible-light-driven methane formation from CO2 with a

​ ​ ​ ​ 10
molecular iron catalyst, Nature 548 (2017) 74-77. Springer Nature and Copyright (2017)
Clearance Center

The proposed reaction pathway involves the transformation of an Fe³⁺ porphyrin


precursor (top left) into a catalytically competent Fe⁰ species through a three-electron
reduction. This Fe⁰ center subsequently drives CO₂ conversion, during which Fe⁺ is
regenerated via electron donation from an excited-state photosensitizer (depicted in the
right-hand catalytic cycle). The resulting CO intermediate binds to Fe²⁺ and undergoes
further reduction—requiring six electrons supplied by the excited sensitizer along with
protons—to ultimately yield methane, potentially via a hypothesized Fe⁺–formyl
(Fe⁺CHO) intermediate (left-hand cycle).

Recently, Pan and colleagues⁶² explored the reactive centers involved in CO₂
electroreduction on nitrogen-doped, atomically dispersed Fe and Co catalysts. Their
experimental findings indicated that Fe–N₄ coordination environments exhibit superior
catalytic performance over Co–N₄ analogs within M–N–C frameworks, as evidenced by
higher Faradaic efficiencies, lower overpotentials, and notable activity.

Complementary theoretical studies (Figure 6) identified edge-bound M–N₂+₂–C₈


configurations, bridging dual armchair-like graphitic domains, as key active motifs in
CO₂ reduction. The transition metal sites, adjacent carbon atoms, flexible bonding
frameworks, and nearby nitrogen donors were recognized as critical *CO and *OH
adsorption regions that facilitate C–O bond dissociation during the CO₂ reduction
process.

​ ​ ​ ​ 11
Figure 6. Atomic structure of M−N4−C10 and M−N2+2−C8 (M=Fe or Co) active
sites.(b) Calculated free energy evolution of CO2 reduction to CO on M−N2+2−C8 sites
under useful electrode potential (U) of 0 V and −0.6 V. (c) The initial and final state for
the COOH dissociation reaction on M−N4−C10 and M−N2+2−C8 sites. In the figure,
The gray, blue, yellow, red, and white balls represent C, N, M, O, and H atoms,
respectively62. Reprinted with permission from Pan et al., Unveiling Active Sites of CO2
Reduction on Nitrogen-Coordinated and Atomically Dispersed Iron and CobaltCatalysts,
ACS Catal. 8 (4) (2018) 3116-3122. Copyright (2018) American Chemical Society.

Nichols et al.⁶³ explored how second-sphere spatial arrangements influence the catalytic
behavior of a conserved iron porphyrin framework in electrochemical CO₂ reduction. To
this end, they synthesized four positional isomers (Figure 7) by varying the placement of
an amide group in the second coordination sphere—either ortho or para, and either
near (proximal) or far (distal) from the porphyrin macrocycle.

Under CO₂ atmosphere and in the presence of phenol (used as a proton donor), cyclic
voltammetry was employed to probe the electrocatalytic activity of each complex. In
comparison, when 100 mM phenol was introduced, Fe–ortho-1-amide showed a slightly
more positive catalytic onset potential than Fe–ortho-2-amide. Both ortho variants
exhibited substantially improved activity over their para-substituted counterparts and the
unmodified Fe–TPP.

​ ​ ​ ​ 12
To gain further insight into the CO₂ reduction mechanism by these amide-functionalized
systems, kinetic data obtained via foot-of-the-wave analysis (FOWA) were evaluated
with respect to varying phenol and CO₂ concentrations. All four catalysts displayed
first-order kinetics with phenol under pseudo-first-order conditions. Among them,
Fe–ortho-2-amide presented the highest observed rate constants. However, its
performance plateaued at elevated phenol levels, likely due to catalyst inhibition or local
CO₂ depletion. Fe–ortho-1-amide followed closely in activity, trailed by
Fe–para-2-amide, while Fe–para-1-amide demonstrated the lowest reactivity.

Catalytic performance was further assessed by plotting overpotential (η) versus


log(TOF) to produce Tafel-style curves⁶⁴. Efficient catalysts appear in the upper-left of
these plots, reflecting high turnover frequencies at low overpotentials. Fe–ortho-2-amide
consistently showed superior TOFs across the entire overpotential range compared to
Fe–ortho-1-amide, which in turn outperformed Fe–TPP and both para variants. The
curve peaks represent maximum TOF achievable under high-driving-force conditions.

Importantly, differences in catalytic onset potentials (E₀cat) among the various


complexes were critical in pinpointing optimal second-sphere amide placements. Prior
studies⁶⁵,⁶⁶ have shown that E₀cat is strongly influenced by electron-donating or
-withdrawing substituents on porphyrin aryl groups. A linear trend typically emerges
between log(TOF ₐₓ) and E₀cat, where greater activity corresponds to more negative
E₀cat values. However, deviations can occur when second-sphere effects—such as
hydrogen bonding or electrostatic interactions—enhance or impair catalytic
performance, as observed in earlier reports⁶⁷.

Figure 7. Positional isomers of amide-functionalized iron tetraphenylporphyrins63

​ ​ ​ ​ 13
2.3.2 Zn-based Metal Complex Molecular Catalyst

Zinc-based catalytic systems have attracted attention due to the metal's low cost, high
availability, and low toxicity. However, reports on zinc-catalyzed hydrosilylation of CO2
remain scarce. Notable examples include two cationic N-heterocyclic carbene (NHC)
zinc complexes, a dicationic NHC-stabilized zinc hydride cluster, and mononuclear
complexes such as \[Tptm]ZnH and \[TismPriBenz]ZnH. These systems have
demonstrated catalytic activity in CO2 transformations.

Donovan et al. reported the synthesis and electrochemical evaluation of two novel Zn(II)
complexes: dichloro\[phenyldi(2-pyridyl)phosphine-κ2-N,N’]zinc(II) and
dichloro\[diphenyl-(2-pyridyl)phosphine-κ1-N]zinc(II). Their electrochemical profiles were
studied via cyclic voltammetry. To assess their reactivity with CO2, 0.20 mol L⁻¹ THF
solutions of each complex were exposed to CO2. Solids were collected post-reaction
and analyzed via IR spectroscopy. The dichloro\[phenyl di(2-pyridyl)phosphine]zinc(II)
complex showed no spectral changes, indicating a lack of stable CO2 adduct formation.
Conversely, dichloro\[diphenyl-(2-pyridyl)phosphine]zinc(II) displayed a new IR band at
1726 cm⁻¹ after 30 minutes, suggesting the formation of a CO2-derived intermediate
(2-CO2).

Further analysis via cyclic voltammetry indicated that the latter complex significantly
reduces the overpotential required for CO2 reduction. In a CO2 environment, an anodic
shift (\~0.6 V) in the reduction onset potential was observed, consistent with a CE
(chemical-electrochemical) mechanism. Mass spectrometry of the electrolysis cell
headspace before and after bulk electrolysis confirmed CO generation. Prior to
electrolysis, the dominant MS peak corresponded to CO2 (m/z 44). Post-electrolysis,
the main peak shifted to m/z 28, indicating CO evolution. This demonstrates that the
dichloro\[diphenyl-(2-pyridyl)phosphine]zinc(II) complex functions as the first molecular
Zn complex capable of lowering the CO2-to-CO reduction overpotential by about 0.6 V,
with the reaction proceeding efficiently on a glassy carbon electrode—an advantage
over more costly Pt or Pd materials.

​ ​ ​ ​ 14
In a related study, Wu et al. reported a heterogeneous zinc–porphyrin complex, zinc(II)
5,10,15,20-tetraphenylporphyrin (PorZn), as an effective electrocatalyst. This system
delivered a Faradaic efficiency of 95% for CO production at −1.7 V vs. SHE in a mixed
organic/aqueous electrolyte and a turnover frequency of 14.4 s⁻¹ per active site. PorZn
electrodes were investigated under CO2-saturated conditions using 0.1 mol L⁻¹ TBAPF6
in a DMF/H2O solvent. At potentials more positive than −1.4 V, H2 was the only
product, while CO production initiated at −1.4 V with a 22% Faradaic efficiency.
Maximum CO efficiency (95%) was achieved at −1.7 V with a current density of \~2.1
mA/cm². This performance remained stable for at least 4 hours.

Given zinc's established role in CO2-to-CO electroreduction, it was necessary to


exclude the formation of metallic Zn during catalysis. In situ and operando X-ray
absorption spectroscopy (XAS) was used to monitor changes in the Zn center. XANES
data showed no detectable changes in the Zn oxidation state as the potential varied
from OCV (−1.7 V vs. SHE) to +0.2 V. Extended X-ray absorption fine structure
(EXAFS) analysis showed minor variations in Zn coordination and bond lengths, likely
due to ligand reduction or substrate binding. These results suggest that the Zn center
plays a critical structural role while remaining redox-innocent.

Cyclic voltammetry of PorZn in CO2-saturated electrolyte revealed a reduction wave at


−1.4 V vs. SHE, corresponding to the onset of CO production. In Ar-saturated
electrolyte, a similar reduction wave was observed, likely linked to protonation events.
However, significant differences in anodic behavior were noted. Under CO2, a single
weak anodic wave (~~−1.45 V) was observed. Under Ar, three anodic peaks (~~−1.53
V, −1.17 V, −0.43 V) appeared, consistent with oxidation and deprotonation of reduced
porphyrin ligands, as reported in prior studies of zinc(II) tetraphenylporphyrin.

To gain mechanistic insight, Wu and colleagues chemically reduced PorZn under inert
conditions using one or two equivalents of sodium naphthalene (NaNap, \~−2.4 V vs.
SHE). The reduced species exhibited UV–Vis bands at 710, 820, and 920 nm, indicative

​ ​ ​ ​ 15
of transient reduced zinc–porphyrin intermediates. Exposure to air led to rapid
reoxidation and regeneration of the original PorZn complex. While complete structural
characterization of these species remains elusive due to instability, the findings confirm
PorZn as the first structurally defined Zn-based molecular electrocatalyst for CO2
reduction to CO. This catalyst operates via a redox-inert Zn(II) center, distinguishing it
from most known transition-metal systems and offering a promising pathway for
selective CO2-to-CO electroreduction.

Figure 8: (Left) Electrochemical CO2 reduction catalyzed by the PorZn electrode in 0.1
mol L-1 TBAPF6 DMF/H2O solution. (A) CO Faradaic efficiencies and CO partial
current densities at different potentials averaged from three measurements. (B) CO
Faradaic efficiencies and total current densities after 5, 60, 120, 180, and 240 min of
electrolysis at −1.7 V vs. SHE. (right) (A) Zn K-edge XANES spectra and (B) Fourier
transforms of Zn K-edge EXAFS spectra of the PorZn catalyst electrode at different
potentials (V vs. SHE)75. Reprinted with permission from Wu et al., Electroreduction of
CO2 Catalyzed by a Heterogenized Zn-Porphyrin Complex with a Redox-Innocent Metal
Center, ACS Central Science 3 (8) (2017) 847-852. Copyright (2017) American
Chemical Society.

​ ​ ​ ​ 16
Wang et al.87 investigated the electrochemical CO2 reduction in aqueous solutions
using synthesized ZIF–8 nanomaterials, which possess unique properties such as a
large surface area, uniform pore size, well-defined morphology, and strong coordination.
Upon CO2 electrochemical reduction, ZIF–8 nanocomposites were found to be highly
effective, with the metal nodes of ZIF–8 proposed as active catalytic sites. The
electrochemical reduction of CO2 using ZIF–8 nanomaterials supported on glassy
carbon (GC) electrodes was evaluated via cyclic voltammetry (CV) in a 0.5 mol L⁻¹ NaCl
aqueous solution. The CV results for different ZIF–8 materials in both Ar and CO2
atmospheres showed increased current and a positive shift in the onset potential of
CO2 catalysis. Specifically, the positive potential shift was approximately 400 mV for
ZIF–8SO4, while ZIF–8AC exhibited slightly less current in CO2 compared to Ar.
Despite this, the current for ZIF–8AC was still distinguishable from the other two
materials (ZIF–8SO4 and ZIF–8NO3) in terms of surface loading uniformity.
Furthermore, electrolysis results showed that ZIF–8AC demonstrated high current for
hydrogen evolution, rather than CO2 reduction.

Controlled-potential electrolysis (CPE) was conducted under CO2 to examine product


distribution and Faradaic efficiency (FE). The potentials used ranged between –1.4 V
and –1.9 V vs. SCE, with gas chromatography (GC) and nuclear magnetic resonance
spectroscopy (NMR) applied for gaseous and liquid phase product analysis. Three main
observations were made: (i) all ZIF–8 materials were active for CO2 reduction, (ii) key
products were hydrogen (H₂) and carbon monoxide (CO), and (iii) only small amounts of
formate were found in the liquid phase. As the applied potential increased, the FE for
CO gradually increased until reaching a maximum, after which it decreased at higher
potentials. The CO Faradaic efficiencies achieved were as follows: ZIF–8SO4 [65.5%],
ZIF–8NO3 [69.8%], and ZIF–8AC [57.7%] at –1.8 V vs. SCE. Notably, ZIF–8SO4
exhibited the widest potential range suitable for CO production (–1.5 to –1.9 V vs. SCE).
Additionally, current densities for H₂ and CO production were examined at –1.8 V vs.
SCE, and ZIF–8SO4 showed the highest efficiency for CO2 reduction compared to
ZIF–8NO3 and ZIF–8AC.

​ ​ ​ ​ 17
Several synthesized ZIF–8 nanocomposites were found to be excellent CO2 reduction
catalysts. By adjusting the zinc source, the chemical reactivity of ZIF–8 could be
tailored, and ZIF–8SO4 was particularly effective in producing CO with high selectivity.
Furthermore, the electrolyte composition played a crucial role in achieving high CO
selectivity, with chloride (Cl⁻) anions enhancing CO2 reduction reactivity, likely due to
superficial anion exchange and the small hydrated ionic radius. These findings highlight
ZIF–8 materials as promising electrocatalysts for CO2 reduction.

2.3.3 Mn-based Metal Complex Molecular Catalyst


Manganese compounds commonly exhibit oxidation states of +2, +4, and +7, with the
Mn+7 state being stable but often reduced to Mn+2. The Mn+1 oxidation state is rare
but can appear in manganese-based organometallic complexes, such as the d6 MnI
tricarbonyl complexes, which are increasingly important in catalysis 88, 89. An example
is the fac-[Mn(bpy)(CO)3Br] complex, where the highest occupied molecular orbital
(HOMO) is dominated by a Mn 3d-orbital and the lowest unoccupied molecular orbital
(LUMO) is influenced by a diimine-based ligand. The electronic and photophysical
characteristics of MnI complexes can be further tuned by modifying the diimine ligand,
adjusting the LUMO energy. The use of manganese as a catalyst is motivated by its
chemical similarities to rhenium (same group, oxidation states, and geometries), with
manganese being 1.3 million times more abundant in Earth's crust than rhenium90.

Numerous reviews have explored transition metal and manganese-based systems for
CO2 reduction 91-94, including a recent review by Sinopoli et al.90, which highlighted
the effectiveness, features, and strategies of Mn carbonyl complexes in CO2 electro-
and photoreduction. Stanbury and colleagues95 also conducted a comprehensive
survey of Mn carbonyl systems, evaluating their activity, stability, and selectivity for CO2
reduction through electro-, photo-, or photo-electroreduction. Their review concluded
that Mn-based carbonyl complexes exhibit significant catalytic potential for CO2
reduction, rivaling their rhenium analogs in terms of turnover numbers (TON) and
turnover frequencies (TOF).

​ ​ ​ ​ 18
Fei et al.96 investigated the post-synthetic metallation of a robust Zr(IV)-based
metal-organic framework (MOF) with bpy metal-chelating linkers, resulting in isolated
Mn(bpy)(CO)3Br moieties within the MOF. When combined with [Ru(dmb)3]2+ as a
redox photosensitizer and 1-benzyl-1,4-dihydronicotinamide (BNAH) as a reducing
agent, the [UiO-67 Mn(bpy)(CO)3Br] system showed high activity and selectivity for the
photocatalytic reduction of CO2 to formate, achieving a turnover number (TON) of 110
after 18 hours of catalysis.

The proposed mechanism for the photocatalytic reaction is as follows: BNAH serves as
a reducing agent, reducing the excited Ru(II) photosensitizer and initiating the reaction.
The reduced photosensitizer then transfers electrons to the Mn catalyst, leading to the
formation of an absorbable Mn(0) species during catalysis. The large pores of UiO-67
enable efficient electron transfer between the Ru(II) photosensitizer and the Mn
complex. Additionally, TEOA may enhance the reaction by providing a proton and
electron (i.e., a hydrogen atom) during catalysis through a Hofmann-type degradation
process (Figure 9). It is still unclear whether TEOA interacts directly with the Mn center
during this process, but it likely aids in the binding of CO2 to the metal center, forming
an O-bound Re−OC(O)OCH2CH2NR2 complex, as observed in previous studies with
rhenium bipyridine catalyst 101. Fei and co-workers suggest that TEOA donates one
proton and one electron to the catalytic cycle, producing a Mn(I)−H complex, into which
CO2 can insert, forming a Mn(I)−OC(O)H complex. Formate (or formic acid after further
protonation) can then be released from the Mn center, regenerating the starting Mn(I)
complex. These conclusions are consistent with prior research on photosensitized
catalysis driven by sacrificial reducing agents 102-107.

​ ​ ​ ​ 19
Figure 9. Suggested mechanism for the formation of formate from the photocatalytic
reaction with UiO-67-Mn(bpy)(CO)3Br96.Reprinted with permission from Fei et al.,
Photocatalytic CO2 Reduction to Formate Using a Mn(I) Molecular Catalyst in a Robust
Metal–Organic Framework, Inorg.Chem. 54 (14) (2015) 6821-6828. Copyright(2015)
American Chemical Society.

According to Franco et al., the first pure organometallic fac-[MnI(CO)3(bis-MeNHC)Br]


complex has revolutionized the field by demonstrating exceptional electrocatalytic
CO2-to-CO reduction activity, achieving over 100 turnovers of CO with high Faradaic
efficiency (ηCO~95%) in anhydrous CH3CN. The catalyst displayed an outstanding
performance, surpassing 2100 turnovers per second (TOFmax), a result significantly
higher than those reported for other manganese-based catalysts. These findings
represent a major advancement in CO2 reduction catalysis, marking the importance of
this Mn-based system in the conversion of CO2 to useful products.
To gain deeper insights into the mechanism of this transformation, cyclic voltammetry
(CV) was employed, revealing the transformation of the [MnI(CO)3(bis-MeNHC)Br]
complex into a five-coordinate [Mn(CO)3(bis-MeNHC)]– core upon reduction. The
experimental CO stretching frequency of this intermediate demonstrated a pronounced
negative charge localized over the manganese atom, consistent with the redox-inert

​ ​ ​ ​ 20
nature of the bis-MeNHC ligand. The reduction process, as evidenced by the
Kohn-Sham orbital analysis, occurs primarily on the metal center, confirming the pivotal
role of the manganese center in the reduction process.

The reduced form of the catalyst, [Mn(CO)3(bis-MeNHC)]–, presents an open HOMO


orbital structure, which is highly accessible to undergo nucleophilic attack. This ability to
undergo nucleophilic attack is in stark contrast to [Mn(CO)3(py-MeNHC)]–, where the
reactivity is limited. The alteration of the catalyst’s properties due to the introduction of
bis-MeNHC instead of py-MeNHC exemplifies the significant influence that ligand
modifications can have on the catalytic efficiency and selectivity. The NHC-based
ligands enhance both the turnover frequency (TOFmax) and the selectivity for CO
production, demonstrating their superior catalytic performance compared to traditional
C^N ligand-based systems.

Franco and colleagues' groundbreaking work highlights the potential of bis-NHC-based


catalysts for the selective reduction of CO2 to CO, demonstrating remarkable efficiency
in aprotic organic solvents. This marks a departure from traditional bpy-based systems,
which typically exhibit inertness in the absence of a defined proton source. The study
not only establishes the efficacy of NHC-ligated manganese complexes but also sets a
new benchmark for electrocatalytic CO2 reduction, underscoring the importance of
ligand design and metal center optimization in achieving high-performance catalysts.

​ ​ ​ ​ 21
Figure 10. Proposed mechanism of CO2 reduction to CO by
fac-[MnI(CO)3(bis-MeNHC)Br] complex 108

Reuillard et al.115 stated the assemblage of the complex [MnBr(2,2'-bipyridine)(CO)3]


attached to a carbon nanotube electrode over a pyrene unit. Molecular catalyst check
allows electrocatalytic CO2 reduction during entire aqueous conditions with a catalytic
onset overpotential of η = 360 mV, and restrained potential electrolysis produced
turnovers at η = 550 mV greater than 1000. The product selectivity is adjustable by
catalyst amendment loading on the nanotube surface.

Figure 11. Proposed mechanism of fac-[MnBr(bpypyr)(CO)3] immobilized on a


CNT sidewall, concentration-dependent dimerization or Mn−H formation, and

​ ​ ​ ​ 22
intermediate-dependent reduction of CO2 to CO or HCOOH115
The [MnBr(bpy)(CO)₃] complex (where bpy = 2,2'-bipyridine) is a non-precious
metal-based model electrocatalyst for CO₂ conversion, owing to the structural versatility
and direct coordination of the bpy ligand. This Mn-based catalyst demonstrated notable
activity with a turnover frequency (TOF) reaching 480 s⁻¹ in acetonitrile. Mechanistic
studies have extensively explored how substituent modifications on the bpy framework
affect catalysis. The Mn complex has been immobilized on carbon nanotubes (CNTs)
using Nafion, polymerized onto p-type silicon, and anchored to TiO₂ surfaces via
phosphonate groups. Among these, the CNT-based system exhibited a record turnover
number (TON) of 112 in MeCN. Notably, the Mn species tends to dimerize on the
electrode surface, a feature supported by spectroelectrochemical UV–Vis studies.
Dimer formation was also confirmed in solution upon reduction. Nonetheless, catalytic
efficiency in aqueous media remains modest, with reported TONs not exceeding 101.
The integration of a pyrene moiety enabled robust attachment to CNTs, facilitating
electrochemical studies of CO₂ reduction both in homogeneous MeCN/H₂O (5%)
systems and under aqueous heterogeneous conditions. Constant potential electrolysis
(CPE) and cyclic voltammetry (CV) experiments were used to assess catalytic behavior.
Surface concentration of the Mn catalyst strongly influenced product distribution
between CO and formate. The formation and role of various intermediates were further
examined through in situ UV–Vis and attenuated total reflection (ATR) infrared
spectroelectrochemistry. Differentiating between CO and HCOO⁻ formation pathways at
varying catalyst loadings is crucial for understanding selectivity and mechanistic
insights.

2.3.4 Nickel-Based Molecular Catalysts

Nickel, a first-row transition metal from group VIIIB, is recognized as a highly promising
alternative to noble metals such as palladium and platinum in the development of
molecular catalysts. Its accessibility to multiple oxidation states (Ni⁰, Niᴵ, Niᴵᴵ, Niᴵᴵᴵ, and
Niᴵⱽ) enables diverse redox behavior essential for catalytic processes. In CO₂ reduction
reactions, catalytic activity generally relies on the reduction of Niᴵᴵ to Niᴵ, which involves

​ ​ ​ ​ 23
a shift in geometry from a square planar (preferred for Niᴵᴵ) to a tetrahedral structure
compatible with Niᴵ. A similar geometrical reorganization occurs during oxygen evolution
reactions (OER), where oxidation from Niᴵᴵ to Niᴵᴵᴵ promotes transition into tetragonal or
octahedral forms. These adaptable electronic and structural characteristics allow ligand
frameworks to be rationally designed to support and stabilize active Ni centers, resulting
in efficient catalytic systems.

In nature, nickel plays a central role in enzymatic systems such as [NiFe] hydrogenases
and [NiFe] carbon monoxide dehydrogenases (CODHs), responsible for reversible H⁺/H₂
and CO₂/CO interconversion, respectively. These biological precedents have inspired
the use of nickel complexes for artificial catalysis, particularly in energy-relevant
transformations. Although numerous reviews have emphasized catalysts based on
earth-abundant metals for hydrogen evolution (HER), OER, and CO₂ reduction,
comprehensive literature focusing on nickel-based molecular catalysts remained scarce
until a recent review by Wang and Jia-Wei et al., which thoroughly surveyed Ni-based
systems for water splitting and CO₂ conversion. Among these, nickel cyclam-based
complexes have emerged with remarkable efficiency and selectivity in CO₂ reduction,
while others show moderate performance and are often limited to non-aqueous
solvents.

A notable advancement was reported by Niu and co-workers, who developed a porous,
crystalline nickel-organic hybrid catalyst via a photochemical synthesis route. This
material, built from a Ni(TPA/TEG) ligand framework, exhibited high defect density and
was active for the photoreduction of CO₂ to CO with an impressive production rate of
~1.6 × 10⁴ μmol h⁻¹ g⁻¹. The process was nearly 100% selective for CO formation with
no measurable H₂ detected over 6 hours of irradiation. Among the tested materials, the
spongy Ni(TPA/TEG) catalyst displayed the highest CO output—95.2 mmol within 2
hours—translating to a CO generation rate of 15,866 μmol h⁻¹ g⁻¹. Over a 6-hour
experiment, the total CO yield reached 136.9 μmol, corresponding to a turnover number
(TON) of 11.5.

​ ​ ​ ​ 24
Varying the amount of catalyst revealed a direct proportionality between catalyst mass
and CO yield up to a point, beyond which the CO production rate per gram decreased.
This suggests that increased catalyst loading may impede electron transfer from the
photosensitizer [Ru(bpy)₃]²⁺, identifying this step as potentially rate-limiting. The catalyst
also demonstrated excellent recyclability and retained both structure and performance
after multiple 2-hour cycles and up to 24 hours of continuous use. Isotopic tracing using
¹³CO₂ and subsequent GC-MS analysis confirmed that the generated CO indeed
originated from CO₂, with a diagnostic m/z signal at 29 indicative of ¹³CO.

Based on these findings, a mechanistic model was proposed wherein [Ru(bpy)₃]²⁺ is


photoexcited and subsequently quenched by the sacrificial donor TEOA, forming a
reduced Ru species that transfers electrons to the nickel catalyst. CO₂ adsorbed onto
the catalyst is then reduced to CO. Further investigations into downstream products
showed the potential formation of liquid fuels such as formic acid, acetic acid, and
ethanol through proton-coupled multi-electron transfer reactions. At near-neutral pH
(~8), pathways were identified for the generation of HCOOH, CH₃COOH, and
CH₃CH₂OH via one-, four-, and eight-electron processes, respectively.

Mechanistically, CH₃COOH formation likely proceeds via sequential hydration of CO to


CHO, then CHOH, CH₂OH, and CH₃OH, followed by coupling with surface-bound CO.
Ethanol formation may involve CHO dehydroxylation to CH, then further protonation to
CH₂ and CH₃, with subsequent C–C bond formation. In contrast, under highly basic
conditions (pH ~13), hydroxide ions drive selective formation of formic acid, though they
may suppress deeper reduction steps and limit acetic acid and ethanol production.
CH₃OH presence under basic conditions suggests reduced C–C coupling efficiency,
highlighting key considerations for designing next-generation CO₂/CO reduction
catalysts

​ ​ ​ ​ 25
Figure 12. Proposed mechanisms for the photocatalytic reduction of CO2 to CO and of
CO to other liquid products. Visible light reduction of the photosensitizer [Ru(bpy)3]2+,
which transfers an electron to the Ni(TPA/TEG) catalyst to convert CO2 to CO and to
Ni(TPA/TEG)-(Ag/Rh) catalysts for the generation of HCOOH, CH3COOH, and
CH3CH2OH from further reduction of CO164.

Kuehnel and colleagues examined a group of nickel terpyridine-based complexes that


self-assemble and serve as molecular catalysts for CO₂-to-CO conversion in
non-aqueous media. By anchoring these complexes onto CdS quantum dots (QDs),
they enabled the catalysts to operate in fully aqueous systems and demonstrated
photocatalytic CO₂ reduction with over 90% selectivity under UV-filtered simulated solar
irradiation (AM 1.5G, 100 mW cm⁻², λ > 400 nm, at pH 6.7). The Ni complexes were
introduced in situ by adding a stock solution to a QD suspension in an aqueous
triethanolamine (TEOA, 0.1 mol L⁻¹) environment. The resulting solvent composition for
[Ni(terpy)₂]²⁺ and [Ni(terpyS)₂]²⁺ was 99:1 (H₂O:CH₃CN), and slightly higher for
[Ni(terpyC)₂]²⁺ and [Ni(terpyP)₂]²⁺ at 99.5:0.5.

Catalyst attachment was verified by UV–Vis analysis, which showed a dependency of


surface loading on the type of anchoring group (Figure 13A). Among the derivatives,
[Ni(terpyS)₂]²⁺ with a thiol functionality showed the most efficient surface attachment,
while others demonstrated reduced binding affinities (Figure 13B). UV–Vis and ATR-IR
spectra confirmed that the molecular integrity of [Ni(terpyS)₂]²⁺ remained intact when
immobilized on CdS QDs supported by mesoporous SnO₂ (Figure 13C). Moreover,
cyclic voltammetry affirmed that the immobilized complex retained its original redox

​ ​ ​ ​ 26
characteristics, validating its functional stability on the quantum dot interface (Figure
13D). Additionally, transmission electron microscopy revealed that catalyst
immobilization had no discernible impact on the structural morphology of the QDs.

Figure 13. Hybrid photocatalyst assembly from CdS QDs and [Ni(terpyX)2]
2+. A) Difference in UV-vis absorption of a [Ni(terpyS)2]
2+ solution before and after stirring with CdS QDs; B) Adsorption efficiency of different
[Ni(terpyX)2] 2+complexes; C) UV-vis spectra of CdS-[Ni(terpyS)2]
2+ hybrid photocatalyst immobilized on a mesoporous SnO2 electrode
and comparison with CdS-BF4 and [Ni(terpyS)2]2+ in solution (spectra scaled and
stacked for clarity); D) Cyclic voltammetry of CdS-[Ni(terpyS)2]
2+ photocatalyst immobilized on a SnO2 electrode172. Reprinted with permission from
Kuehnel et al., Selective Photocatalytic CO2 Reduction in Water through Anchoring of a
Molecular Ni Catalyst on CdS Nanocrystals, J. Am. Chem. Soc. 139 (21) (2017)
7217-7223. Copyright (2017) American Chemical Society.

The photocatalytic efficiency of CdS–[Ni(terpyX)₂]²⁺ hybrid assemblies was assessed


under CO₂-saturated aqueous conditions using simulated solar irradiation and
triethanolamine (TEOA) as a sacrificial electron donor. In contrast to the unmodified
CdS–[Ni(terpy)₂]²⁺ system, which predominantly generated H₂ with only minimal CO

​ ​ ​ ​ 27
detected, the derivatives equipped with anchoring moieties exhibited significantly
enhanced CO₂ reduction performance. Among these, the thiol-functionalized complex
[Ni(terpyS)₂]²⁺ displayed superior catalytic activity and selectivity, achieving a CO
selectivity of 92.2% after 4 hours, in comparison to just 10.2% and 3.9% for
[Ni(terpyP)₂]²⁺ and [Ni(terpyC)₂]²⁺, respectively.

Interestingly, this observed product selectivity did not align with trends in homogeneous
electrocatalysis but rather corresponded with the adsorption efficiencies of the
complexes on CdS. This outcome suggests that the interaction between the molecular
catalyst and the semiconductor interface is pivotal to effective photocatalytic behavior in
aqueous conditions. Under optimized conditions, approximately 20 catalytic turnovers
were reached over 24 hours of visible light exposure with the CdS–[Ni(terpyS)₂]²⁺
catalyst. The CO selectivity remained high (>90%) for the initial 8 hours but began to
diminish thereafter, favoring increased H₂ production by the 24-hour mark.
ICP-OES analysis of quantum dots recovered from the reaction medium revealed that
this shift in selectivity was associated with a progressive detachment of [Ni(terpyS)₂]²⁺
from the CdS surface. Notably, the integrity of the CdS–BF₄⁻ nanoparticles remained
intact throughout the process. Minor redshifts in UV–Vis absorption spectra indicated
negligible aggregation and no major photocorrosion.

Furthermore, reintroducing fresh [Ni(terpyS)₂]²⁺ after 20 hours restored CO production


while reducing H₂ evolution, whereas supplementing with Ni(BF₄)₂ alone only promoted
hydrogen formation. Reducing the initial ratio of catalyst to QDs decreased CO
selectivity but had little influence on the total number of CO turnovers relative to the
amount of [Ni(terpyS)₂]²⁺ present. These findings suggest a TON_CO value of
approximately 20 as a practical stability limit for this catalyst system.

3. Conclusions and future perspective

This general overview of the recent research on metal complex molecular catalysts for
CO2 reduction shows their contributions to the growing and dynamic field that is gaining

​ ​ ​ ​ 28
new insights into science at an ever-increasing rate. In order to increase metal
complexes' molecular catalysts application, it is imperative to overcome their poor
stability which is a significant challenge. We suggest that one of the possible means to
achieve this stability could be to ensure complexes are heterogeneous, through their
substrate’s immobilization. Notwithstanding, the limited examples we have presented in
this review, (e.g., porous structures, or covalent grafting on electrodes through ligand
functionalization), suggest that the concept is realizable. Succinctly, the advancing
knowledge acquired in several reduction mechanisms might give-way for more efficient
selective and prolonged metal complex molecular catalysts development. Additionally,
most metal complexes molecular catalysts are light sensitive and suffer from
photostability issues in any of these two ways; (i) by ligand photodissociation reactions
and (ii) photoisomerization. This particular problem is an important obstacle towards the
efficient metal complex photocatalysts for CO2 reduction advancement. Unarguably,
more improvements are still required, and other pathways for CO2 catalytic reduction
may perhaps offer ample fruitful opportunities.

5. Acknowledgments

It gives me great pleasure to find an opportunity to express my deep and sincere


gratitude to my course coordinator Dr Manas Kumar Biswas. I do very respectfully
recollect his constant encouragement,kind attention and keen interest throughout
the course of report writing on assigned topics. I am highly indebted to him for the
way he modeled and structured the format of report writing with his valuable tips and
suggestions. I am extremely grateful to the Department of chemistry, Asutosh
College, for extending all the facilities that enabled me to make the report.

I humbly extend my sense of gratitude to other faculty members, laboratory staff,


library staff and administration of this Institute for providing me with their valuable
help and time with a congenital working environment. Last but not least, I would like

to convey my heartiest thanks to all my classmates who have helped me with their

valuable suggestions during the drafting of the report.

​ ​ ​ ​ 29
6. References

1.​ Asadi, M., Kim, K., Liu, C., Addepalli, A. V., Abbasi, P., Yasaei, P., Phillips, P.,
Behranginia, A., Cerrato, J. M., Haasch, R., Zapol, P., Kumar, B., Klie, R. F.,
Abiade, J.,Curtiss, L. A., Salehi-Khojin, A., Nanostructured transition metal
dichalcogenide electrocatalysts for CO2 reduction in ionic liquid, Science 353
(6298) (2016) 467-470.https://doi.org/10.1126/science.aaf4767.
2.​ Collin, J. P., Sauvage, J. P., Electrochemical reduction of carbon dioxide
mediated by molecular catalysts, Coord. Chem. Rev. 93 (2) (1989) 245-268.
https://doi.org/10.1016/0010-8545(89)80018-9.
3.​ Savéant, J.-M., Molecular catalysis of electrochemical reactions. Mechanistic
aspects, Chem. Rev. 108 (7) (2008) 2348-2378.
https://doi.org/10.1021/cr068079z.
4.​ Takeda, H., Cometto, C., Ishitani, O.,Robert, M., Electrons, photons, protons and
earth-abundant metal complexes for molecular catalysis of CO2 reduction, ACS
Catal. 7 (1) (2017) 70-88. https://doi.org/10.1021/acscatal.6b02181.
5.​ Li, L., Yan, J., Wang, T., Zhao, Z. J., Zhang, J., Gong, J., Guan, N., Sub-10 nm
rutile titanium dioxide nanoparticles for efficient visible-light-driven photocatalytic
hydrogen production, Nat. Commun. 6 article number: 5881
(2015).https://doi.org/10.1038/ncomms6881.
6.​ Wang, J. C., Zhang, L., Fang, W. X., Ren, J., Li, Y. Y., Yao, H. C., Wang, J. S., Li,
Z. J., Enhanced photoreduction CO2 activity over direct z-schem α-Fe2O3/Cu2O
heterostructures under visible light irradiation, ACS Appl. Mater. Interfaces 7 (16)
(2015) 8631-8639. https://doi.org/10.1021/acsami.5b00822.
7.​ Hisatomi, T., Kubota, J., Domen, K., Recent advances in semiconductors for
photocatalytic and photoelectrochemical water splitting, Chem. Soc. Rev. 43
(2014) 7520-7535.https://doi.org/10.1039/C3CS60378D.
8.​ Ma, Y., Wang, X., Jia, Y., Chen, X., Han, H., Li, C., Titanium dioxide-based
nanomaterials for photocatalytic fuel generations, Chem. Rev.114 (19) (2014)
9987-10043.https://doi.org/10.1021/cr500008u.

​ ​ ​ ​ 30
9.​ Kang, D., Kim, T. W., Kubota, S. R., Cardiel, A. C., Cha, H. G., Choi, K. S.,
Electrochemical synthesis of photoelectrodes and catalysts for use in solar water
splitting, Chem. Rev. 115 (23) (2015)
12839-12887.https://doi.org/10.1021/acs.chemrev.5b00498.
10.​Tong, H., Ouyang, S., Bi, Y., Umezawa, N., Oshikiri, M., Ye, J.,
Nano-photocatalytic materials: possibilities and challenges, Adv. Mater.24 (2)
(2012) 229-251. https://doi.org/10.1002/adma.201102752.

​ ​ ​ ​ 31

You might also like