ArticlePDF Available

Recent Progress of Hydrogenation and Hydrogenolysis Catalysts Derived from Layered Double Hydroxides

Authors:

Abstract and Figures

Layered double hydroxides (LDHs), also known as hydrotalcite−like compounds, are widely used in many fields due to their unique structural advantages. Based on LDHs, a wide range of metal catalysts could be synthesized with high metal dispersion, tunable acid−base properties, facile but flexible preparation methods, strong metal−support interaction, and thermal stability. Owing to these outstanding advantages, LDH−derived materials manifest great potential as catalysts, particularly in hydrogenation and hydrogenolysis reactions. More than 200 papers published in the past five years in this field clearly indicated the rapid development of these materials. In this respect, it is imperative and essential to provide a timely review to summarize the current progress and motivate greater research effort on hydrogenation and hydrogenolysis catalysts derived from LDHs. In this review, the applications of LDH−derived materials as heterogeneous catalysts in various hydrogenation and hydrogenolysis reactions were comprehensively discussed. Hydrogenation of unsaturated chemical bonds, hydrodeoxygenation of oxygenated compounds, hydrogenolysis of carbon–carbon bonds and hydrogenation of nitrites and nitriles were described. This review demonstrates the extraordinary potentials of LDH−derived catalysts in hydrogenation and hydrogenolysis reactions, and it is undoubted that LDH−derived catalysts will play an even more significant role in the foreseeable future.
Content may be subject to copyright.
Catalysts 2022, 12, 1484. https://doi.org/10.3390/catal12111484 www.mdpi.com/journal/catalysts
Review
Recent Progress of Hydrogenation and Hydrogenolysis
Catalysts Derived from Layered Double Hydroxides
Zhihui Wang
1
, Wei Zhang
1
,
Cuiqing Li
2,
* and Chen Zhang
2,
*
1
School of New Materials and Chemical Engineering, Beijing Institute of Petrochemical Technology,
Beijing 102617, China
2
Beijing Key Laboratory of Enze Biomass Fine Chemicals, Beijing Institute of Petrochemical Technology,
Beijing 102617, China
* Correspondence: licuiqing@bipt.edu.cn (C.L.); zhangc@bipt.edu.cn (C.Z.)
Abstract: Layered double hydroxides (LDHs), also known as hydrotalcitelike compounds, are
widely used in many fields due to their unique structural advantages. Based on LDHs, a wide range
of metal catalysts could be synthesized with high metal dispersion, tunable acidbase properties,
facile but flexible preparation methods, strong metalsupport interaction, and thermal stability.
Owing to these outstanding advantages, LDHderived materials manifest great potential as cata-
lysts, particularly in hydrogenation and hydrogenolysis reactions. More than 200 papers published
in the past five years in this field clearly indicated the rapid development of these materials. In this
respect, it is imperative and essential to provide a timely review to summarize the current progress
and motivate greater research effort on hydrogenation and hydrogenolysis catalysts derived from
LDHs. In this review, the applications of LDHderived materials as heterogeneous catalysts in var-
ious hydrogenation and hydrogenolysis reactions were comprehensively discussed. Hydrogenation
of unsaturated chemical bonds, hydrodeoxygenation of oxygenated compounds, hydrogenolysis of
carbon–carbon bonds and hydrogenation of nitrites and nitriles were described. This review
demonstrates the extraordinary potentials of LDHderived catalysts in hydrogenation and hydro-
genolysis reactions, and it is undoubted that LDHderived catalysts will play an even more signif-
icant role in the foreseeable future.
Keywords: layered double hydroxides; hydrotalcite; hydrogenation; hydrogenolysis; catalysis
1. Introduction
Hydrogenation is one the most fundamental processes in the chemical industry with
diverse applications. In hydrogenation, atomic hydrogen is added to unsaturated chemi-
cal bonds in the presence of homogeneous or heterogeneous catalysts. Hydrogenolysis is
another important type of reaction where a carbon–carbon bond or carbon–heteroatom
bond is cleaved by hydrogen. In the past century, hydrogenation and hydrogenolysis are
particularly important in the petrochemical industry to saturate alkenes or aromatics, re-
move sulfur or nitrogen atoms, increase fuel stability and decrease toxicity [1]. Recently,
hydrogenation and hydrogenolysis attracted extensive attention for new objectives. Bio-
mass utilization [2,3] and CO
2
hydrogenation [4,5], which are crucial processes for a sus-
tainable human society, also heavily rely on the technological development of hydrogena-
tion and hydrogenolysis. Using renewable H
2
produced from ecofriendly processes [6,7],
hydrogenation and hydrogenolysis in biomass upgrade and CO
2
hydrogenation could
greatly reduce the dependency on fossil fuels. In most industrial cases, hydrogenation or
hydrogenolysis are catalyzed by supported metal catalysts based on transition metals
such as Ni, Cu, Pd, or Pt [2,8,9].
Layered double hydroxides (LDHs) are layered inorganic materials consisting of pos-
itively charged brucite Mg(OH)
2
like layers and interlayer chargecompensating anions,
Citation: Wang, Z.; Zhang, W.; Li,
C.; Zhang, C. Recent Progress of
Hydrogenation and Hydrogenolysis
Catalysts Derived from Layered
Double Hydroxides. Catalysts 2022,
12, 1484. https://doi.org/
10.3390/catal12111484
Academic Editor: Vladimir Sobolev
Received: 21 October 2022
Accepted: 17 November 2022
Published: 20 November 2022
Publisher’s Note: MDPI stays neu-
tral with regard to jurisdictional
claims in published maps and institu-
tional affiliations.
Copyright: © 2022 by the authors. Li-
censee MDPI, Basel, Switzerland.
This article is an open access article
distributed under the terms and con-
ditions of the Creative Commons At-
tribution (CC BY) license (https://cre-
ativecommons.org/licenses/by/4.0/).
Catalysts 2022, 12, 1484 2 of 32
as shown in Figure 1 [10]. A typical type of LDHs is hydrotalcite (HT), a natural mineral
with chemical composition of Mg
6
Al
2
(OH)
16
CO
3
·4H
2
O [11]. LDHs and natural hy-
drotalcite share similar structure, in which metal cations can be replaced by divalent and
trivalent metal ions with similar atomic radii in the same crystal, and interlayer anions
can be replaced by other intercalating anions. LDHs, also called hydrotalcitelike com-
pounds, could be expressed with the general formula [M
2+(1x)
·M
3+x
(OH)
2
]
x+
·[(A
n
)
x/n
·zH
2
O]
x
[12]. Common M
2+
ions include Mg
2+
, Cu
2+
, Zn
2+
and Ce
2+
, while common M
3+
include Al
3+
,
Fe
3+
, and Cr
3+
. These cations can be atomically and uniformly dispersed in the hydrotalcite
layer. The formed cation layers lack interlayer interaction and have a negative net charge,
which must be balanced by the inserted anions, and water molecules might also enter the
interlayer region [13]. Interlayer anions are exchangeable, such as CO
32
, NO
3
, SO
42
, Cl
,
F
, acetate and salicylate. The x value in the general formula represents the molar ratio of
M
2+
/(M
2+
+
M
3+
), and the range of its value has great influence on the composition and
structure of hydrotalcite materials [10]. Both the magnitude of the charge density on the
laminate and the extent to which the homocrystalline substitution of metal cations occurs
depend on the x value, and a genuine LDH phase is usually found only at 0.2 < x < 0.4
[14,15].
The unique structure, tunability of the layer and interlayer element composition, de-
lamination property, structural topological transformations, and the confinement effect of
LDHs provide LDHderived materials great application potentials in many fields [14,16].
In heterogeneous catalysis, LDH can be applied as catalyst itself, catalyst support or cata-
lyst precursor in many reactions such as C
C coupling, Narylation, oxidation, hydro-
genation, hydrogenolysis, etc. Although several review articles on catalytic applications
of LDHs have been published [12,17–21], reviews focusing on hydrogenation and hydro-
genolysis are still scarce to the best of the authors’ knowledge. Furthermore, in the past
few years, a large number of research articles on hydrogenation and hydrogenolysis over
LDHderived catalysts were published, demonstrating the significant development in the
field. Therefore, we believe it is timely to provide a comprehensive and indepth sum-
mary on the latest progress of LDHderived catalysts for hydrogenation and hydrogen-
olysis. This review will focus on the applications of LDHderived materials as hydrogena-
tion and hydrogenolysis catalysts, highlighting physiochemical properties, structural
properties and reactivity advantages of LDHderived catalysts. Thermocatalytic pro-
cesses catalyzed by heterogeneous catalysts will be the emphasis of this review, while
electrocatalytic or photocatalytic research will not be discussed due to their distinct nature
in mechanisms.
Figure 1. A Schematic illustration of LDH structure.
Catalysts 2022, 12, 1484 3 of 32
2. Type of LDHDerived Catalysts for Hydrogenation and Hydrogenolysis
Generally, three types of LDHderived catalysts are prepared and investigated:
LDHsupported catalysts, LDHderived mixed metal oxides (MMOs), and LDHderived
intermetallic compounds (IMCs). Table 1 shows several examples of different types of
LDHderived catalysts and their reported physiochemical properties. It could be noticed
in the table that high specific surface area and small particle size could be readily achieved
with catalysts from LDHs. This section will describe in detail the general advantages and
synthesis protocols of these three types of LDHderived materials.
Table 1. Examples of LDHderived catalysts prepared with different methods.
Catalyst Preparation Method
Surface Area
(m2/g)
Particle Size
(nm) Other properties Ref.
PdAg/ZnTiLDH Coprecipitation and photo-
chemical reduction 131 5 Pore volume 2.63 cm3·g1; mean
pore size 80.3 nm [22]
Pt/CoAlMMO Coprecipitation and reduc-
tion–deposition 87 4–5
Pore volume 0.29 cm3·g1; mean
pore diameter 3.83 nm [23]
CuMgAlMMO Urea hydrolysis 213 Pore volume 0.95 cm3·g1 [24]
CuCoAlMMO Coprecipitation 92.9 3.1 Pore volume 0.69 cm3·g1 [25]
NiMoAlMMO Coprecipitation and ion ex-
change 93 4.6
Pore volume 0.13 cm3·g1; total
acid sites 2.077 mmol·g1 [26]
NiInIMC Coprecipitation 126.4 5.8 [27]
NiMoIMC In situ coreduction 189.2 18.6 [28]
2.1. LDHSupported Catalysts
Because heterogeneous catalysts for hydrogenation and hydrogenolysis usually re-
quire the participation of metallic sites, pristine LDHs could not be used directly as hy-
drogenation and hydrogenolysis catalysts. Nevertheless, LDHs act as perfect support ma-
terials due to intrinsic structural superiorities. High specific surface area, hierarchical pore
structure, abundant surface defects, strong metalsupport interaction, surface acidbase
sites with tunable strength or concentration, or inherent confinement effects have been
reported for LDHs [14,20].
Major synthesis methods of LDHs include coprecipitation [29], urea hydrolysis, hy-
drothermal synthesis [30] and the solgel method [15]. Different synthesis methods di-
rectly affect the physical and chemical properties of the obtained LDH materials [31–33].
The most facile and common method is coprecipitation. By coprecipitation, crystallinity,
particle size distribution, and stability of LDHs can be precisely controlled by adjusting
the synthesis parameters such as pH, temperature, aging time, mixing rate, cation ratio
and solution concentration [31,32]. Coprecipitation can be combined with the anion ex-
change method to modulate the structure and function of LDHbased materials by insert-
ing ions or molecules between layers while keeping the basic structure unchanged, mod-
ifying the physiochemical properties or the surface properties of the material.
2.2. LDHDerived Mixed Metal Oxide Catalysts
Mixed metal oxides (MMOs), which could be synthesized by the heat treatment of
LDHs, showed high catalytic performance and received broad attention in hydrogenation
and hydrogenolysis. LDHderived MMOs inherit LDH characteristics such as controlled
composition, high specific surface area, and uniform morphology. In addition, by con-
verting metal hydroxides into metal oxides, highly dispersed metal oxides particles were
generated, which could act as acid or base sites. The nature, intensity and concentration
of these acid/base sites on MMOs could be tuned by controlling the type and molar ratio
of metal cations in the precursors, preparation method, calcination temperature and type
of interlayer anions [20,34,35]. The high surface area and tunable acid/base property, as
Catalysts 2022, 12, 1484 4 of 32
well as other advantages, make LDHderived MMOs an ideal catalyst support material
compared to the conventional metal oxide support. MMOs could also be used inde-
pendently as hydrogenation/hydrogenolysis catalysts. For MMOs with hydrogena-
tionactive metals such as Ni or Cu, the reduction of MMOs could generate highly dis-
persed metallic sites in situ which are catalytically active for hydrogenation/hydrogenol-
ysis. Compared with supported metal catalysts prepared by impregnation, reduced
MMOs as hydrogenation/hydrogenolysis catalysts are often more reactive, due to the high
metal dispersion, abundant acid/base sites and strong metal–support interaction, as
shown later in this review.
The calcination of LDHs to generate MMOs is a topological transformation process.
During thermal transformation three temperature domains of LDH structure changes can
be observed. The first structural change at around 190 °C corresponds to the dehydration
of loosely bonded interlayer water, accompanied by the decrease in the layer spacing,
while the layered structure is retained [36]. The second temperature domain in the range
of 200–400 °C indicated dehydroxylation and the collapse of the layered structure of
LDHs, in situ forming highly dispersed metal oxides [37]. The third domain in the range
of 500–1000 °C is ascribed to the formation of sintered metal particles and spinel phases
[36,38]. The structure and function of MMOs can be controlled by changing the topological
transformation parameters such as heating rate, calcination temperature and calcination
atmosphere of LDH precursors during the thermal treatment, leading to the modification
of the catalytic performance of MMOs [10,17,34,39,40].
2.3. Intermetallic Compound Catalysts
Intermetallic compounds (IMCs), also called ordered alloy, is a special type of alloy
composed of two (or more) metal elements with specific crystal structure and atomic com-
position. IMCs can be expressed as AxBy, where x and y are small integers, usually 1, 2 and
3 [41,42]. Compared with conventional alloys, IMCs exhibit completely or partially or-
dered surface structure. Figure 2 showed the structure of NiSb and NiBi IMCs catalysts
prepared by Wei’s group [43]. It could be noticed that high metal dispersion and uniform
alloy composition were obtained with IMCs, which were important for catalytic reactivity.
Due to the unique electronic and geometric properties, IMCs promote the activated ad-
sorption of substrate molecules with specific patterns or configurations, thus achieving
improved catalyst activity, selectivity and stability [41,44,45].
Figure 2. (B1) SEM, (B2) TEM and particle size distribution, (B3) elemental mapping, and (B4) lattice
spacing images of NiSb IMC; (C1) SEM, (C2) TEM and particle size distribution, (C3) elemental
mapping, and (C4) lattice spacing images of NiBi IMC [43]. (Copyright 2020, Elsevier).
The preparation of IMCs can be categorized into three methods, controlled colloid
synthesis, inorganic capsule synthesis, and layered double hydroxide synthesis [42]. The
Catalysts 2022, 12, 1484 5 of 32
synthetic method based on LDHs has particular advantages. The composition flexibility
of LDH precursors enable the possibility of synthesizing a variety of IMC catalysts [45,46].
Metal elements of LDH layers have a high degree of dispersion at the atomic level, which
is conducive to the formation of metal catalysts with high metal dispersion after the calci-
nationreduction process. The preparation of IMC nanocrystals using LDH precursors can
be achieved by two methods, endogenous method or exogenous method. For the endog-
enous method, two metal components of IMCs (e.g., Ni, Co, Cu, Ga) are introduced sim-
ultaneously into the host layer of LDHs, followed by the calcination and reduction of LDH
precursors, affording IMC catalysts. For noble metal elements (Au, Rh, Pd, Ir, Pt, etc.) or
heavy metal elements (Sb, Sn, Bi, Pb, etc.) which are difficult or impossible to introduce
into the LDH layer, the exogenous method can be adopted, in which noble metal or heavy
metal salts are mixed with LDH precursors by impregnation, followed by calcination and
reduction to obtain IMC catalysts.
3. Hydrogenation of CarbonOxygen Unsaturated Bonds
3.1. Hydrogenation of Ketones and Aldehydes
The hydrogenation of carbonyl groups in ketones and aldehydes to form hydroxy
groups is an important process particularly in synthetic chemistry. This section will re-
view research work on ketone or aldehyde hydrogenation catalyzed by LDHderived cat-
alysts. The selective hydrogenation of αβ unsaturated aldehydes including crotonalde-
hyde, citral and cinnamaldehyde will also be discussed. The hydrogenation of furfural, a
highly important furan aldehyde derived from hemicellulose, will be reviewed in the next
section separately.
For hydrogenation over heterogeneous catalysts, metal dispersion is crucial because
high dispersion would naturally mean more active sites for reactions to take place.
Nonprecious metal catalysts such as Ni, Cu and Co could be synthesized by the calcina-
tion of corresponding LDH precursors prepared by coprecipitation. Because M2+ and M3+
ions are uniformly distributed in the brucite layers, the calcination of LDHs could in situ
generate highly dispersed metal particles and more catalytic active sites for hydrogena-
tion. For instance, Dragoi’s group have shown that a reduced CuMgAl catalyst synthe-
sized from LDH precursors exhibited particle sizes of 2.6 to 6.5 nm and high activity in
cinnamaldehyde hydrogenation to cinnamyl alcohol [40]. Another study using CuM-
gAlMMO of 1.4 to 2.4 nm metal particle size for benzyl aldehyde hydrogenation, ob-
tained 85–93% selectivity of benzyl alcohol [47].
As shown in Table 2, precious metals are particularly effective in selective hydro-
genation of unsaturated aldehydes. For precious metal catalysts such as Pt, Pd or Ru, us-
ing LDHs or LDHderived MMOs as support could also generate catalysts with high pre-
cious metal dispersion. Many examples were found in the literature reporting metal par-
ticle sizes less than 5 nm using the conventional impregnation method on LDHs/MMOs
owing to the abundant surface basicity to anchor metal species [48], such as Ru/MgAlHT
[49], Pd/MgAlMMO [50] and Au/MgAlMMO [51] for aldehyde/ketone hydrogenation.
Notably, due to the special interlayer galleries of LDHs, an alternative way to load metals
to the support other than impregnation is by intercalating metalcontaining anions in
LDHs by the ion exchange method. With this method, researchers were able to obtain a
Pt/MgAlHT catalyst with an average diameter of Pt nanoparticle being only 1.9 nm using
PtCl62 [52]. Generally, as the catalyst particle size decreases, the number of low coordina-
tion sites such as the edges and corners of the metal increases, providing more active sites
for the reaction and thus enhancing the catalytic activity of the catalyst [23,27,53].
Catalysts 2022, 12, 1484 6 of 32
Table 2. Catalytic performances of LDHderived catalysts for hydrogenation of ketones and alde-
hydes.
Catalyst Substrate Reaction Conditions Conversion (%) Product Selectivity (%) Ref.
Pt/ZnSnAl/C 2Pentenal 80 °C; 3.0 MPa H2 28.5 2Pentenol, 92.0 [54]
Cu/MgAlHT Benzaldehyde 250 °C 68 Benzyl alcohol, 93 [47]
Ru/MgAlHT Benzaldehyde 100 °C; 3.5 MPa H2 >63 Cyclohexanemethanol, 91.9 [49]
Au/Mg2AlO Crotonaldehyde 120 °C; 0.93 MPa H2 23.6 Crotyl alcohol, 62 [51]
Pt/MgAlLDH Cinnamaldehyde 80 °C; 2 MPa H2 92.6 Cinnamyl alcohol, 75.5 [52]
Pt/CoAlMMO Cinnamaldehyde 70 °C; 2 MPa H2 99.7 Cinnamyl alcohol, 72.5 [23]
Au/ZnAl Cinnamaldehyde 130 °C; 1.5 MPa H2 100 Cinnamyl alcohol, 95.7 [53]
Pt/MgAlLDH Cinnamaldehyde 60 °C; 1 MPa H2 79.7 Cinnamyl alcohol, 85.4 [55]
PtGa/MgAlGa Cinnamaldehyde 70 °C; 3 MPa H2 52.8 Cinnamyl alcohol, 70.7 [56]
Ir/MgAlFe Cinnamaldehyde 60 °C; 3 MPa H2 94.4 Cinnamyl alcohol, 79.1 [57]
Pt/CoAlLDH Cinnamaldehyde 70 °C; 3 MPa H2 94.3 Cinnamyl alcohol, 91.9 [58]
CoGaIMC Cinnamaldehyde 100 °C; 2 MPa H2 100 Cinnamyl alcohol, 96 [59]
NiZnAl/C Citral 140 °C; 1 MPa H2 100 Citronellol, 92.3 [60]
CoSnIMC Citral 160 °C; 4.0 MPa H2 100 Citronellol, 67.6 [61]
CuZnAlMMO Citral 80 °C; 1.0 MPa H2 99.8 Allylic alcohol, 75.1 [62]
NiBiIMC Unsaturated aldehydes 100 °C; 2 MPa H2 >90 Unsaturated alcohol, >93.2 [43]
NiInIMC Unsaturated aldehydes 120–145 °C; 3 MPa H2 >56 Unsaturated alcohol, >44 [27]
Pt/MgCoAl Unsaturated aldehydes 80 °C; 2 MPa H2 >87 Unsaturated alcohol, >80 [63]
Because metal particle size plays an important role in the reactivity of the catalyst,
particularly for the selective hydrogenation of αβ unsaturated aldehydes, several meth-
ods were developed for tuning particle size in LDHrelated catalysts. Xiang et al. pre-
pared HTsupported Pt nanocrystal catalysts for the selective hydrogenation of cinnamal-
dehyde [55]. It is noticed in this research that the selectivity of cinnamyl alcohol was
closely related to Pt particle size, while the size of the Pt nanocrystals could be finely tuned
by controlling the amount of surfactant added. Another study provides an example of
using different alcohols including ethylene glycol, methanol and ethanol for the reduction
of Pt on MgAlLDHs, generating different sizes of reduced Pt particles with distinct reac-
tivity in cinnamaldehyde hydrogenation [64].
For LDHderived catalysts, the fact that active metal particles were incorporated in
the LDH or MMO matrix indicates strong metalsupport interaction (SMSI) [65]. SMSI is
a specific interaction effect between the support and the metal nanoparticles (NP). This
interaction may lead to metal–metal bonding or the formation of intermetallic compounds
in the catalyst [66], enhancing the stability of the metal particles in the catalyst. In addition,
the topological transformation from LDHs to MMOs can be exploited to create SMSI by
stabilizing the active metal particles in the oxide matrix, to prevent leaching and agglom-
eration [67–69]. Wang et al. reported a slight increase in particle size from 1.2 nm to 1.6
nm for the Pt catalyst supported on ZnSnAlMMOs after 12 h of 2pentenal hydrogena-
tion reaction [54]. Metal–support interaction could also electronically and geometrically
affect metal atoms, shifting the product distribution in selective hydrogenation [56,57,67].
For instance, the addition of ZnO into NiZnAlMMOs could deactivate C=C adsorption
on Ni sites, inducing increased citronellol selectivity in citral hydrogenation [60]. Oxygen
vacancies, which play important roles in carbonyl hydrogenation, could also be readily
generated by SMSI. Miao et al. reported the formation of oxygen vacancies at the metal–
support interface of Pd/MgAlLDHs [70] and Pt/CoAlLDHs [58]. It is found that these
oxygen vacancies are activation sites for C=O bonds, and significantly enhance the selec-
tivity toward C=O hydrogenation over C=C bond hydrogenation. Although SMSI is im-
portant for catalysts, its actual mechanism on product selectivity is still complex and
casedependent. It is proposed in a study that SMSI could enhance the C=C bond adsorp-
tion on Pd sites on MgAlLDHs, leading to an increased selectivity toward citronellal in
citral hydrogenation [71], while another research on cinnamaldehyde hydrogenation over
Catalysts 2022, 12, 1484 7 of 32
Pt/CoMgAl suggested that SMSI strengthened the carbonyl adsorption and inhibited C=C
hydrogenation [63].
Wei’s groups synthesized a series of intermetallic compound (IMC) catalysts using
LDH precursors, including CoGa, CoIn, NiBi, CoSn and NiIn, for selective hydrogenation
of unsaturated aldehydes [43,59,61]. It could be concluded from their research that by
forming IMC catalysts, the electron density of active metal (Ni or Co) was modified due
to electron transfer, leading to a change in adsorption conformation or adsorption strength
which eventually affected product selectivity. The promising results from IMCs, as well
as the convenient preparation method using LDH precursors, suggest a promising pro-
spect of IMCs in heterogeneous catalysis.
For αβ unsaturated aldehydes, total hydrogenation of both C=C bonds and C=O
bonds will also decrease yields toward desired products, and in most cases selective hy-
drogenation is pursued rather than total hydrogenation. The tunable acid–base properties
of LDHs or MMOs play an important role in product distribution of selective hydrogena-
tion of αβ unsaturated aldehydes [72,73]. It is shown that CuZnAlMMO catalysts with
the most Lewis acid sites which adsorbed citral molecules via the carbonyl group showed
the highest selectivity toward allylic alcohol products such as geraniol and nerol [62]. An-
other study on Pt/MgAlMMO also suggested that catalysts with stronger acidity were
more prone to unsaturated alcohol production [74]. It is deduced that Lewis acid sites,
such as metal cations, are adsorption and activation sites for C=O bond, and thus the se-
lectivity of C=O bond hydrogenation products is closely associated with Lewis acidity of
catalysts. LDHderived catalysts often exhibit a close vicinity to Lewis acid sites for car-
bonyl adsorption and metal sites for hydrogenation, inducing synergistic effects for effi-
cient hydrogenation of carbonyl bond in aldehyde or ketone hydrogenation [75].
3.2. Hydrogenation of Furfural
Furfural (FAL) is an important biomass platform chemical bridging biomass feed-
stock and bioderived chemicals with an annual production exceeding 652 kilotons [76].
FAL could be converted from lignocellulose and their chemical versatility implies plenti-
ful potential applications in the production of renewable C5 chemicals [77]. C=C double
bond and C=O double bond in furfural molecules require selective hydrogenation capa-
bility from catalysts to yield desired products. The most prevalent product derived from
furfural is furfuryl alcohol (FOL), as more than 65% of FAL produced worldwide was
converted into FOL [78]. The hydrogenation of FAL to yield furfuryl alcohol (FOL) could
be achieved by various metals [77]. Nonprecious metal Cu and Nibased catalysts pre-
pared from LDH precursors were studied for carbonyl group hydrogenation to produce
furfuryl alcohol, such as CuAl [79,80], CuCr [81,82], NiAl [83,84], NiSn [85], and CuNiAl
[86,87]. Advantages of preparing Cu or Ni catalysts from LDH precursors are the small
metal particle size and high stability, due to the uniform atom distribution in LDH pre-
cursors. Yang et al. reported the NiSn intermetallic compound catalyst prepared from
LDH precursors exhibited a monometallic Ni particle with a size of around 10 nm. The
high dispersion and the electronic modification effect of Sn resulted in a 99% selectivity
toward FOL from FAL at 100 °C and 2 MPa, and a high stability after more than 30 h [85].
Considering that the C=O bond from the furfural molecule could interact with the
basic site on the catalyst surface, metal catalysts with surface basicity were synthesized,
with Mg being the most commonly used metal due to the facile preparation from
MgAlhydrotalcite. NiMgAlMMOs synthesized by calcining hydrotalcite precursors
displayed high reactivity for the hydrogenation of FAL to FOL, with the synergistic effect
between metallic Ni species to dissociate H2 and surface basic sites generated by Mg in-
troduction to activate furfural via carbonyl groups [88]. Villaverde et al. prepared CuM-
gAl catalysts derived from HTlike phases possessed high copper dispersion as well as
strong interactions between metallic copper and magnesium–aluminum support, result-
ing in better activity, selectivity and stability than impregnated Cu catalysts [89,90]. Since
Catalysts 2022, 12, 1484 8 of 32
CuNiAl, NiMgAl, and CuMgAl all showed outstanding reactivity in furfural hydrogena-
tion, it is reasonable that CuNiMgAl would also display desirable catalytic properties
[91,92], and it is demonstrated that the uniform distribution of highly dispersed CuNi
particles and surface basicity were crucial for the enhanced catalytic performances [93].
3.3. Hydrogenation of Levulinic Acid
Levulinic acid (LA) is an important biomass platform chemical which can be pro-
duced by the acidcatalyzed hydrolysis of lignocellulose [94]. One of the most important
downstream chemicals from LA is gammavalerolactone (GVL), a stable and nontoxic
lactone that has broad applications and could be further converted into valuable com-
pounds [95]. For example, GVL could be transformed into 1,4pentanediol (1,4PDO) by
a hydrogenative ringopening reaction, and 1,4PDO could go through an intermolecular
dehydration to form 2methyltetrahydrofuran (2MTHF) [26,96]. Two mechanisms for LA
hydrogenation into GVL were proposed based on literature research. In the first mecha-
nism [97], LA was initially hydrogenated to form 4hydroxypentanoic acid on metallic
sites, followed by the lactonization to yield GVL at acid or basic sites. The second mecha-
nism [98] started with an isomerization of LA to the enol form, followed by the lactoniza-
tion to produce angelica lactone, and ends with hydrogenation to produce GVL. Some
studies stated that the hydrogenation was the ratedetermining step, while the lactoniza-
tion reaction proceeded rapidly [99]. High hydrogen pressure (0.5–3 MPa) is usually nec-
essary for high conversion of LA, so that the sufficient amount of dissolved hydrogen in
solvent enables LA hydrogenation [100]. Many catalysts based on LDHs or MMOs were
reported to be efficient for converting LA to GVL with high yields. Based on the published
research, it seems that MMO catalysts based on Cu, Ni or Co showed comparably high
selectivity toward GVL from LA. For example, CuAlMMO [82], NiAlMMO [101,102]
and CoAlMMO [103] catalysts were synthesized by different research groups and all
showed comparably high selectivity (>85%) toward GVL. Analogously, CoMoAl [104],
NiMgAl [101] and CuMgAl [105] catalysts synthesized by calcination of LDH precursors
also showed dominant GVL production. Noble metal supported on LDH or MMO mate-
rials were also efficient for LA hydrogenation. Notably, MMOssupported noble metal
catalysts with a much lower reaction temperature (40–80 °C) were reported recently, com-
pared to a reaction temperature of 140–260 °C used by most nonnoble metal catalysts. Pt
supported on MgAlMMO demonstrated a nearly complete LA conversion into GVL at
40 °C and 50 bar [106], while Ru supported on MgAlMMO or MgLaMMO also showed
a GVL selectivity higher than 99% at 80 °C and 5 bar [107].
Although high selectivity toward GVL from LA was achieved by many catalysts, a
major concern in LA hydrogenation catalysts that should be dealt with is catalyst deacti-
vation. Several mechanisms were proposed, including particle sintering, coke deposition,
or metal leaching [108,109]. Because LA conversion into GVL produces water in lactoni-
zation, it is probable that MgO and Al2O3 may undergo phase transition in the hydrother-
mal environment, with brucite MgO or γ−Al2O3 transformed into periclase or boehmite,
respectively, leading to undesired changes in pore structure and surface property [110–
112]. Deactivation of LDH or MMO catalysts in LA hydrogenation was also observed and
studied for its mechanisms. Hu’s research reported Ni leaching in NiMgAlMMO cata-
lysts in LA hydrogenation, mainly due to the detrimental effect of water to Mg or Al lead-
ing to structure collapse [101]. Hussain et al. observed significant decreases of LA conver-
sion in a timeonstream study of MgAlMMO. After characterization on spent catalysts,
it is confirmed that in addition to the coke deposition on catalyst surface, the MgAlMMO
catalyst adsorbed water to regain the layered structure due to the memory effect, though
this deactivation was largely reversible by calcination at 450 °C in air [113]. Considering
the structure destruction during the reaction, metal particle sintering becomes inevitable,
as observed with Cu/MgAlMMO catalysts that Cu particles sized increased from 2 to 5.3
Catalysts 2022, 12, 1484 9 of 32
nm after LA hydrogenation at 260 °C [114]. Therefore, the stability of LDH or MMO cata-
lysts for LA hydrogenation is an unavoidable issue before the commercialization of the
catalytic process.
3.4. Catalytic Hydrogenation of Monosaccharides
The conversion of biomass feedstocks to fuels or chemicals is an important approach
to generate renewable energy resources. Lignocellulose depolymerization or fermentation
could produce monosaccharides or sugar alcohols, which could be further transformed to
various types of chemicals [115]. For example, cellulose or hemicellulose could be hydro-
lyzed into C5 and C6 sugar alcohols such as glucose, xylose, and arabinose. These sugar
alcohols were listed by the US Department of Energy among the top 12 sugarderived
building block chemicals, which are stable and versatile to upgrade to commodity chem-
icals [116]. Yamaguchi et al. prepared hydrotalcite supported nickel phosphide with high
activity and stability for the hydrogenation of glucose [117], xylose [118], and maltose
[119] under mild condition. The authors indicated that the hydrotalcite support activated
the carbonyl group of the saccharide molecules and donated electrons to Ni, while molec-
ular hydrogen was activated by unsaturated Ni sites from Ni2P. Another study on sugar
hydrogenation using a physical mixture of PtSn/Al2O3 and MgAlMMO proposed that
the addition of MgAlMMO could provide an alkaline environment in the reaction system
to facilitate ring opening reactions of sugar molecules [120], which is a key step for car-
bonyl hydrogenation [121]. While supported LDH catalysts showed synergistic effects be-
tween support and active metal, LDHderived MMOs were also active in sugar hydro-
genation. Wu et al. have shown that reduced CuNiAl and CuNiAlM (M = Mg, Co, Cr and
Fe) catalysts were effective for the hydrogenation of glucose and fructose into sorbitol and
mannitol [122,123]. PérezRamírez et al. also reported HTderived CuNiAl catalysts
could catalyze transfer hydrogenation of these sugars to the corresponding polyols [124].
In both research, characterization results indicated that metallic Cu and Ni were formed
in the reduced catalysts, which could act as active sites for hydrogenation.
4. Hydrogenation of CarbonCarbon Unsaturated Bonds
4.1. Partial Hydrogenation of Alkynes
The production of ethylene in oil refinery is often accompanied with acetylene as the
byproduct. Because acetylene could poison Ziegler–Natta catalysts and significantly de-
teriorate the quality of polyethylene, a partial hydrogenation process is needed to elimi-
nate ethylene and not overhydrogenate ethylene to ethane [125]. Acetylene hydrogenation
is considered to be a structuresensitive reaction in which catalyst composition plays a
vital role in product selectivity [126]. Ma et al. obtained a Pd/MgAlLDHs/Al2O3 catalyst
by in situ synthesis of Pd/MgAlLDHs on spherical Al2O3 surface, and another
Pd/MgOAl2O3 catalyst by calcination and reduction of synthesized
Pd/MgAlLDHs/Al2O3 [127]. The results indicated that both catalysts had larger surface
area, lower surface acidity, uniform Pd particle size, and strong metalsupport interac-
tion, leading to higher catalytic activity and selectivity than impregnated Pd/Al2O3 [128].
The high dispersion of Pd on the HT surface is presumably originated from the acidic sites
on the support surface which weakened the electron density of Pd, favored the formation
of lowcoordinated Pd sites and enhanced dispersion [129,130]. Because Ag could effec-
tively inhibit the formation of coke and improve the activity and stability of the catalyst
[125], PdAg bimetallic catalysts supported on LDHderived MMOs were also studied for
acetylene hydrogenation. Research on PdAg supported on LDHderived ZnOAl2O3
found that compared with Al2O3, ZnOAl2O3MMO could significantly inhibit the over-
hydrogenation of acetylene and decrease oligomer formation [131]. Research on PdAg
supported by NiTiMMOs for acetylene hydrogenation found that a large number of Ti3+
defective sites existed on the NiTi support, acting as active centers to activate hydrogen
and increasing the electron density Pd, thus promoting the desorption of ethylene and
Catalysts 2022, 12, 1484 10 of 32
improving reaction selectivity [132]. When switching NiTi to MgTi as the PdAg support,
the electronic effect between Ti3+ and Pd was still present, combined with the acid–base
property of the MgTi support, contributing to nearly 100% conversion and 83.8% selectiv-
ity to ethylene at 70 °C [133].
In addition to being used as a catalyst support, LDH materials after calcination and
reduction can also be used directly as acetylene hydrogenation catalysts. Rives et al.
[134,135] prepared a series of NiZnAlCr hydrotalcitelike catalysts for the selective hy-
drogenation of acetylene. The introduction of Cr inhibited the formation of coke, while
the addition of Zn enhanced the metal–support interaction and improved the dispersion
of Ni. Liu et al. [136] prepared CuNiMgAl nanoalloy catalysts for acetylene partial hydro-
genation using LDHs as the precursor. High metal dispersion and structural homogeneity
of NiCu nanocrystals were observed, and this LDHderived CuNiMgAl showed better
selectivity, anticoking ability and stability than the CuNi/MgAlHT catalysts prepared
by the impregnation method. The addition of iron in Cubased MMOs could further im-
prove selectivity and reduce oligomerization, because iron served as a structural promoter
to disperse active metals [137]. Moreover, in another research investigating
Cu/FeyMgOxtype catalysts derived from LDHs, iron is proposed to facilitate the for-
mation of bifunctional interfacial active site Cuδ−Fe0.16δ+MgOx, which is effective in activat-
ing acetylene and hydrogen and promoting the desorption of ethylene [138].
Similar to acetylene partial hydrogenation, the partial hydrogenation of phenylacety-
lene to styrene for the elimination of phenylacetylene is an important pretreatment step
in styrene polymerization process. Several LDHbased catalysts have shown excellent
performance in this reaction. Pd/HT catalysts were synthesized for the phenylacetylene
hydrogenation, and it is proposed that the layered structure of HT may impose steric re-
striction on reactants, making Pd/HT more stereoselective in the product than the conven-
tional supported Pd catalysts [139]. Duan’s group presented a novel series of nickel phos-
phide catalysts from LDH precursors with high selectivity toward styrene. It is claimed
that phosphorus incorporation increased the Ni–Ni bond length and decreased Ni elec-
tron density, leading to the desorption of styrene so that further hydrogenation was
avoided [140].
4.2. Hydrogenation of Aromatic Ring
Aromatic compounds in transportation fuels are responsible for reduced cetane
number and increased particulate emission, as well as potential health hazard accompa-
nied by the emission [141–143]. Strict legislative restrictions on aromatic content all over
the world encourage research efforts on aromatic rings hydrogenation as an effective strat-
egy to eliminate aromatic compounds in fuels [144]. The urgent need for lignin upgrade
in recent years also inspires research on efficient aromatic hydrogenation [145]. Bai et al.
investigated the hydrogenation of phenol over a series of Nibased MMO catalysts and
over 90% selectivity of cyclohexanol was obtained at 110–150 °C [146–148]. It is pointed
out that the strong metal–support interaction of MMO catalysts induced high metal dis-
persion and prevented agglomeration or sintering. Strong metal–support interaction was
also important in aromatic hydrogenation over Pd supported by MMOs [149]. Highly dis-
persed Pd particles were formed on CoCeAlMMO surface with a particle size smaller
than 4 nm, providing plentiful active sites for H2 dissociative adsorption and ring hydro-
genation. The addition of Ce generated abundant oxygen vacancies, which activated phe-
nol by forming phenoxy and accelerated the hydrogenation reaction. Reaction tempera-
ture and pressure also significantly affected hydrogenation activity, as relatively high
temperature and high pressure were more preferable for benzyl ring hydrogenation prod-
ucts [150,151].
Catalysts 2022, 12, 1484 11 of 32
5. Hydrogenolysis of Oxygenated Compounds
5.1. Hydrogenolysis of Esters
Ester hydrogenation and hydrogenolysis are important reactions with particular in-
terest in biomass upgrade. Esters, including fatty acid esters, lactones, and levulinate es-
ters, could be either directly extracted or chemically converted from biomass feedstock.
Because esters themselves have limited uses in the chemical industry, hydrodeoxygena-
tion or hydrogenolysis is usually necessary to convert esters into valuable commodities
such as alcohols or alkanes. Triacylglycerols from nonedible plant oils or animal fats
could be converted into fatty acid esters by transesterification reactions. Due to their high
oxygen content and low heating value as fuels [152], researchers attempted to upgrade
fatty acid esters to dieselrange hydrocarbon (C15C18) by hydrodeoxygenation. Cao et al.
utilized LDHderived NiCuAl catalysts for the hydrogenolysis of soybean and waste
cooking oils, and more than 80% yield for dieselrange hydrocarbons was obtained at 260
°C and 3 MPa of H2 [153]. Lewis acid sites (Al3+) was thought to be active sites for decar-
bonylation by interacting with the oxygen atom and cleaving acyl CO bond. Another
study [154] using NiGaMgAl to catalyze the hydrodeoxygenation of methyl laurate
(C11H23COOCH3) also present a 99% yield toward C11+C12 hydrocarbons at 400 °C and 3
MPa. The promotional effect of Ga to Ni suppressed the CC bond hydrogenolysis, ac-
counting for the high hydrocarbon yield.
Converting esters to diols which could be used as monomers for polyester materials
is another attractive tactic with practical needs. Ethylene glycol, an important monomer
in the polymer industry, could be produced by an indirect pathway from syngas via di-
methyl oxalate. Wei’s group prepared CuZrMgAlMMO catalysts for dimethyl oxalate
hydrogenation and obtained 99.5% yield of ethylene glycol under 180 °C and 2 MPa [155].
According to in situ characterization techniques, the authors attributed the high reactivity
to the CuOZr metalsupport interfacial sites, where carbon–oxygen bonds were acti-
vated at Zrrelated oxygen vacancies and hydrogenated at adjacent Cu sites. Li’s group
demonstrated that 1,4pentanediol could be produced from bioderived ethyl levulinate
by hydrodeoxygenation over CuCoAlMMO catalysts, where Lewis acid sites such as
electrophile Cu+ or electrondeficient CoOx were considered to be crucial for carbonyl ad-
sorption and activation [25].
5.2. Hydrodeoxygenation of Lignin Derivatives
Lignin is a crosslinked threedimensional amorphous polymer composed of aro-
matic units, accounting for 10–35 wt.% in lignocellulosic material [145,156]. The rapid de-
velopment of the biomass industry cogenerated more than 100 million tons of lignin an-
nually [157]. Various aromatic chemicals could be produced by lignin valorization, yet
only a small portion of lignin was actually utilized for chemical production. Lignin depol-
ymerization is a complex and challenging task, with the pivotal issue being the cleavage
of CC bond and CO bond. Hydrogenolysis is an effective way to break CO linkage,
and various LDHderived catalysts were demonstrated to be highly reactive for lignin
hydrogenolysis due to its high metal dispersion and adjustable acid–base properties.
Flowerlike Ni2PAl2O3 catalysts were obtained by reduction of red phosphorus with
NiAlLDH precursors at 500 °C by Li’s group [158]. Ni2PAl2O3 showed high alkane se-
lectivity in hydrogenolysis of poplar lignin oil. The high reactivity was attributed to the
high exposure of cantilevered conical Ni sites with strong CO bond break capability and
plentiful acid sites to activate the substrate and H2, as shown in Figure 3. Other research
stated that acid sites, especially Lewis acid sites, facilitated deoxygenation of lignin deriv-
atives by adsorbing the electronrich phenyl ring or oxygen atoms, promoting ring hy-
drogenation or deoxygenation of methoxy or carbonyl [159,160]. Oxygen vacancies are
also important for constructing high reactivity LDHderived catalysts for lignin hydro-
genolysis. Reducible oxide such as CoOx [161] or CeOx [162] in MMO catalysts could form
Catalysts 2022, 12, 1484 12 of 32
surface oxygen vacancies to adsorb or activate oxygencontaining functional groups,
greatly benefiting lignin deoxygenation reaction.
Figure 3. Reaction mechanism of guaiacol hydrodeoxygenation over NiP catalysts [158]. (Copyright
2022, American Chemical Society).
Among the common linkages that exist in lignin, the most recurring linkage is the
βO4 linkage, constituting 50% of all linkages [145]. Therefore, model compounds with
βO4 linkages were often synthesized to investigate catalyst reactivity in lignin depoly-
merization [163–165]. Wang et al. prepared a NiMgAlC composite catalyst using ligno-
sulfonate as the carbon precursor to impose electronic modifications to Ni [164]. The syn-
thesized catalysts showed promising results in terms of conversion and selectivity in hy-
drogenolysis of model compounds and real lignin. The lamellar structure of LDHs could
also be utilized to improve catalytic performance. Beckham and colleagues presented an
inspiring study, where HTderived catalysts with intercalated nitrates produced phenol
monomers 2–3 times more than catalysts without nitrates in hydrogenolysis of model
compounds [166]. It is proposed that intercalated nitrates not only increased accessibility
to basic active sites but also directly participated in the hydrogenolysis process, as the
activity could be recovered after the nitrate reservoir was depleted and replenished.
5.3. Hydrodeoxygenation of Furfural
C5 polyols are important types of chemicals with huge demands in the polymer in-
dustry. Such polyols are traditionally produced from petroleum resources. Recently, re-
searchers have developed the catalytic transformation of furfural to various polyols via
hydrodeoxygenation. For furfural conversion to pentanediols (PDO), although early in
1930s Adkins and Connor presented CuCr catalysts to convert furfuryl alcohol to pen-
tanediols [167], MMO catalysts for catalytic conversion of furfurals to pentanediols were
extensively studied in the last decade. Initially, Pt supported on CoAlMMO [168] or
MgAlHT [169] synthesized by two separate research groups showed high reactivity in
furfural hydrogenolysis toward 1,5pentanediol and 1,2pentanediol, respectively, at
110–140 °C and 1.5–3 MPa. The role of these supports is related to the basicity which might
change adsorption preference leading to ringopening reactions. Subsequently, Huang’s
group noticed that CuMgAlMMO [170] or CuAlMMO [171] alone could also catalyze
Catalysts 2022, 12, 1484 13 of 32
the formation of pentanediols at a relatively high pressure of 6 MPa. The distinct differ-
ence in reactivity and selectivity between conventional impregnated Cu/Al2O3 catalyst
and coprecipitated CuAlMMO catalyst demonstrated the advantage of the much smaller
Cu particle size of the MMO catalyst (1.9 nm) compared to the impregnated catalyst (16.7
nm) [171]. Another advantage of MMO catalysts is the partially reduced metal species on
the surface, as several studies presented positive correlations between partially reduced
metal species and pentanediol selectivity [24,172,173]. Based on published literature, the
mechanism of furfural conversion to pentanediols could be summarized, as shown in Fig-
ure 4. Furfural is first hydrogenated to produce furfuryl alcohol, and then furfuryl alcohol
will go through CO bond scission to afford 1,2pentanediol or 1,5pentanediol depend-
ing on the catalyst property. Acid–base active sites [24,174–177] or partially reduced metal
species [172,173] are proposed to influence the adsorption configuration of furfuryl alco-
hol, resulting in the activation of different CO bond and different pentanediol products.
Figure 4. Reaction mechanism of furfural hydrogenolysis over CuMgAlMMO catalysts [170]. (Cop-
yright 2022, Royal Society of Chemistry).
5.4. 5Hydroxymethylfurfural Hydrodeoxygenation
Cellulose could be transformed into 5hydroxymethylfurfural (HMF), a platform
chemical with various downstream products [77]. Many important compounds can be
produced from HMF. For example, 2,5dimethylfuran (DMF) by HMF hydrodeoxygena-
tion could be used as fuel additive due to its highoctane number and energy density
[178,179]. HMF could also be converted to 2,5furandimethanol (FDM) [180,181] via hy-
drogenation of carbonyl group, to 2,5tetrahydrofurandimethanol (THFDM) [180,182] via
furan ring hydrogenation of FDM, or to 2,5dimethyltetrahydrofuran (DMTHF) [183–185]
via furan ring hydrogenation of DMF. FDM and THFDM could be used as building blocks
in polymer synthesis, and DMTHF is a potential fuel additive [77]. As shown in Table 3,
the selectivity toward DMF, FDM, or DMTHF could be controlled by choosing adequate
reaction conditions or metal catalysts [180,184].
Catalysts 2022, 12, 1484 14 of 32
Table 3. Catalytic performances of LDHderived catalysts for 5hydroxymethylfurfural hydrode-
oxygenation 1.
Catalyst Reaction Condition Conversion (%) Major Product Selectivity (%) Ref.
NiAlMMO 80 °C; 2 MPa H2 96.0 THFDM, 74 [180]
Ru/MgAlHT 220 °C; 1 MPa H2 100.0 DMF, 58 [181]
NiAlMMO 180 °C; 1.2 MPa H2 100.0 DMTHF, 97.4 [182]
NiCu/HT 90 °C; 1 MPa H2 99 DMF, 67 [185]
CoZnAlMMO 130 °C; 0.7 MPa H2 >99.9 DMF, 74.2 [186]
CuCoNiAlMMO 180 °C; 1 MPa H2, 99.8 DMF, 95.3 [187]
CuZnAlMMO 180 °C; 1.2 MPa H2 100 DMF, 90.1 [188]
CoNC/NiAlMMO 170 °C; 1.5 MPa H2 99.9 DMF, 100 [189]
NiZnAl 100 °C; 1.5 MPa H2 100 FDM, 98.2 [190]
Cu@Co/CoAlOx 180 °C; 1.2 MPa H2 100 DHTMF
,
83.6 [191]
CuCoCeMMO 210 °C; 1.5 MPa H2 100 DMF, 96.5 [192]
NiCoAlMMO 120 °C; 4 MPa H2 100 1,2,6HTO, 64.5 [193]
Cu1.5Mg1.5Al 150 °C, 6 MPa H2 100 1,2HDO, 40 [194]
1. THFDM, 2,5tetrahydrofurandimethanol; FDM, 2,5furandimethanol; DMF, 2,5dimethylfuran;
DMTHF, 2,5dimethyltetrahydrofuran; 1,2,6HTO, 1,2,6hexanetriol; 1,2HDO, 1,2hexanediol.
The transformation of HMF to afford DMF could be accomplished by various metal
catalysts. CuAl [80], NiAl [180,182], and CuNiAlMMO [195] catalysts with high metal
dispersion showed excellent performance for the conversion of HMF to DMF, with DMF
yields higher than 90%. In contrast, CoAl showed relatively lower DMF selectivity
[186,187], likely due to its lower hydrogenation reactivity [196]. The conversion of HMF
into DMF requires multifunctional catalysts with dual active sites catalyzing C=O hydro-
genation and CO hydrogenolysis. Wang et al. [188]. studied the HMF hydrodeoxygena-
tion over CuZnAlMMO catalysts, claiming that Cu+ species acted as CO cleavage sites
and Cu0 species acted as C=O adsorption sites. The synergy between Cu+ and Cu0 species
promoted the selective transformation of multifunctional groups of HMF molecules, thus
improving catalyst reactivity and selectivity. However, these active sites might be suscep-
tible to deactivation caused by changes in the metal chemical state [189]. This issue could
be tackled by the incorporation of Zn generating strong CuOZn interaction by charge
compensation, stabilizing active Cu+ species, influencing adsorption configuration and
preventing Cu agglomeration [186,190]. In another study on LDHderived CuCoAl cata-
lysts, Wang et al. pointed out that interfacial sites played key roles in the formation of
different reaction intermediates in the catalytic hydrodeoxygenation of HMF [191]. C=O
bond in the reactant was hydrogenated by the oxygen vacancies from CoOx site to form
hydroxy groups, while C=C bond was hydrogenated by CuCo interface positions. A sim-
ilar interfacial effect is observed in reduced CuCoCe [192] and CoMgFe [197] catalysts as
well, where oxygen vacancies formed on CuCeOx or CoFeOx interface activated the CO
bond and promoted CO cleavage.
The hydrogenolysis of HMF to hexanediols or hexanetriols over LDHderived cata-
lysts were also probed by researchers. NiCoAlMMO [193] and Pt supported on
MgAlMMO [198] were effective in converting HMF to 1,2,6hexanetriol at 120–160 °C
and 3–4 MPa. A 64.5% yield was obtained with a 0.5Ni2.5CoAl catalyst at 120 °C and 4
MPa. 1,2Hexanediol could be produced over CuMgAlMMO catalysts at 150 °C and 4–
6 MPa [194]. The structure–activity relation could be inferred by two correlations reported
in Hu’s research [194], as shown in Figure 5: a positive correlation between basic sites and
catalyst turnover frequency (TOF), and a negative correlation between Cu particle size
and catalyst TOF. These observations hinted at the crucial role of Cu particle size and
basicity, which controlled the catalysts reactivity and selectivity in HMF hydrogenolysis,
analogous to furfural hydrogenolysis. Conceivably, HMF or its hydrodeoxygenated prod-
uct 5methyl furfuryl alcohol were adsorbed on active sites in a tilted mode with CO
Catalysts 2022, 12, 1484 15 of 32
bonds interacting with active sites, leading to the CO bond cleavage to form polyols
[193,199].
Figure 5. Influence of the molar ratio of Cu/Mg on the (a) distribution of basic sites and (b) Cu
particle size. (c) Correlation between the Cu particle size and TOF value and (d) basic sites and TOF
value [194]. (Copyright 2022, American Chemical Society).
6. CO2 Hydrogenation
The chemical fixation of CO2 into valueadded products is a pivotal process in CO2
utilization to reduce CO2 emission, decrease fossil fuel usage, and alleviate the greenhouse
effect. Numerous research on catalytic conversion of CO2 were conducted in recent years,
and exciting progress was made for the hydrogenation of CO2 into various chemicals in-
cluding methane, methanol, carbonates, carboxylic acid, etc. [200,201]. Compared to reg-
ular hydrogenation catalysts, unique advantages in adsorption capacity of LDHderived
catalysts for CO2 hydrogenation are recognized to further promote catalyst performance
through adsorption enhancement [202]. Because a number of reviews were published in
the past few years regarding the broad CO2 hydrogenation field [9,201,203,204] or specific
CO2 hydrogenation reaction [205], the goal of this review is to analyze the applications of
LDH or MMO catalysts in CO2 hydrogenation, focusing on CO2 thermocatalytic conver-
sion into methanol and methane.
6.1. CO2 Conversion to Methanol
The CO2 hydrogenation to methanol is a promising strategy for CO2 utilization, as
methanol is a key feedstock that can be industrially converted to a wide range of chemicals
such as light olefins and gasoline [206,207]. For example, the methanoltoolefin (MTO)
process by Dalian Institute of Chemical Physics, China, has been commercialized with an
annual production of 7.16 Mt per year by the end of 2018 [208]. Methanol could be pro-
duced from syngas (CO2/CO/H2) using an industrial Cu/ZnO/Al2O3 catalysts synthesized
from zincian malachite, (Cu,Zn)2(OH)2CO3, by the coprecipitation of metal nitrates (Cu,
Zn and Al) [209]. It has been conclusively demonstrated that the superior reactivity of
Cu/ZnO/Al2O3 stemmed from the structure defects of the Cu surface and the promotional
effect of Zn to Cu [210]. Owing to the advantages of the MMO catalysts mentioned above,
Catalysts 2022, 12, 1484 16 of 32
research efforts were devoted to preparing CuZnAlbased catalysts from LDH precursors
aiming to obtain catalysts more reactive than the industrial Cu/ZnO/Al2O3 catalyst. Beh-
rens and coworkers presented CuZnAlMMO catalysts through the coprecipitation of ni-
trate solution at 25 °C and a constant pH value of 8 ± 0.7. The synthesized CuZnAl was
intrinsically more active in CO2 hydrogenation than the industrial CuZnAl catalyst when
normalized by Cu surface area, due to the highly dispersed Cu particles (7 nm) and strong
interfacial interaction between embedded Cu particles and ZnAl2O4 matrix [211].
To further improve the catalyst reactivity of CuZnAl catalysts, the addition of fourth
metal elements was explored in CO2 hydrogenation to methanol. A wide range of metals
have been added to CuZnAl by coprecipitation of metal precursors, with different metals
exhibiting different modification effects. The addition of zirconium to CuZnAl could in-
crease the basicity, strengthen the adsorption of CO2 and promote the formation of meth-
anol [212,213]. Similar effects were observed for Mn, La, Ce and Y, as a series of CuZnAlX
(X = Mn, La, Ce, Zr and Y) catalysts were synthesized in a study from Sun’s group, and
modified CuZnAlX catalysts showed a higher portion of strong basic sites, CO2 conver-
sion and methanol selectivity than the unmodified CuZnAl [214]. Yttrium modified CuZ-
nAl possessed a higher surface area and more dispersed Cu particles, resulting in a higher
methanol yield per gram of catalysts than the unmodified CuZnAl [215]. The addition of
indium to CuZnAl increased the methanol selectivity by inhibiting CO formation, yet de-
creased TOF of methanol formation [216].
Modifications on synthesis procedures were also attempted to induce structural
transformation of the CuZnAlbased MMO catalysts. Different alumina sources [217,218],
Zn precursors [219,220], phase of precipitated precursors [221], atomic ratios [222–224],
sequence of precursor addition [225,226], calcination temperature [227], precipitation pH
[228] and precipitation agents [229] in synthesis procedures of MMO catalysts commonly
resulted in diverse catalytic activity, selectivity or stability for CO2 hydrogenation. These
phenomena were often attributed to the difference in acid–base properties, exposed Cu
surface area, particle size, phase composition, etc. However, due to the structural com-
plexity of active sites in methanol synthesis [210,230], it is not easy to acquire a direct
correlation between catalyst physiochemical properties and catalytic performance. More
indepth research on catalyst structure, especially in situ characterizations, is essential to
unveil the structure–activity relationship for CuZnAl catalysts.
For the industrial Cu/ZnO/Al2O3 catalyst, the role of alumina is considered to be a
structural promoter that prevents Cu particles from aggregation [231]. Kühl et al. con-
ducted experiments to partially replace Al with trivalent cations Cr and Ga [232]. The
substitution of Al by Cr reduced the interaction between Cu and oxide matrix, causing Cu
particle growth during CO2 hydrogenation. In contrast, the substitution of Al by Ga turns
out to be rather effective in improving catalyst activity, and this improvement was as-
cribed to the fact that Ga addition could stabilize Cu phase and benefit the catalytic activ-
ity of Cu particles. Other research presented consistent conclusions. CuZnGaMMO cat-
alysts showed higher Cu surface area and dispersion than CuZnAl using the aqueous
miscible organic solvent treatment (AMOST) method, which involves an additional step
of treating LDHs precipitates with an organic solvent [233,234].
6.2. CO2 Conversion to Methane
The conversion of CO2 to methane, namely CO2 methanation, is also known as the
Sabatier reaction first reported in 1897. In the current age of pursuing carbon neutrality,
Sabatier reaction is experiencing its renaissance, as CO2 methanation becomes a conven-
ient strategy to convert CO2 to a widely used energy resource in a sustainable manner.
Two types of reaction mechanisms for CO2 methanation were proposed. The first mecha-
nism presumes the direct methanation of CO2 to methane via formate, carbonate or bicar-
bonate, while the second mechanism requires the conversion of CO2 to CO before CO is
converted to CH4 [235–237]. The actual mechanism may vary depending on the catalyst
Catalysts 2022, 12, 1484 17 of 32
type, reaction temperature, etc. As CuZnbased catalysts constitute the majority of cata-
lysts for methanol synthesis, Nibased catalysts represent the most important category of
CO2 methanation catalysts [236]. Another important factor for the catalyst of methane for-
mation is the amount and strength of basic sites [238,239]. In this respect, MMO catalysts
derived from MgAlLDHs were studied comprehensively due to its advantageous capa-
bility of combining tunable basic sites with dispersed metal particles.
The strong dependence of CO2 conversion to catalyst basicity, especially medium
strength basic sites, was reported by many researchers, as indicated by the positive corre-
lation between medium strength basic sites and CO2 conversion or TOF [240–242]. Basic
sites, such as OH and O2, could adsorb and activate CO2 to form carbonate or bicarbonate
species [243,244]. It is noticed that increased basicity was related to the increased amount
of bicarbonatelike species for NiMgAlMMO catalysts, while bicarbonate is the main
CO2 adsorption product and an important intermediate for methane formation at low
temperatures [245]. It is also speculated that higher basicity promotes the hydrogenation
rate of surfaceadsorbed carbonates [246]. Moreover, medium strength basic sites pro-
mote monodentate carbonates formation, as monodentate is more readily hydrogenated
to form CH4 than bidentate carbonates [238,245]. The introduction of certain metal pro-
moters is a facile way to manipulate MMO catalyst basicity. Metals including V [247], Y
[240], Fe [246,248], Mn [242], Zr [249], Ce [250], La [251–253] and Cu [241] have been
demonstrated to result in increased concentration of basic sites. Apart from changes in
basicity, metal promoters could also adjust the metal–support interaction and increase Ni
dispersion. Ho and colleagues [253] claimed that CO2 interacted more strongly with
NiLaAl than NiAl due to higher basicity and more dispersed Ni particles, leading to a
much higher lowtemperature activity in CO2 methanation of NiLaAl. In situ characteri-
zation studies suggested that both dissociative activation (forming CO) and associative
activation (forming bicarbonate and carbonate) of CO2 were simultaneously observed for
NiLaAl, while for NiAl, CO2 was successively activated by the dissociative pathway and
then associative.
Oxygen vacancies are also proposed as a crucial factor for CO2 methanation, as
shown in Figure 6 [250]. Oxygen vacancies could act as Lewis acid sites to interact with
electronrich oxygen atoms from CO2 for CO2 chemisorption and activation [254]. DFT
calculations confirmed that oxygen vacancies on simulated the NiCeAl surface contrib-
uted to the lower CO2 adsorption energy than the NiAl surface [255]. Oxygen vacancies
also facilitate the formation of active oxygen species, which could react with CO2 to pro-
duce monodentate or bidentate carbonates, accelerating the methanation reaction
[249,250]. He et al. confirmed that NiZrAl ternary MMO catalysts showed more oxygen
vacancies and basic sites than NiAl binary counterpart.
Figure 6. The role of oxygen vacancies in CO2 methanation via (a) formate pathway and (b) CO
pathway [250]. (Copyright 2022, Elsevier).
Catalysts 2022, 12, 1484 18 of 32
7. Hydrogenation in CC Coupling Processes
In recent years, the demand for bioderived fuels or chemicals stimulates extensive
interests in converting biomass feedstock into various types of renewable fuels. Regarding
the fact that most biomass platform chemicals consist of oxygencontaining functional
groups such as hydroxy or carbonyl groups, CC coupling or CC bond formation via
aldol condensation, dehydration/hydrodeoxygenation, and hydrogenation are effective
strategies to produce downstream chemicals, especially in the synthesis of long chain hy-
drocarbons such as jet fuel in the range of C8C16. Regarding the advantages of LDHde-
rived MMOs simultaneously possessing metallic, acidic and basic sites, they become ideal
candidates either as catalyst supports or as the catalyst themselves to achieve the cascade
hydroconversion of CC coupling for biomass upgrade. In some cases [256–262], CC cou-
pling to produce chemicals was accomplished by a twostage process, with the first stage
aimed at aldol condensation and dehydration using HT or MMO catalysts, and the second
stage aimed at hydrogenation using nonHT and nonMMO catalysts. These studies,
which did not use HT or MMO as hydrogenation catalysts, will not be discussed in this
review. The following section will examine literature work on HT or MMOcatalyzed
hydrogenation in the CC coupling processes.
Acetone, an inexpensive byproduct in lignocellulose fermentation, could undergo al-
dol condensation with itself or other chemicals to produce products with longer chains.
The transformation of acetone to methyl isobutyl ketone (MIBK), an important chemical
with extensive applications, is a typical example of the multifunctionality of LDHderived
materials. MIBK could be converted from acetone via three consecutive steps: aldol con-
densation catalyzed by bases, dehydration catalyzed by acids, and hydrogenation cata-
lyzed by metals [263]. Regarding the advantages of LDHderived MMOs, they are cata-
lytically active as supports (Pd on ZrCr [263], MgAl [264,265] and CoAl [266]) or catalysts
themselves (NiMgAl [267] and CuAl [268]) to achieve the onepot conversion of acetone
to MIBK. Because the sequence of the three steps are crucial for MIBK production from
acetone, the concentration and the strength of acidic, basic, and metallic sites are expected
to be balanced in order to minimize side products [264]. In order to obtain the optimal
catalysts, catalysts with different concentrations of acid or base sites were prepared and
tested. According to experimental results, it is generally believed that excessively high
acidity or basicity would lead to overcondensation products and reduce MIBK yields
[264,267].
Cyclopentanone (CPO, C5H8O), a bioderived chemical produced by aqueous phase
furfural rearrangement, could be transformed to jetfuel range cycloalkanes by CC cou-
pling reactions. Cai and coworkers synthesized a bifunctional Ni/MgAl/active carbon cat-
alyst using a MgAlHT precursor in an integrated CC coupling/hydrogenation process
for the conversion of CPO with over 80% yield toward C10 or C15 alkanes [269,270]. The
authors ascribed the remarkable performance to the enhanced strength of the basic sites
of MgAlMMO, promoting trimerization of CPO more effectively than conventional ox-
ide support such as MgO, Al2O3, NaY or TiO2, as well as the strong hydrodeoxygenation
activity of Ni than Fe, Co or Cu. Another research combining MgAlMMO with Pd in
CPO trimerization also obtained high trimer yield [271]. In ad dition t o CPO, CC coupling
of other ketones were also attempted to produce elongated alkanes using LDH or MMO
catalysts. Sheng et al. performed MIBK selfcondensation over Pdmodified MgAlHT to
produce dodecanol, with MgAlHT support acting as a base catalyst for selfaldol con-
densation [272].
Alcohols could also be used for CC coupling after dehydrogenation to form alde-
hydes or ketones, followed by aldol condensation, dehydration and hydrogenation. For
these processes, metallic sites on MMO catalysts could catalyze the dehydrogenation of
the initial alcohol and the hydrogenation of the unsaturated ketones, while acid–base sites
catalyze condensation and dehydration reaction. A few alcohols were studied for their
CC coupling applications, including 2hexanol [273], propanol [274], octanol [275] and
Catalysts 2022, 12, 1484 19 of 32
ethanol [276,277]. It is commonly believed in these studies that numbers of metal, acid,
and base sites should be delicately designed to prevent side reactions and byproducts.
8. Hydrogenation of Nitrites and Nitriles
8.1. Hydrogenation of Organic Nitrites
The selective hydrogenation of aromatic nitro compounds to produce anilines is im-
portant in synthetic chemistry. The major challenges in this process are to selectively re-
duce nitro groups and not affect other functional groups. Various types of noble or
nonnoble metal catalysts were developed to achieve this goal [278]. For catalysts with
noble metals such as Pd or Pt, key factors that determine their catalytic performance are
metal dispersion and the resistance to particle agglomeration. LDHs or MMOs, which ex-
hibit hierarchical structure able to immobilize active species, stand out as desirable cata-
lyst support. Wang et al. showed that Pd supported on MgAlLDHs demonstrated higher
catalytic activity and selectivity than Pd/SiO2 or Pd/C in substituted nitrobenzene hydro-
genation, likely due to the confining effect of metallic Pd in layers of LDHs [279]. Similar
conclusions were reached in another study studying Pd supported on MgAlMMOs,
MgO and γ−Al2O3 in nitrobenzene hydrogenation [48]. Results showed that Pd/MMOs
were superior to the latter two in terms of turnover frequency and stability, and authors
ascribed the superiority to the high dispersion of Pd particles on MMOs and the anchoring
effect of MMOs to prevent Pd aggregation. Nonnoble MMO catalysts such as CuZnAl
also showed high activity in the hydrogenation of nitro groups [280]. The activity of CuZ-
nAl was attributed to the CuZnOx active sites, similar to the one used in CO2 conversion,
as mentioned above.
A more challenging case in aromatic nitro compound hydrogenation is the selective
hydrogenation of nitro groups when other readily reducible groups are presented, as in
the case of nitrostyrene hydrogenation to yield vinylaniline. Corma and Serna have re-
ported a breakthrough by using Au/TiO2 and Au/Fe2O3 to hydrogenate 3nitrostyrene to
selectively produce 3vinylaniline [281]. Recently, a series of gold catalysts supported on
hydrotalcites also showed promising activity. Zhang’s group discovered that thiolated
Au25 nanoclusters supported on calcined ZnAlHT showed complete 3nitrostyrene con-
version in a wide temperature window from 90 to 135 °C with 3vinylaniline selectivity
higher than 98%. The high selectivity was attributed to the ZnAlHT support adsorbing
nitro groups rather than vinyl groups [282]. Followup research comparing MgAl,
ZnAl, and NiAlHT as catalyst support observed that the amount of basic sites were in
the order of MgAl > NiAl > ZnAl, while the 3vinylaniline yield followed the order of
MgAl < NiAl < ZnAl, suggesting that medium to weak basicity of the HT surface might
benefit the reactivity [283]. Because nitro groups were generally believed to readily adsorb
on basic surface, the role of calcined HT is speculated to be adsorption sites for nitro
groups, in synergy with the gold particles which activate hydrogen. Nonetheless, the
choice of HT support used for nitrostyrene hydrogenation catalysts still request special
attention, as another research indicated that the selectivity toward vinyl hydrogenation
could be enhanced by changing HT composition [284].
8.2. Hydrogenation of Nitriles
Hydrogenation of nitrile is an important industrial route for amines production. To
inhibit byproducts such as secondary or tertiary imines, bases such as alkali bases or am-
monia are commonly added along with metal catalysts [285]. The inherent basicity of
MgAlMMO materials makes them potential candidates for the hydrogenation of nitriles
to produce amines. Tichit and coworkers prepared a series of Nicontaining catalysts
based on LDH precursors for catalytic hydrogenation of nitriles. They discovered that the
introduction of Mg in LDH precursors was effective in decreasing the surface acidity of
the catalyst, decreasing the adsorption strength of primary amines, and thus preventing
further coupling reaction between amines and imines. [286,287]. Cao et al. formulated a
Catalysts 2022, 12, 1484 20 of 32
core–shell catalyst by coating nickelbased nanocomposites with LDHs or layered double
oxides (LDOs), to catalyze the selective hydrogenation of benzonitrile to Nbenzylaniline
or benzylmethylamine. It was found that the structural and acidbase properties of the
LDH/LDOcoated nanocomposites could be switched by calcination or hydration accord-
ing to the memory effect of LDHs [288]. Their following work on LDHderived NiM-
gAlMMO catalysts for the hydrogenation of benzonitrile also showed that the presence
of strong metal–support interactions in NiMgAl catalysts could effectively inhibit the
leaching or aggregation of Ni nanoparticles, accounting for the excellent stability of NiM-
gAlMMO than the impregnated Ni/Mg0.75Al0.25 catalysts [289]. The strong metal–support
interaction was also observed in another study, in which a highly stable and active Co2P
supported on HT was prepared and studied for the ammoniafree selective hydrogena-
tion of various nitriles to corresponding primary amines [290]. Compared to Al2O3, SiO2,
or carbon support, HT support stabilized Co species and prevented their oxidation, lead-
ing to enhanced reactivity and stability.
9. Future and Prospect of LDHDerived Catalysts for Hydrogenation and Hydrogen-
olysis
Hydrogenation and hydrogenolysis are important reactions particularly in the pet-
rochemical industry, biomass upgrade and CO2 conversion. In the past decade, research-
ers have carried out extensive research on the application of LDHderived catalysts in
hydrogenation and hydrogenolysis. As reported in numerous studies mentioned above,
superior reactivity of LDHs and their derived catalysts undoubtedly demonstrates out-
standing advantages over conventional catalysts in the following aspects:
(1) Various facile preparation methods, including coprecipitation, hydrothermal meth-
ods, ion exchange, urea hydrolysis, etc., could be used to prepare LDHderived cat-
alysts. These preparation methods are generally mature and wellestablished, facili-
tating the wide application and scaleup production of LDH materials.
(2) The appropriate amount and strength of acidity/basicity are crucial for the design of
multifunctional catalysts, and these properties could be achieved with LDHderived
materials. LDHs and MMOs possess high concentrations of acid/base sites, which are
key sites for adsorption and reaction. For example, as discussed in this review, acidic
sites could catalyze deoxygenation reactions, while basic sites could interact with
carbonyl groups or catalyze aldol condensation. More importantly, the acid–base
properties of LDHs and their derivative materials can be finely modulated by con-
trolling the main layer element composition, interlayer ion species of the LDH pre-
cursor, and other synthesis parameters.
(3) Metal particle size is crucial for the reactivity and selectivity of metal catalysts. In
hydrogenation or hydrogenolysis reactions, metal catalysts with high metal disper-
sion will provide more active sites for dissociative activation of hydrogen and surface
reaction to take place, resulting in increased reaction rates and decreased catalyst us-
ages. For LDHsupported metal catalysts, hydroxy groups and interlayer galleries
could promote the dispersion of metal particles. For MMO or IMC catalysts, because
metal cations are uniformly distributed in the atomic level within LDH layers, the
calcination of LDH precursors would generate MMOs with highly dispersed metal
atoms. Therefore, LDHderived materials become ideal catalysts for hydrogenation
or hydrogenolysis due to their capability of generating small metal particles.
(4) For MMO or IMC catalysts, when LDHs are calcined at high temperatures, active
metal atoms are immobilized in the metal oxide matrix. This interaction between ac-
tive metals and oxide matrix results in the formation of strong metal–support inter-
actions (SMSI). SMSI is important not only in effectively preventing particle aggre-
gation or sintering of the active metal during reaction but also modifying the elec-
tronic properties and catalytic reactivity of active metals.
Catalysts 2022, 12, 1484 21 of 32
Although much progress was made in the field of hydrogenation or hydrogenolysis
catalysis over LDH materials, several challenges remained as follows:
(1) The synthesis procedures of LDHs were studied thoroughly, but for LDHderived
catalysts more optimization and mechanism research on the preparation methods are
still necessary. For example, for MMO catalysts, the relationship between prepara-
tion parameters (e.g., metal precursors, precipitation pH, crystallization time, calci-
nation temperature, reduction temperature, etc.) and physiochemical properties is
still vague or casedependent.
(2) The structure of LDHderived catalysts is also currently unclear, which deserves
more characterization efforts or theoretical predictions. The surface composition of
LDHderived catalysts, the electronic and geometric interactions between metal na-
noparticles and neighboring components, and the origin of acidity/basicity of
LDHderived materials, are largely unknown, leading to difficulties in studying
structure–reactivity relationships and catalyst design.
(3) LDHs show a unique “memory effect” and good reversible topological conversion
properties after heat treatment over a wide range of temperatures. Accordingly, how
LDHderived catalysts go through structure transformation during the reaction pro-
cess also needs indepth research. For biomassrelated hydrogenation or hydrogen-
olysis research, this issue is important because water is often present either as reac-
tant or product, which might lead to topological transformation of LDH derivates.
More research on structure transformation of LDHderived catalysts before, during,
or after reaction will be valuable for a broader application of LDHderived catalysts.
10. Conclusions
In this review, the applications of LDHderived materials in the field of catalytic hy-
drogenation and hydrogenolysis are comprehensively summarized. LDHderived cata-
lysts in hydrogenation and hydrogenolysis could be categorized into three types:
LDHsupported catalysts, mixed metal oxides, and intermetallic compounds, while dif-
ferent types of catalysts exhibit different structure and physiochemical properties. The
tunability in composition and property of these LDHderived materials makes them ver-
satile in many applications. As demonstrated by numerous examples mentioned above,
LDHderived catalysts showed superior reactivity owing to their unique advantages.
Therefore, based on the current research progress, it could be envisaged that research on
hydrogenation or hydrogenolysis over LDHderived catalysts will keep on increasing
rapidly in the future, and industrial applications of LDHderived catalysts will also be
expanding due to their desirable characteristics.
Author Contributions: Z.W.: Investigation, Writing—Original Draft, W.Z.: Validation, Supervision.
C.L.: Writing—Review and Editing. C.Z.: Conceptualization, Writing—Review and Editing, Fund-
ing acquisition. All authors have read and agreed to the published version of the manuscript.
Funding: The authors would like to acknowledge funding support provided by National Natural
Science Foundation of China (No. 21905027) and Beijing Education Committee Science and Tech-
nology Project (No. KM202010017007).
Conflicts of Interest: The authors declare no conflict of interest.
References
1. Ramachandran, R.; Menon, R.K. An overview of industrial uses of hydrogen. Int. J. Hydrog. Energy 1998, 23, 593–598.
2. Corma, A.; Iborra, S.; Velty, A. Chemical routes for the transformation of biomass into chemicals. Chem. Rev. 2007, 107, 2411–
2502.
3. Gallezot, P. Conversion of biomass to selected chemical products. Chem. Soc. Rev. 2012, 41, 1538–1558.
4. Wang, W.-H.; Himeda, Y.; Muckerman, J.T.; Manbeck, G.F.; Fujita, E. CO2 hydrogenation to formate and methanol as an alter-
native to photo- and electrochemical CO2 reduction. Chem. Rev. 2015, 115, 12936–12973.
5. Wang, W.; Wang, S.; Ma, X.; Gong, J. Recent advances in catalytic hydrogenation of carbon dioxide. Chem. Soc. Rev. 2011, 40,
3703–3727.
Catalysts 2022, 12, 1484 22 of 32
6. Qureshi, F.; Yusuf, M.; Kamyab, H.; Vo, D.-V.N.; Chelliapan, S.; Joo, S.-W.; Vasseghian, Y. Latest eco-friendly avenues on hy-
drogen production towards a circular bioeconomy: Currents challenges, innovative insights, and future perspectives. Renew.
Sustain. Energy Rev. 2022, 168, 112916.
7. Yusuf, M.; Farooqi, A.S.; Keong, L.K.; Hellgardt, K.; Abdullah, B. Contemporary trends in composite ni-based catalysts for CO2
reforming of methane. Chem. Eng. Sci. 2021, 229, 116072.
8. Ruppert, A.M.; Weinberg, K.; Palkovits, R. Hydrogenolysis goes bio: From carbohydrates and sugar alcohols to platform chem-
icals. Angew. Chem. Int. Ed. 2012, 51, 2564–2601.
9. Yang, H.; Zhang, C.; Gao, P.; Wang, H.; Li, X.; Zhong, L.; Wei, W.; Sun, Y. A review of the catalytic hydrogenation of carbon
dioxide into value-added hydrocarbons. Catal. Sci. Technol. 2017, 7, 4580–4598.
10. Fan, G.; Li, F.; Evans, D.G.; Duan, X. Catalytic applications of layered double hydroxides: Recent advances and perspectives.
Chem. Soc. Rev. 2014, 43, 7040–7066.
11. Cavani, F.; Trifirò, F.; Vaccari, A. Hydrotalcite-type anionic clays: Preparation, properties and applications. Catal. Today 1991,
11, 173–301.
12. Feng, J.; He, Y.; Liu, Y.; Du, Y.; Li, D. Supported catalysts based on layered double hydroxides for catalytic oxidation and hy-
drogenation: General functionality and promising application prospects. Chem. Soc. Rev. 2015, 44, 5291–5319.
13. Yu, J.; Wang, Q.; O’Hare, D.; Sun, L. Preparation of two dimensional layered double hydroxide nanosheets and their applica-
tions. Chem. Soc. Rev. 2017, 46, 5950–5974.
14. Wang, Q.; O’Hare, D. Recent advances in the synthesis and application of layered double hydroxide (ldh) nanosheets. Chem.
Rev. 2012, 112, 4124–4155.
15. Yan, K.; Wu, G.; Jin, W. Recent advances in the synthesis of layered, double-hydroxide-based materials and their applications
in hydrogen and oxygen evolution. Energy Technol. 2016, 4, 354–368.
16. Zhao, M.-Q.; Zhang, Q.; Huang, J.-Q.; Wei, F. Hierarchical nanocomposites derived from nanocarbons and layered double hy-
droxides—Properties, synthesis, and applications. Adv. Funct. Mater. 2012, 22, 675–694.
17. Xu, Z.P.; Zhang, J.; Adebajo, M.O.; Zhang, H.; Zhou, C. Catalytic applications of layered double hydroxides and derivatives.
Appl. Clay Sci. 2011, 53, 139–150.
18. Takehira, K. Recent development of layered double hydroxide-derived catalysts—Rehydration, reconstitution, and supporting,
aiming at commercial application. Appl. Clay Sci. 2017, 136, 112–141.
19. Zhang, F.; Xiang, X.; Li, F.; Duan, X. Layered double hydroxides as catalytic materials: Recent development. Catal. Surv. Asia
2008, 12, 253.
20. Xu, M.; Wei, M. Layered double hydroxide-based catalysts: Recent advances in preparation, structure, and applications. Adv.
Funct. Mater. 2018, 28, 1802943.
21. Yan, K.; Liu, Y.; Lu, Y.; Chai, J.; Sun, L. Catalytic application of layered double hydroxide-derived catalysts for the conversion
of biomass-derived molecules. Catal. Sci. Technol. 2017, 7, 1622–1645.
22. Shen, Y.; Yin, K.; An, C.; Xiao, Z. Design of a difunctional Zn-Ti LDHs supported PdAu catalyst for selective hydrogenation of
phenylacetylene. Appl. Surf. Sci. 2018, 456, 1–6.
23. Tian, Z.; Li, Q.; Hou, J.; Pei, L.; Li, Y.; Ai, S. Platinum nanocrystals supported on coal mixed metal oxide nanosheets derived
from layered double hydroxides as catalysts for selective hydrogenation of cinnamaldehyde. J. Catal. 2015, 331, 193–202.
24. Fu, X.; Ren, X.; Shen, J.; Jiang, Y.; Wang, Y.; Orooji, Y.; Xu, W.; Liang, J. Synergistic catalytic hydrogenation of furfural to 1,2-
pentanediol and 1,5-pentanediol with LDO derived from CuMgAl hydrotalcite. Mol. Catal. 2021, 499, 111298.
25. Wu, J.; Gao, G.; Sun, P.; Long, X.; Li, F. Synergetic catalysis of bimetallic cuco nanocomposites for selective hydrogenation of
bioderived esters. ACS Catal. 2017, 7, 7890–7901.
26. Zhang, G.; Li, W.; Fan, G.; Yang, L.; Li, F. Controlling product selectivity by surface defects over moox-decorated ni-based
nanocatalysts for γ-valerolactone hydrogenolysis. J. Catal. 2019, 379, 100–111.
27. Li, C.; Chen, Y.; Zhang, S.; Xu, S.; Zhou, J.; Wang, F.; Wei, M.; Evans, D.G.; Duan, X. Ni–in intermetallic nanocrystals as efficient
catalysts toward unsaturated aldehydes hydrogenation. Chem. Mater. 2013, 25, 3888–3896.
28. Liu, W.; Yang, Y.; Chen, L.; Xu, E.; Xu, J.; Hong, S.; Zhang, X.; Wei, M. Atomically-ordered active sites in nimo intermetallic
compound toward low-pressure hydrodeoxygenation of furfural. Appl. Catal. B 2021, 282, 119569.
29. Wang, Y.; Chen, Z.; Zhang, M.; Liu, Y.; Luo, H.; Yan, K. Green fabrication of nickel-iron layered double hydroxides nanosheets
efficient for the enhanced capacitive performance. Green Energy Environ. 2022, 7, 1053–1061.
30. Liu, B.; Xu, S.; Zhang, M.; Li, X.; Decarolis, D.; Liu, Y.; Wang, Y.; Gibson, E.K.; Catlow, C.R.A.; Yan, K. Electrochemical upgrading
of biomass-derived 5-hydroxymethylfurfural and furfural over oxygen vacancy-rich nicomn-layered double hydroxides
nanosheets. Green Chem. 2021, 23, 4034–4043.
31. Bukhtiyarova, M.V. A review on effect of synthesis conditions on the formation of layered double hydroxides. J. Solid State
Chem. 2019, 269, 494–506.
32. Theiss, F.L.; Ayoko, G.A.; Frost, R.L. Synthesis of layered double hydroxides containing mg2+, zn2+, ca2+ and al3+ layer cations
by co-precipitation methods—A review. Appl. Surf. Sci. 2016, 383, 200–213.
33. Othman, M.R.; Helwani, Z.; Martunus; Fernando, W.J.N. Synthetic hydrotalcites from different routes and their application as
catalysts and gas adsorbents: A review. Appl. Organomet. Chem. 2009, 23, 335–346.
34. Debecker, D.P.; Gaigneaux, E.M.; Busca, G. Exploring, tuning, and exploiting the basicity of hydrotalcites for applications in
heterogeneous catalysis. Chem.—Eur. J. 2009, 15, 3920–3935.
Catalysts 2022, 12, 1484 23 of 32
35. He, S.; An, Z.; Wei, M.; Evans, D.G.; Duan, X. Layered double hydroxide-based catalysts: Nanostructure design and catalytic
performance. Chem. Commun. 2013, 49, 5912–5920.
36. Yang, W.; Kim, Y.; Liu, P.K.T.; Sahimi, M.; Tsotsis, T.T. A study by in situ techniques of the thermal evolution of the structure
of a Mg–Al–CO3 layered double hydroxide. Chem. Eng. Sci. 2002, 57, 2945–2953.
37. Zhao, X.; Zhang, F.; Xu, S.; Evans, D.G.; Duan, X. From layered double hydroxides to zno-based mixed metal oxides by thermal
decomposition: Transformation mechanism and uv-blocking properties of the product. Chem. Mater. 2010, 22, 3933–3942.
38. Kim, Y.; Yang, W.; Liu, P.K.T.; Sahimi, M.; Tsotsis, T.T. Thermal evolution of the structure of a mgalco3 layered double hy-
droxide: Sorption reversibility aspects. Ind. Eng. Chem. Res. 2004, 43, 4559–4570.
39. Takehira, K.; Shishido, T. Preparation of supported metal catalysts starting from hydrotalcites as the precursors and their im-
provements by adopting “memory effect”. Catal. Surv. Asia 2007, 11, 1–30.
40. Dragoi, B.; Ungureanu, A.; Chirieac, A.; Ciotonea, C.; Rudolf, C.; Royer, S.; Dumitriu, E. Structural and catalytic properties of
mono- and bimetallic nickel–copper nanoparticles derived from MgNi(Cu)Al-LDHs under reductive conditions. Appl. Catal. A
2015, 504, 92–102.
41. Armbrüster, M.; Schlögl, R.; Grin, Y. Intermetallic compounds in heterogeneous catalysis—A quickly developing field. Sci.
Technol. Adv. Mater. 2014, 15, 034803.
42. Yang, Y.; Wei, M. Intermetallic compound catalysts: Synthetic scheme, structure characterization and catalytic application. J.
Mater. Chem. A 2020, 8, 2207–2221.
43. Yu, J.; Yang, Y.; Chen, L.; Li, Z.; Liu, W.; Xu, E.; Zhang, Y.; Hong, S.; Zhang, X.; Wei, M. Nibi intermetallic compounds catalyst
toward selective hydrogenation of unsaturated aldehydes. Appl. Catal. B 2020, 277, 119273.
44. Furukawa, S.; Komatsu, T. Intermetallic compounds: Promising inorganic materials for well-structured and electronically mod-
ified reaction environments for efficient catalysis. ACS Catal. 2017, 7, 735–765.
45. Marakatti, V.S.; Peter, S.C. Synthetically tuned electronic and geometrical properties of intermetallic compounds as effective
heterogeneous catalysts. Prog. Solid State Chem. 2018, 52, 1–30.
46. Zang, Y.; Wang, Y.; Gao, F.; Gu, J.; Qu, J. Highly active two-dimensional niinalc intermetallic compounds derived from al-
substituted layered double hydroxides for CO2 hydrogenation reduction. Fuel 2021, 299, 120929.
47. Jayesh, T.B.; Itika, K.; Ramesh Babu, G.V.; Rama Rao, K.S.; Keri, R.S.; Jadhav, A.H.; Nagaraja, B.M. Vapour phase selective
hydrogenation of benzaldehyde to benzyl alcohol using Cu supported Mg-Al hydrotalicite catalyst. Catal. Commun. 2018, 106,
73–77.
48. Sangeetha, P.; Shanthi, K.; Rao, K.S.R.; Viswanathan, B.; Selvam, P. Hydrogenation of nitrobenzene over palladium-supported
catalysts—Effect of support. Appl. Catal. A 2009, 353, 160–165.
49. Chaudhari, C.; Sato, K.; Miyahara, S.-i.; Yamamoto, T.; Toriyama, T.; Matsumura, S.; Kusuda, K.; Kitagawa, H.; Nagaoka, K.
The effect of Ru precursor and support on the hydrogenation of aromatic aldehydes/ketones to alcohols. ChemCatChem 2022,
14, e202200241.
50. Martínez-Ortiz, M.J.; de la Rosa-Guzmán, M.A.; Vargas-García, J.R.; Flores-Moreno, J.L.; Castillo, N.; Guzmán-Vargas, A.; Mo-
randi, S.; Pérez-Gutiérrez, R.M. Selective hydrogenation of cinnamaldehyde using pd catalysts supported on Mg/Al mixed
oxides: Influence of the pd incorporation method. Can. J. Chem. Eng. 2018, 96, 297–306.
51. Chen, H.-Y.; Chang, C.-T.; Chiang, S.-J.; Liaw, B.-J.; Chen, Y.-Z. Selective hydrogenation of crotonaldehyde in liquid-phase over
Au/Mg2AlO hydrotalcite catalysts. Appl. Catal. A 2010, 381, 209–215.
52. Duan, J.; Wang, D.; Cui, R.; Zhang, H.; Zhang, B.; Guan, H.; Zhao, Y. In-situ incorporation of Pt nanoparticles on layered double
hydroxides for selective conversion of cinnamaldehyde to cinnamyl alcohol. ChemistrySelect 2021, 6, 13890–13896.
53. Tan, Y.; Liu, X.; Zhang, L.; Liu, F.; Wang, A.; Zhang, T. Producing of cinnamyl alcohol from cinnamaldehyde over supported
gold nanocatalyst. Chin. J. Catal. 2021, 42, 470–481.
54. Wang, H.; Lan, X.; Wang, S.; Ali, B.; Wang, T. Selective hydrogenation of 2-pentenal using highly dispersed pt catalysts sup-
ported on ZnSnAl mixed metal oxides derived from layered double hydroxides. Catal. Sci. Technol. 2020, 10, 1106–1116.
55. Xiang, X.; He, W.; Xie, L.; Li, F. A mild solution chemistry method to synthesize hydrotalcite-supported platinum nanocrystals
for selective hydrogenation of cinnamaldehyde in neat water. Catal. Sci. Technol. 2013, 3, 2819–2827.
56. Hui, T.; Miao, C.; Feng, J.; Liu, Y.; Wang, Q.; Wang, Y.; Li, D. Atmosphere induced amorphous and permeable carbon layer
encapsulating PtGa catalyst for selective cinnamaldehyde hydrogenation. J. Catal. 2020, 389, 229–240.
57. Lin, W.; Cheng, H.; Li, X.; Zhang, C.; Zhao, F.; Arai, M. Layered double hydroxide-like Mg3Al1xFex materials as supports for ir
catalysts: Promotional effects of Fe doping in selective hydrogenation of cinnamaldehyde. Chin. J. Catal. 2018, 39, 988–996.
58. Miao, C.; Zhang, F.; Cai, L.; Hui, T.; Feng, J.; Li, D. Identification and insight into the role of ultrathin LDH-induced dual-
interface sites for selective cinnamaldehyde hydrogenation. ChemCatChem 2021, 13, 4937–4947.
59. Yang, Y.; Rao, D.; Chen, Y.; Dong, S.; Wang, B.; Zhang, X.; Wei, M. Selective hydrogenation of cinnamaldehyde over co-based
intermetallic compounds derived from layered double hydroxides. ACS Catal. 2018, 8, 11749–11760.
60. Yang, L.; Jiang, Z.; Fan, G.; Li, F. The promotional effect of ZnO addition to supported ni nanocatalysts from layered double
hydroxide precursors on selective hydrogenation of citral. Catal. Sci. Technol. 2014, 4, 1123–1131.
61. Zhou, J.; Yang, Y.; Li, C.; Zhang, S.; Chen, Y.; Shi, S.; Wei, M. Synthesis of Co–Sn intermetallic nanocatalysts toward selective
hydrogenation of citral. J. Mater. Chem. A 2016, 4, 12825–12832.
62. Li, W.; Fan, G.; Yang, L.; Li, F. Surface lewis acid-promoted copper-based nanocatalysts for highly efficient and chemoselective
hydrogenation of citral to unsaturated allylic alcohols. Catal. Sci. Technol. 2016, 6, 2337–2348.
Catalysts 2022, 12, 1484 24 of 32
63. Li, C.; Ke, C.; Han, R.; Fan, G.; Yang, L.; Li, F. The remarkable promotion of In Situ formed Pt-cobalt oxide interfacial sites on
the carbonyl reduction to allylic alcohols. Mol. Catal. 2018, 455, 78–87.
64. Wang, Y.; He, W.; Wang, L.; Yang, J.; Xiang, X.; Zhang, B.; Li, F. Highly active supported Pt nanocatalysts synthesized by alcohol
reduction towards hydrogenation of cinnamaldehyde: Synergy of metal valence and hydroxyl groups. Chem.—Asian J. 2015, 10,
1561–1570.
65. Rudolf, C.; Dragoi, B.; Ungureanu, A.; Chirieac, A.; Royer, S.; Nastro, A.; Dumitriu, E. Nial and coal materials derived from
takovite-like LDHs and related structures as efficient chemoselective hydrogenation catalysts. Catal. Sci. Technol. 2014, 4, 179–
189.
66. Tauster, S.J.; Fung, S.C.; Garten, R.L. Strong metal-support interactions. Group 8 noble metals supported on titanium dioxide.
J. Am. Chem. Soc. 1978, 100, 170–175.
67. Liu, C.; Nan, C.; Fan, G.; Yang, L.; Li, F. Facile synthesis and synergistically acting catalytic performance of supported bimetallic
pdni nanoparticle catalysts for selective hydrogenation of citral. Mol. Catal. 2017, 436, 237–247.
68. Zhang, Y.; Wei, S.; Lin, Y.; Fan, G.; Li, F. Dispersing metallic platinum on green rust enables effective and selective hydrogena-
tion of carbonyl group in cinnamaldehyde. ACS Omega 2018, 3, 12778–12787.
69. Zhao, J.; Xu, S.; Wu, H.; You, Z.; Deng, L.; Qiu, X. Metal-support interactions on Ru/CaAlOx catalysts derived from structural
reconstruction of Ca-Al layered double hydroxides for ammonia decomposition. Chem. Commun. 2019, 55, 14410–14413.
70. Miao, C.; Hui, T.; Liu, Y.; Feng, J.; Li, D. Pd/MgAl-LDH nanocatalyst with vacancy-rich sandwich structure: Insight into inter-
facial effect for selective hydrogenation. J. Catal. 2019, 370, 107–117.
71. Han, R.; Nan, C.; Yang, L.; Fan, G.; Li, F. Direct synthesis of hybrid layered double hydroxide–carbon composites supported pd
nanocatalysts efficient in selective hydrogenation of citral. RSC Adv. 2015, 5, 33199–33207.
72. Chen, Y.; Liu, W.; Yin, P.; Ju, M.; Wang, J.; Yang, W.; Yang, Y.; Shen, C. Synergistic effect between Ni single atoms and acid–
base sites: Mechanism investigation into catalytic transfer hydrogenation reaction. J. Catal. 2021, 393, 1–10.
73. Basu, S.; Pradhan, N.C. Kinetics of acetone hydrogenation for synthesis of isopropyl alcohol over cu-al mixed oxide catalysts.
Catal. Today 2020, 348, 118–126.
74. Santiago-Pedro, S.; Tamayo-Galván, V.; Viveros-García, T. Effect of the acid–base properties of the support on the performance
of Pt catalysts in the partial hydrogenation of citral. Catal. Today 2013, 213, 101–108.
75. Shen, M.; Zhao, G.; Nie, Q.; Meng, C.; Sun, W.; Si, J.; Liu, Y.; Lu, Y. Ni-foam-structured Ni–Al2O3 ensemble as an efficient catalyst
for gas-phase acetone hydrogenation to isopropanol. ACS Appl. Mater. Interfaces 2021, 13, 28334–28347.
76. Jaswal, A.; Singh, P.P.; Mondal, T. Furfural—A versatile, biomass-derived platform chemical for the production of renewable
chemicals. Green Chem. 2022, 24, 510–551.
77. Chen, S.; Wojcieszak, R.; Dumeignil, F.; Marceau, E.; Royer, S. How catalysts and experimental conditions determine the selec-
tive hydroconversion of furfural and 5-hydroxymethylfurfural. Chem. Rev. 2018, 118, 11023–11117.
78. Mariscal, R.; Maireles-Torres, P.; Ojeda, M.; Sadaba, I.; Lopez Granados, M. Furfural: A renewable and versatile platform mol-
ecule for the synthesis of chemicals and fuels. Energy Environ. Sci. 2016, 9, 1144–1189.
79. Li, X.; Liu, T.; Shao, S.; Yan, J.; Zhang, H.; Cai, Y. Catalytic transfer hydrogenation of biomass-derived oxygenated chemicals
over hydrotalcite-like copper catalyst using methanol as hydrogen donor. Biomass Convers. Biorefin. 2022.
https://doi.org/10.1007/s13399-021-02209-y.
80. Zhang, J.; Chen, J. Selective transfer hydrogenation of biomass-based furfural and 5-hydroxymethylfurfural over hydrotalcite-
derived copper catalysts using methanol as a hydrogen donor. ACS Sustain. Chem. Eng. 2017, 5, 5982–5993.
81. Yan, K.; Chen, A. Efficient hydrogenation of biomass-derived furfural and levulinic acid on the facilely synthesized noble-
metal-free Cu–Cr catalyst. Energy 2013, 58, 357–363.
82. Yan, K.; Liao, J.; Wu, X.; Xie, X. A noble-metal free cu-catalyst derived from hydrotalcite for highly efficient hydrogenation of
biomass-derived furfural and levulinic acid. RSC Adv. 2013, 3, 3853–3856.
83. Ramos, R.; Peixoto, A.F.; Arias-Serrano, B.I.; Soares, O.S.G.P.; Pereira, M.F.R.; Kubička, D.; Freire, C. Catalytic transfer hydro-
genation of furfural over Co3O4Al2O3 hydrotalcite-derived catalyst. ChemCatChem 2020, 12, 1467–1475.
84. Pan, Z.; Jiang, H.; Gong, B.; Luo, R.; Zhang, W.; Wang, G.-H. Layered double hydroxide derived nial-oxide hollow nanospheres
for selective transfer hydrogenation with improved stability. J. Mater. Chem. A 2020, 8, 23376–23384.
85. Yang, Y.; Chen, L.; Chen, Y.; Liu, W.; Feng, H.; Wang, B.; Zhang, X.; Wei, M. The selective hydrogenation of furfural over
intermetallic compounds with outstanding catalytic performance. Green Chem. 2019, 21, 5352–5362.
86. Luo, L.; Yuan, F.; Zaera, F.; Zhu, Y. Catalytic hydrogenation of furfural to furfuryl alcohol on hydrotalcite-derived CuxNi3xaloy
mixed-metal oxides. J. Catal. 2021, 404, 420–429.
87. Aldureid, A.; Medina, F.; Patience, G.S.; Montané, D. Ni-Cu/Al2O3 from layered double hydroxides hydrogenates furfural to
alcohols. Catalysts 2022, 12, 390.
88. Manikandan, M.; Venugopal, A.K.; Prabu, K.; Jha, R.K.; Thirumalaiswamy, R. Role of surface synergistic effect on the perfor-
mance of Ni-based hydrotalcite catalyst for highly efficient hydrogenation of furfural. J. Mol. Catal. A Chem. 2016, 417, 153–162.
89. Villaverde, M.M.; Bertero, N.M.; Garetto, T.F.; Marchi, A.J. Selective liquid-phase hydrogenation of furfural to furfuryl alcohol
over Cu-based catalysts. Catal. Today 2013, 213, 87–92.
90. Villaverde, M.M.; Garetto, T.F.; Marchi, A.J. Liquid-phase transfer hydrogenation of furfural to furfuryl alcohol on Cu-Mg-Al
catalysts. Catal. Commun. 2015, 58, 6–10.
Catalysts 2022, 12, 1484 25 of 32
91. Xu, C.; Zheng, L.; Liu, J.; Huang, Z. Furfural hydrogenation on nickel-promoted cu-containing catalysts prepared from hy-
drotalcite-like precursors. Chin. J. Chem. 2011, 29, 691–697.
92. Xu, C.; Zheng, L.; Deng, D.; Liu, J.; Liu, S. Effect of activation temperature on the surface copper particles and catalytic properties
of Cu–Ni–Mg–Al oxides from hydrotalcite-like precursors. Catal. Commun. 2011, 12, 996–999.
93. Wu, J.; Gao, G.; Li, J.; Sun, P.; Long, X.; Li, F. Efficient and versatile CuNi alloy nanocatalysts for the highly selective hydrogena-
tion of furfural. Appl. Catal. B 2017, 203, 227–236.
94. Kang, S.; Fu, J.; Zhang, G. From lignocellulosic biomass to levulinic acid: A review on acid-catalyzed hydrolysis. Renew. Sustain.
Energy Rev. 2018, 94, 340–362.
95. Yan, K.; Yang, Y.; Chai, J.; Lu, Y. Catalytic reactions of gamma-valerolactone: A platform to fuels and value-added chemicals.
Appl. Catal. B 2015, 179, 292–304.
96. Shao, Y.; Sun, K.; Li, Q.; Liu, Q.; Zhang, S.; Liu, Q.; Hu, G.; Hu, X. Copper-based catalysts with tunable acidic and basic sites for
the selective conversion of levulinic acid/ester to γ-valerolactone or 1,4-pentanediol. Green Chem. 2019, 21, 4499–4511.
97. Li, W.; Fan, G.; Yang, L.; Li, F. Highly efficient vapor-phase hydrogenation of biomass-derived levulinic acid over structured
nanowall-like nickel-based catalyst. ChemCatChem 2016, 8, 2724–2733.
98. Gundeboina, R.; Gadasandula, S.; Velisoju, V.K.; Gutta, N.; Kotha, L.R.; Aytam, H.P. Ni-Al-Ti hydrotalcite based catalyst for
the selective hydrogenation of biomass-derived levulinic acid to γ-valerolactone. ChemistrySelect 2019, 4, 202–210.
99. Yan, K.; Chen, A. Selective hydrogenation of furfural and levulinic acid to biofuels on the ecofriendly Cu–Fe catalyst. Fuel 2014,
115, 101–108.
100. Zhang, J.; Chen, J.; Guo, Y.; Chen, L. Effective upgrade of levulinic acid into γ-valerolactone over an inexpensive and magnetic
catalyst derived from hydrotalcite precursor. ACS Sustain. Chem. Eng. 2015, 3, 1708–1714.
101. Shao, Y.; Sun, K.; Fan, M.; Wang, J.; Gao, G.; Zhang, L.; Zhang, S.; Hu, X. Selective conversion of levulinic acid to gamma-
valerolactone over Ni-based catalysts: Impacts of catalyst formulation on sintering of nickel. Chem. Eng. Sci. 2022, 248, 117258.
102. Gundeboina, R.; Velisoju, V.K.; Gutta, N.; Medak, S.; Aytam, H.P. Influence of surface lewis acid sites for the selective hydro-
genation of levulinic acid to γ-valerolactone over Ni–Cu–Al mixed oxide catalyst. React. Kinet. Mech. Catal. 2019, 127, 601–616.
103. Long, X.; Sun, P.; Li, Z.; Lang, R.; Xia, C.; Li, F. Magnetic co/al2o3 catalyst derived from hydrotalcite for hydrogenation of
levulinic acid to γ-valerolactone. Chin. J. Catal. 2015, 36, 1512–1518.
104. Wang, L.; Yang, Y.; Yin, P.; Ren, Z.; Liu, W.; Tian, Z.; Zhang, Y.; Xu, E.; Yin, J.; Wei, M. Moox-decorated Co-based catalysts
toward the hydrodeoxygenation reaction of biomass-derived platform molecules. ACS Appl. Mater. Interfaces 2021, 13, 31799–
31807.
105. Gupta, S.S.R.; Kantam, M.L. Selective hydrogenation of levulinic acid into γ-valerolactone over Cu/Ni hydrotalcite-derived
catalyst. Catal. Today 2018, 309, 189–194.
106. Siddiqui, N.; Pendem, C.; Goyal, R.; Khatun, R.; Khan, T.S.; Samanta, C.; Chiang, K.; Shah, K.; Ali Haider, M.; Bal, R. Study of
γ-valerolactone production from hydrogenation of levulinic acid over nanostructured pt-hydrotalcite catalysts at low temper-
ature. Fuel 2022, 323, 124272.
107. Swarna Jaya, V.; Sudhakar, M.; Naveen Kumar, S.; Venugopal, A. Selective hydrogenation of levulinic acid to γ-valerolactone
over a Ru/Mg–LaO catalyst. RSC Adv. 2015, 5, 9044–9049.
108. Abdelrahman, O.A.; Luo, H.Y.; Heyden, A.; Román-Leshkov, Y.; Bond, J.Q. Toward rational design of stable, supported metal
catalysts for aqueous-phase processing: Insights from the hydrogenation of levulinic acid. J. Catal. 2015, 329, 10–21.
109. Wright, W.R.H.; Palkovits, R. Development of heterogeneous catalysts for the conversion of levulinic acid to γ-valerolactone.
ChemSusChem 2012, 5, 1657–1667.
110. Varkolu, M.; Velpula, V.; Burri, D.R.; Kamaraju, S.R.R. Gas phase hydrogenation of levulinic acid to γ-valerolactone over sup-
ported Ni catalysts with formic acid as hydrogen source. New J. Chem. 2016, 40, 3261–3267.
111. Li, D.; Tian, Z.; Cai, X.; Li, Z.; Zhang, C.; Zhang, W.; Song, Y.; Wang, H.; Li, C. Nature of polymeric condensates during furfural
rearrangement to cyclopentanone and cyclopentanol over Cu-based catalysts. New J. Chem. 2021, 45, 22767–22777.
112. Gundekari, S.; Srinivasan, K. In Situ generated Ni(0)@boehmite from NiAl-LDH: An efficient catalyst for selective hydrogena-
tion of biomass derived levulinic acid to γ-valerolactone. Catal. Commun. 2017, 102, 40–43.
113. Hussain, S.; Velisoju, V.K.; Rajan, N.P.; Kumar, B.P.; Chary, K.V.R. Synthesis of γ-valerolactone from levulinic acid and formic
acid over Mg-Al hydrotalcite like compound. ChemistrySelect 2018, 3, 6186–6194.
114. Mitta, H.; Seelam, P.K.; Chary, K.V.R.; Mutyala, S.; Boddula, R.; Inamuddin; Asiri, A.M. Efficient vapor-phase selective hydro-
genolysis of bio-levulinic acid to γ-valerolactone using cu supported on hydrotalcite catalysts. Glob. Chall. 2018, 2, 1800028.
115. Alonso, D.M.; Wettstein, S.G.; Dumesic, J.A. Bimetallic catalysts for upgrading of biomass to fuels and chemicals. Chem. Soc.
Rev. 2012, 41, 8075–8098.
116. Werpy, T.; Petersen, G.; Aden, A.; Bozell, J.; Holladay, J.; White, J.; Manheim, A.; Eliot, D.; Lasure, L.; Jones, S. Top Value Added
Chemicals from Biomass: Volume 1—Results of Screening for Potential Candidates from Sugars and Synthesis Gas; DTIC Document;
DTIC: Fort Belvoir, VA, USA, 2004.
117. Yamaguchi, S.; Fujita, S.; Nakajima, K.; Yamazoe, S.; Yamasaki, J.; Mizugaki, T.; Mitsudome, T. Air-stable and reusable nickel
phosphide nanoparticle catalyst for the highly selective hydrogenation of d-glucose to d-sorbitol. Green Chem. 2021, 23, 2010–
2016.
118. Yamaguchi, S.; Mizugaki, T.; Mitsudome, T. Efficient d-xylose hydrogenation to d-xylitol over a hydrotalcite-supported nickel
phosphide nanoparticle catalyst. Eur. J. Inorg. Chem. 2021, 2021, 3327–3331.
Catalysts 2022, 12, 1484 26 of 32
119. Yamaguchi, S.; Fujita, S.; Nakajima, K.; Yamazoe, S.; Yamasaki, J.; Mizugaki, T.; Mitsudome, T. Support-boosted nickel phos-
phide nanoalloy catalysis in the selective hydrogenation of maltose to maltitol. ACS Sustain. Chem. Eng. 2021, 9, 6347–6354.
120. Tathod, A.; Kane, T.; Sanil, E.S.; Dhepe, P.L. Solid base supported metal catalysts for the oxidation and hydrogenation of sugars.
J. Mol. Catal. A Chem. 2014, 388–389, 90–99.
121. Attia, S.; Schmidt, M.C.; Schröder, C.; Weber, J.; Baumann, A.-K.; Schauermann, S. Keto–enol tautomerization as a first step in
hydrogenation of carbonyl compounds. J. Phys. Chem. C 2019, 123, 29271–29277.
122. Zhang, J.; Wu, S.; Liu, Y.; Li, B. Hydrogenation of glucose over reduced Ni/Cu/Al hydrotalcite precursors. Catal. Commun. 2013,
35, 23–26.
123. Zhang, J.; Xu, S.; Wu, S.; Liu, Y. Hydrogenation of fructose over magnetic catalyst derived from hydrotalcite precursor. Chem.
Eng. Sci. 2013, 99, 171–176.
124. Scholz, D.; Aellig, C.; Mondelli, C.; Pérez-Ramírez, J. Continuous transfer hydrogenation of sugars to alditols with bioderived
donors over Cu–Ni–Al catalysts. ChemCatChem 2015, 7, 1551–1558.
125. McCue, A.J.; Anderson, J.A. Recent advances in selective acetylene hydrogenation using palladium containing catalysts. Front.
Chem. Sci. Eng. 2015, 9, 142–153.
126. Kim, W.-J.; Moon, S.H. Modified pd catalysts for the selective hydrogenation of acetylene. Catal. Today 2012, 185, 2–16.
127. Feng, J.-T.; Ma, X.-Y.; Evans, D.G.; Li, D.-Q. Enhancement of metal dispersion and selective acetylene hydrogenation catalytic
properties of a supported pd catalyst. Ind. Eng. Chem. Res. 2011, 50, 1947–1954.
128. Ma, X.-Y.; Chai, Y.-Y.; Evans, D.G.; Li, D.-Q.; Feng, J.-T. Preparation and selective acetylene hydrogenation catalytic properties
of supported Pd catalyst by the In Situ precipitationreduction method. J. Phys. Chem. C 2011, 115, 8693–8701.
129. He, Y.; Fan, J.; Feng, J.; Luo, C.; Yang, P.; Li, D. Pd nanoparticles on hydrotalcite as an efficient catalyst for partial hydrogenation
of acetylene: Effect of support acidic and basic properties. J. Catal. 2015, 331, 118–127.
130. Liu, Y.; He, Y.; Zhou, D.; Feng, J.; Li, D. Catalytic performance of Pd-promoted Cu hydrotalcite-derived catalysts in partial
hydrogenation of acetylene: Effect of Pd–Cu alloy formation. Catal. Sci. Technol. 2016, 6, 3027–3037.
131. Gao, X.; Zhou, Y.; Jing, F.; Luo, J.; Huang, Q.; Chu, W. Layered double hydroxides derived ZnO-Al2O3 supported Pd-Ag cata-
lysts for selective hydrogenation of acetylene. Chin. J. Chem. 2017, 35, 1009–1015.
132. Liu, Y.N.; Feng, J.T.; He, Y.F.; Sun, J.H.; Li, D.Q. Partial hydrogenation of acetylene over a NiTi-layered double hydroxide sup-
ported PdAg catalyst. Catal. Sci. Technol. 2015, 5, 1231–1240.
133. Liu, Y.; Zhao, J.; He, Y.; Feng, J.; Wu, T.; Li, D. Highly efficient pdag catalyst using a reducible Mg-Ti mixed oxide for selective
hydrogenation of acetylene: Role of acidic and basic sites. J. Catal. 2017, 348, 135–145.
134. Rives, V.; Labajos, F.M.; Trujillano, R.; Romeo, E.; Royo, C.; Monzón, A. Acetylene hydrogenation on Ni–Al–Cr oxide catalysts:
The role of added Zn. Appl. Clay Sci. 1998, 13, 363–379.
135. Monzón, A.; Romeo, E.; Royo, C.; Trujillano, R.; Labajos, F.M.; Rives, V. Use of hydrotalcites as catalytic precursors of multime-
tallic mixed oxides. Application in the hydrogenation of acetylene. Appl. Catal. A 1999, 185, 53–63.
136. Liu, Y.; Zhao, J.; Feng, J.; He, Y.; Du, Y.; Li, D. Layered double hydroxide-derived Ni-Cu nanoalloy catalysts for semi-hydro-
genation of alkynes: Improvement of selectivity and anti-coking ability via alloying of Ni and Cu. J. Catal. 2018, 359, 251–260.
137. Bridier, B.; Pérez-Ramírez, J. Cooperative eects in ternary CuNiFe catalysts lead to enhanced alkene selectivity in alkyne
hydrogenation. J. Am. Chem. Soc. 2010, 132, 4321–4327.
138. Fu, F.; Liu, Y.; Li, Y.; Fu, B.; Zheng, L.; Feng, J.; Li, D. Interfacial bifunctional effect promoted non-noble Cu/FeYMgOx catalysts
for selective hydrogenation of acetylene. ACS Catal. 2021, 11, 11117–11128.
139. Mastalir, Á.; Király, Z. Pd nanoparticles in hydrotalcite: Mild and highly selective catalysts for alkyne semihydrogenation. J.
Catal. 2003, 220, 372–381.
140. Chen, Y.; Li, C.; Zhou, J.; Zhang, S.; Rao, D.; He, S.; Wei, M.; Evans, D.G.; Duan, X. Metal phosphides derived from hydrotalcite
precursors toward the selective hydrogenation of phenylacetylene. ACS Catal. 2015, 5, 5756–5765.
141. Cooper, B.H.; Donnis, B.B.L. Aromatic saturation of distillates: An overview. Appl. Catal. A 1996, 137, 203–223.
142. Stanislaus, A.; Cooper, B.H. Aromatic hydrogenation catalysis: A review. Catal. Rev. 1994, 36, 75–123.
143. Kim, K.-H.; Jahan, S.A.; Kabir, E.; Brown, R.J.C. A review of airborne polycyclic aromatic hydrocarbons (pahs) and their human
health effects. Environ. Int. 2013, 60, 71–80.
144. Qi, S.-C.; Wei, X.-Y.; Zong, Z.-M.; Wang, Y.-K. Application of supported metallic catalysts in catalytic hydrogenation of arenes.
RSC Adv. 2013, 3, 14219–14232.
145. Sun, Z.H.; Fridrich, B.; de Santi, A.; Elangovan, S.; Barta, K. Bright side of lignin depolymerization: Toward new platform chem-
icals. Chem. Rev. 2018, 118, 614–678.
146. Zhu, T.; Dong, J.; Niu, L.; Chen, G.; Ricardez-Sandoval, L.; Wen, X.; Bai, G. Highly dispersed ni/nicaalox nanocatalyst derived
from ternary layered double hydroxides for phenol hydrogenation: Spatial confinement effects and basicity of the support.
Appl. Clay Sci. 2021, 203, 106003.
147. Dong, J.; Wen, X.; Zhu, T.; Qin, J.; Wu, Z.; Chen, L.; Bai, G. Hierarchically nanostructured bimetallic NiCo/MgxNiYO catalyst
with enhanced activity for phenol hydrogenation. Mol. Catal. 2020, 485, 110846.
148. Dong, J.; Zhu, T.; Li, H.; Sun, H.; Wang, Y.; Niu, L.; Wen, X.; Bai, G. Biotemplate-assisted synthesis of layered double oxides
confining ultrafine Ni nanoparticles as a stable catalyst for phenol hydrogenation. Ind. Eng. Chem. Res. 2019, 58, 14688–14694.
149. Li, H.; Wang, X.; Liu, Y.; He, Y.; Feng, J.; Li, D. Pd nanoparticles loaded on coalce layered double oxide nanosheets for phenol
hydrogenation. ACS Appl. Nano Mater. 2021, 4, 11820–11829.
Catalysts 2022, 12, 1484 27 of 32
150. Sreenavya, A.; Sahu, A.; Sakthivel, A. Hydrogenation of lignin-derived phenolic compound eugenol over ruthenium-containing
nickel hydrotalcite-type materials. Ind. Eng. Chem. Res. 2020, 59, 11979–11990.
151. Liu, M.; Zhang, J.; Zheng, L.; Fan, G.; Yang, L.; Li, F. Significant promotion of surface oxygen vacancies on bimetallic coni
nanocatalysts for hydrodeoxygenation of biomass-derived vanillin to produce methylcyclohexanol. ACS Sustain. Chem. Eng.
2020, 8, 6075–6089.
152. Ramírez-Verduzco, L.F.; Rodríguez-Rodríguez, J.E.; Jaramillo-Jacob, A.d.R. Predicting cetane number, kinematic viscosity, den-
sity and higher heating value of biodiesel from its fatty acid methyl ester composition. Fuel 2012, 91, 102–111.
153. Cao, X.; Long, F.; Wang, F.; Zhao, J.; Xu, J.; Jiang, J. Chemoselective decarboxylation of higher aliphatic esters to diesel-range
alkanes over the NiCu/Al2O3 bifunctional catalyst under mild reaction conditions. Renew. Energy 2021, 180, 1–13.
154. Zhao, N.; Zheng, Y.; Chen, J. Remarkably reducing carbon loss and H2 consumption on Ni–Ga intermetallic compounds in
deoxygenation of methyl esters to hydrocarbons. J. Energy Chem. 2020, 41, 194–208.
155. Cui, G.; Zhang, X.; Wang, H.; Li, Z.; Wang, W.; Yu, Q.; Zheng, L.; Wang, Y.; Zhu, J.; Wei, M. ZrO2-x modified Cu nanocatalysts
with synergistic catalysis towards carbon-oxygen bond hydrogenation. Appl. Catal. B 2021, 280, 119406.
156. Li, C.Z.; Zhao, X.C.; Wang, A.Q.; Huber, G.W.; Zhang, T. Catalytic transformation of lignin for the production of chemicals and
fuels. Chem. Rev. 2015, 115, 11559–11624.
157. Bajwa, D.S.; Pourhashem, G.; Ullah, A.H.; Bajwa, S.G. A concise review of current lignin production, applications, products and
their environmental impact. Ind. Crops Prod. 2019, 139, 111526.
158. Jia, Z.; Ji, N.; Diao, X.; Li, X.; Zhao, Y.; Lu, X.; Liu, Q.; Liu, C.; Chen, G.; Ma, L.; et al. Highly selective hydrodeoxygenation of
lignin to naphthenes over three-dimensional flower-like Ni2P derived from hydrotalcite. ACS Catal. 2022, 12, 1338–1356.
159. Yue, X.; Zhang, L.; Sun, L.; Gao, S.; Gao, W.; Cheng, X.; Shang, N.; Gao, Y.; Wang, C. Highly efficient hydrodeoxygenation of
lignin-derivatives over ni-based catalyst. Appl. Catal. B 2021, 293, 120243.
160. Wang, Z.; Wang, A.; Yang, L.; Fan, G.; Li, F. Supported Ru nanocatalyst over phosphotungstate intercalated Zn-Al layered
double hydroxide derived mixed metal oxides for efficient hydrodeoxygenation of guaiacol. Mol. Catal. 2022, 528, 112503.
161. Xu, Q.; Shi, Y.; Yang, L.; Fan, G.; Li, F. The promotional effect of surface Ru decoration on the catalytic performance of Co-based
nanocatalysts for guaiacol hydrodeoxygenation. Mol. Catal. 2020, 497, 111224.
162. De Saegher, T.; Lauwaert, J.; Hanssen, J.; Bruneel, E.; Van Zele, M.; Van Geem, K.; De Buysser, K.; Verberckmoes, A. Monome-
tallic cerium layered double hydroxide supported Pd-Ni nanoparticles as high performance catalysts for lignin hydrogenolysis.
Materials 2020, 13, 691.
163. Du, B.; Chen, C.; Sun, Y.; Liu, B.; Yang, Y.; Gao, S.; Zhang, Z.; Wang, X.; Zhou, J. Ni–Mg–Al catalysts effectively promote depol-
ymerization of rice husk lignin to bio-oil. Catal. Lett. 2020, 150, 1591–1604.
164. Wang, M.; Zhang, X.; Li, H.; Lu, J.; Liu, M.; Wang, F. Carbon modification of nickel catalyst for depolymerization of oxidized
lignin to aromatics. ACS Catal. 2018, 8, 1614–1620.
165. Wang, H.-T.; Li, Z.-K.; Yan, H.-L.; Lei, Z.-P.; Yan, J.-C.; Ren, S.-B.; Wang, Z.-C.; Kang, S.-G.; Shui, H.-F. Catalytic hydrogenolysis
of lignin and model compounds over highly dispersed Ni-Ru/Al2O3 without additional H2. Fuel 2022, 326, 125027.
166. Kruger, J.S.; Cleveland, N.S.; Zhang, S.; Katahira, R.; Black, B.A.; Chupka, G.M.; Lammens, T.; Hamilton, P.G.; Biddy, M.J.;
Beckham, G.T. Lignin depolymerization with nitrate-intercalated hydrotalcite catalysts. ACS Catal. 2016, 6, 1316–1328.
167. Adkins, H.; Connor, R. The catalytic hydrogenation of organic compounds over copper chromite. J. Am. Chem. Soc. 1931, 53,
1091–1095.
168. Xu, W.; Wang, H.; Liu, X.; Ren, J.; Wang, Y.; Lu, G. Direct catalytic conversion of furfural to 1,5-pentanediol by hydrogenolysis
of the furan ring under mild conditions over Pt/CO2AlO4 catalyst. Chem. Commun. 2011, 47, 3924–3926.
169. Mizugaki, T.; Yamakawa, T.; Nagatsu, Y.; Maeno, Z.; Mitsudome, T.; Jitsukawa, K.; Kaneda, K. Direct transformation of furfural
to 1,2-pentanediol using a hydrotalcite-supported platinum nanoparticle catalyst. ACS Sustain. Chem. Eng. 2014, 2, 2243–2247.
170. Liu, H.; Huang, Z.; Zhao, F.; Cui, F.; Li, X.; Xia, C.; Chen, J. Efficient hydrogenolysis of biomass-derived furfuryl alcohol to 1,2-
and 1,5-pentanediols over a non-precious Cu–Mg3AlO4.5 bifunctional catalyst. Catal. Sci. Technol. 2016, 6, 668–671.
171. Liu, H.; Huang, Z.; Kang, H.; Xia, C.; Chen, J. Selective hydrogenolysis of biomass-derived furfuryl alcohol into 1,2- and 1,5-
pentanediol over highly dispersed cu-Al2O3 catalysts. Chin. J. Catal. 2016, 37, 700–710.
172. Sulmonetti, T.P.; Hu, B.; Lee, S.; Agrawal, P.K.; Jones, C.W. Reduced Cu–Co–Al mixed metal oxides for the ring-opening of
furfuryl alcohol to produce renewable diols. ACS Sustain. Chem. Eng. 2017, 5, 8959–8969.
173. Tan, J.; Su, Y.; Hai, X.; Huang, L.; Cui, J.; Zhu, Y.; Wang, Y.; Zhao, Y. Conversion of furfuryl alcohol to 1,5-pentanediol over
cucoal nanocatalyst: The synergetic catalysis between Cu, CoOx and the basicity of metal oxides. Mol. Catal. 2022, 526, 112391.
174. Shao, Y.; Wang, J.; Du, H.; Sun, K.; Zhang, Z.; Zhang, L.; Li, Q.; Zhang, S.; Liu, Q.; Hu, X. Importance of magnesium in Cu-based
catalysts for selective conversion of biomass-derived furan compounds to diols. ACS Sustain. Chem. Eng. 2020, 8, 5217–5228.
175. Zhu, Y.; Zhao, W.; Zhang, J.; An, Z.; Ma, X.; Zhang, Z.; Jiang, Y.; Zheng, L.; Shu, X.; Song, H.; et al. Selective activation of C–OH,
C–O–C, or CC in furfuryl alcohol by engineered pt sites supported on layered double oxides. ACS Catal. 2020, 10, 8032–8041.
176. Shao, Y.; Wang, J.; Sun, K.; Gao, G.; Li, C.; Zhang, L.; Zhang, S.; Xu, L.; Hu, G.; Hu, X. Selective hydrogenation of furfural and
its derivative over bimetallic Nife-based catalysts: Understanding the synergy between Ni sites and Ni–Fe alloy. Renew. Energy
2021, 170, 1114–1128.
177. Shao, Y.; Guo, M.; Wang, J.; Sun, K.; Zhang, L.; Zhang, S.; Hu, G.; Xu, L.; Yuan, X.; Hu, X. Selective conversion of furfural into
diols over Co-based catalysts: Importance of the coordination of hydrogenation sites and basic sites. Ind. Eng. Chem. Res. 2021,
60, 10393–10406.
Catalysts 2022, 12, 1484 28 of 32
178. Roman-Leshkov, Y.; Barrett, C.J.; Liu, Z.Y.; Dumesic, J.A. Production of dimethylfuran for liquid fuels from biomass-derived
carbohydrates. Nature 2007, 447, 982–985.
179. Hoang, A.T.; Ölçer, A.I.; Nižetić, S. Prospective review on the application of biofuel 2,5-dimethylfuran to diesel engine. J. Energy
Inst. 2021, 94, 360–386.
180. Perret, N.; Grigoropoulos, A.; Zanella, M.; Manning, T.D.; Claridge, J.B.; Rosseinsky, M.J. Catalytic response and stability of
nickel/alumina for the hydrogenation of 5-hydroxymethylfurfural in water. ChemSusChem 2016, 9, 521–531.
181. Nagpure, A.S.; Venugopal, A.K.; Lucas, N.; Manikandan, M.; Thirumalaiswamy, R.; Chilukuri, S. Renewable fuels from bio-
mass-derived compounds: Ru-containing hydrotalcites as catalysts for conversion of hmf to 2,5-dimethylfuran. Catal. Sci. Tech-
nol. 2015, 5, 1463–1472.
182. Kong, X.; Zheng, R.; Zhu, Y.; Ding, G.; Zhu, Y.; Li, Y.-W. Rational design of Ni-based catalysts derived from hydrotalcite for
selective hydrogenation of 5-hydroxymethylfurfural. Green Chem. 2015, 17, 2504–2514.
183. Hansen, T.S.; Barta, K.; Anastas, P.T.; Ford, P.C.; Riisager, A. One-pot reduction of 5-hydroxymethylfurfural via hydrogen trans-
fer from supercritical methanol. Green Chem. 2012, 14, 2457–2461.
184. Kumalaputri, A.J.; Bottari, G.; Erne, P.M.; Heeres, H.J.; Barta, K. Tunable and selective conversion of 5-hmf to 2,5-furandimetha-
nol and 2,5-dimethylfuran over copper-doped porous metal oxides. ChemSusChem 2014, 7, 2266–2275.
185. Gupta, D.; Kumar, R.; Pant, K.K. Hydrotalcite supported bimetallic (Ni-Cu) catalyst: A smart choice for one-pot conversion of
biomass-derived platform chemicals to hydrogenated biofuels. Fuel 2020, 277, 118111.
186. An, Z.; Wang, W.; Dong, S.; He, J. Well-distributed cobalt-based catalysts derived from layered double hydroxides for efficient
selective hydrogenation of 5-hydroxymethyfurfural to 2,5-methylfuran. Catal. Today 2019, 319, 128–138.
187. Xia, J.; Gao, D.; Han, F.; Lv, R.; Waterhouse, G.I.N.; Li, Y. Hydrogenolysis of 5-hydroxymethylfurfural to 2,5-dimethylfuran over
a modified coal-hydrotalcite catalyst. Front. Chem. 2022, 10, 907649.
188. Wang, Q.; Yu, Z.; Feng, J.; Fornasiero, P.; He, Y.; Li, D. Insight into the effect of dual active Cu0/Cu+ sites in a Cu/ZnO-Al2O3
catalyst on 5-hydroxylmethylfurfural hydrodeoxygenation. ACS Sustain. Chem. Eng. 2020, 8, 15288–15298.
189. Ma, N.; Song, Y.; Han, F.; Waterhouse, G.I.N.; Li, Y.; Ai, S. Highly selective hydrogenation of 5-hydroxymethylfurfural to 2,5-
dimethylfuran at low temperature over a Co–N–C/NiAl-MMO catalyst. Catal. Sci. Technol. 2020, 10, 4010–4018.
190. Kong, X.; Zhu, Y.; Zheng, H.; Zhu, Y.; Fang, Z. Inclusion of Zn into metallic Ni enables selective and effective synthesis of 2,5-
dimethylfuran from bioderived 5-hydroxymethylfurfural. ACS Sustain. Chem. Eng. 2017, 5, 11280–11289.
191. Wang, Q.; Feng, J.; Zheng, L.; Wang, B.; Bi, R.; He, Y.; Liu, H.; Li, D. Interfacial structure-determined reaction pathway and
selectivity for 5-(hydroxymethyl)furfural hydrogenation over Cu-based catalysts. ACS Catal. 2020, 10, 1353–1365.
192. Wang, X.; Zhang, C.; Zhang, Z.; Gai, Y.; Li, Q. Insights into the interfacial effects in cu-co/ceox catalysts on hydrogenolysis of 5-
hydroxymethylfurfural to biofuel 2,5-dimethylfuran. J. Colloid Interface Sci. 2022, 615, 19–29.
193. Yao, S.; Wang, X.; Jiang, Y.; Wu, F.; Chen, X.; Mu, X. One-step conversion of biomass-derived 5-hydroxymethylfurfural to 1,2,6-
hexanetriol over Ni–Co–Al mixed oxide catalysts under mild conditions. ACS Sustain. Chem. Eng. 2014, 2, 173–180.
194. Shao, Y.; Wang, J.; Sun, K.; Gao, G.; Fan, M.; Li, C.; Ming, C.; Zhang, L.; Zhang, S.; Hu, X. Cu-based nanoparticles as catalysts
for selective hydrogenation of biomass-derived 5-hydroxymethylfurfural to 1,2-hexanediol. ACS Appl. Nano Mater. 2022, 5,
5882–5894.
195. Li, W.; Fan, G.; Yang, L.; Li, F. Highly efficient synchronized production of phenol and 2,5-dimethylfuran through a bimetallic
ni–cu catalyzed dehydrogenation–hydrogenation coupling process without any external hydrogen and oxygen supply. Green
Chem. 2017, 19, 4353–4363.
196. Xia, J.; Gao, D.; Han, F.; Li, Y.; Waterhouse, G.I.N. Efficient and selective hydrogenation of 5-hydroxymethylfurfural to 2,5-
dimethylfuran over a non-noble CoNCx/NiFeO catalyst. Catal. Lett. 2022, 152, 3400–3413.
197. Zhao, J.; Liu, M.; Fan, G.; Yang, L.; Li, F. Efficient transfer hydrogenolysis of 5-hydromethylfurfural to 2,5-dimethylfuran over
cofe bimetallic catalysts using formic acid as a sustainable hydrogen donor. Ind. Eng. Chem. Res. 2021, 60, 5826–5837.
198. Kataoka, H.; Kosuge, D.; Ogura, K.; Ohyama, J.; Satsuma, A. Reductive conversion of 5-hydroxymethylfurfural to 1,2,6-hex-
anetriol in water solvent using supported pt catalysts. Catal. Today 2020, 352, 60–65.
199. Fulignati, S.; Antonetti, C.; Licursi, D.; Pieraccioni, M.; Wilbers, E.; Heeres, H.J.; Raspolli Galletti, A.M. Insight into the hydro-
genation of pure and crude HMF to furan diols using Ru/C as catalyst. Appl. Catal. A 2019, 578, 122–133.
200. Farooqi, A.S.; Yusuf, M.; Zabidi, N.A.M.; Sanaullah, K.; Abdullah, B. Chapter 3—CO2 conversion technologies for clean fuels
production. In Carbon Dioxide Capture and Conversion; Nanda, S., Vo, D.-V.N., Nguyen, V.-H., Eds.; Elsevier: Amsterdam, The
Netherlands, 2022; pp 37–63.
201. Ra, E.C.; Kim, K.Y.; Kim, E.H.; Lee, H.; An, K.; Lee, J.S. Recycling carbon dioxide through catalytic hydrogenation: Recent key
developments and perspectives. ACS Catal. 2020, 10, 11318–11345.
202. Fang, X.; Chen, C.; Jia, H.; Li, Y.; Liu, J.; Wang, Y.; Song, Y.; Du, T.; Liu, L. Progress in adsorption-enhanced hydrogenation of
CO2 on layered double hydroxide (LDH) derived catalysts. J. Ind. Eng. Chem. 2021, 95, 16–27.
203. Álvarez, A.; Bansode, A.; Urakawa, A.; Bavykina, A.V.; Wezendonk, T.A.; Makkee, M.; Gascon, J.; Kapteijn, F. Challenges in the
greener production of formates/formic acid, methanol, and dme by heterogeneously catalyzed CO2 hydrogenation processes.
Chem. Rev. 2017, 117, 9804–9838.
204. Dewangan, N.; Hui, W.M.; Jayaprakash, S.; Bawah, A.-R.; Poerjoto, A.J.; Jie, T.; Jangam, A.; Hidajat, K.; Kawi, S. Recent progress
on layered double hydroxide (ldh) derived metal-based catalysts for CO2 conversion to valuable chemicals. Catal. Today 2020,
356, 490–513.
Catalysts 2022, 12, 1484 29 of 32
205. Zhong, J.; Yang, X.; Wu, Z.; Liang, B.; Huang, Y.; Zhang, T. State of the art and perspectives in heterogeneous catalysis of CO2
hydrogenation to methanol. Chem. Soc. Rev. 2020, 49, 1385–1413.
206. Olah, G.A. Beyond oil and gas: The methanol economy. Angew. Chem. Int. Ed. 2005, 44, 2636–2639.
207. Goeppert, A.; Czaun, M.; Jones, J.-P.; Surya Prakash, G.K.; Olah, G.A. Recycling of carbon dioxide to methanol and derived
products—Closing the loop. Chem. Soc. Rev. 2014, 43, 7995–8048.
208. Yang, M.; Fan, D.; Wei, Y.; Tian, P.; Liu, Z. Recent progress in methanol-to-olefins (mto) catalysts. Adv. Mater. 2019, 31, 1902181.
209. Spencer, M.S. The role of zinc oxide in Cu/ZnO catalysts for methanol synthesis and the water–gas shift reaction. Top. Catal.
1999, 8, 259.
210. Behrens, M.; Studt, F.; Kasatkin, I.; Kühl, S.; Hävecker, M.; Abild-Pedersen, F.; Zander, S.; Girgsdies, F.; Kurr, P.; Kniep, B.-L.; et
al. The active site of methanol synthesis over Cu/ZnO/Al2O3 industrial catalysts. Science 2012, 336, 893–897.
211. Kühl, S.; Tarasov, A.; Zander, S.; Kasatkin, I.; Behrens, M. Cu-based catalyst resulting from a Cu,Zn,Al hydrotalcite-like com-
pound: A microstructural, thermoanalytical, and in situ xas study. Chem.—Eur. J. 2014, 20, 3782–3792.
212. Gao, P.; Li, F.; Zhan, H.; Zhao, N.; Xiao, F.; Wei, W.; Zhong, L.; Wang, H.; Sun, Y. Influence of Zr on the performance of
Cu/Zn/Al/Zr catalysts via hydrotalcite-like precursors for CO2 hydrogenation to methanol. J. Catal. 2013, 298, 51–60.
213. Mureddu, M.; Lai, S.; Atzori, L.; Rombi, E.; Ferrara, F.; Pettinau, A.; Cutrufello, M.G. Ex-ldh-based catalysts for CO2 conversion
to methanol and dimethyl ether. Catalysts 2021, 11, 615.
214. Gao, P.; Li, F.; Zhao, N.; Xiao, F.; Wei, W.; Zhong, L.; Sun, Y. Influence of modifier (Mn, La, Ce, Zr and Y) on the performance
of Cu/Zn/Al catalysts via hydrotalcite-like precursors for CO2 hydrogenation to methanol. Appl. Catal. A 2013, 468, 442–452.
215. Gao, P.; Zhong, L.; Zhang, L.; Wang, H.; Zhao, N.; Wei, W.; Sun, Y. Yttrium oxide modified Cu/ZnO/Al2O3 catalysts via hy-
drotalcite-like precursors for CO2 hydrogenation to methanol. Catal. Sci. Technol. 2015, 5, 4365–4377.
216. Stangeland, K.; Chamssine, F.; Fu, W.; Huang, Z.; Duan, X.; Yu, Z. CO2 hydrogenation to methanol over partially embedded cu
within Zn-Al oxide and the effect of indium. J. CO2 Util. 2021, 50, 101609.
217. Zhang, F.; Liu, Y.; Xu, X.; Yang, P.; Miao, P.; Zhang, Y.; Sun, Q. Effect of al-containing precursors on Cu/Zno/Al2O3 catalyst for
methanol production. Fuel Process. Technol. 2018, 178, 148–155.
218. Bahmani, M.; Vasheghani Farahani, B.; Sahebdelfar, S. Preparation of high performance nano-sized Cu/ZnO/Al2O3 methanol
synthesis catalyst via aluminum hydrous oxide sol. Appl. Catal. A 2016, 520, 178–187.
219. Zhao, F.; Fan, L.; Xu, K.; Hua, D.; Zhan, G.; Zhou, S.-F. Hierarchical sheet-like Cu/Zn/Al nanocatalysts derived from ldh/mof
composites for CO2 hydrogenation to methanol. J. CO2 Util. 2019, 33, 222–232.
220. Zhao, F.; Zhan, G.; Zhou, S.-F. Intercalation of laminar cu–al ldhs with molecular tcpp(m) (M = Zn, Co, Ni, and Fe) towards
high-performance CO2 hydrogenation catalysts. Nanoscale 2020, 12, 13145–13156.
221. Kim, J.; Jeong, C.; Baik, J.H.; Suh, Y.-W. Phases of Cu/Zn/Al/Zr precursors linked to the property and activity of their final
catalysts in CO2 hydrogenation to methanol. Catal. Today 2020, 347, 70–78.
222. Frusteri, L.; Cannilla, C.; Todaro, S.; Frusteri, F.; Bonura, G. Tailoring of hydrotalcite-derived cu-based catalysts for CO2 hydro-
genation to methanol. Catalysts 2019, 9, 1058.
223. Gao, P.; Li, F.; Xiao, F.; Zhao, N.; Sun, N.; Wei, W.; Zhong, L.; Sun, Y. Preparation and activity of Cu/Zn/Al/Zr catalysts via
hydrotalcite-containing precursors for methanol synthesis from CO2 hydrogenation. Catal. Sci. Technol. 2012, 2, 1447–1454.
224. Gao, P.; Xie, R.; Wang, H.; Zhong, L.; Xia, L.; Zhang, Z.; Wei, W.; Sun, Y. Cu/Zn/Al/Zr catalysts via phase-pure hydrotalcite-like
compounds for methanol synthesis from carbon dioxide. J. CO2 Util. 2015, 11, 41–48.
225. Gao, P.; Li, F.; Xiao, F.; Zhao, N.; Wei, W.; Zhong, L.; Sun, Y. Effect of hydrotalcite-containing precursors on the performance of
Cu/Zn/Al/Zr catalysts for CO2 hydrogenation: Introduction of Cu2+ at different formation stages of precursors. Catal. Today 2012,
194, 9–15.
226. Zhang, F.; Zhang, Y.; Yuan, L.; Gasem, K.A.M.; Chen, J.; Chiang, F.; Wang, Y.; Fan, M. Synthesis of Cu/Zn/Al/Mg catalysts on
methanol production by different precipitation methods. Mol. Catal. 2017, 441, 190–198.
227. Hou, X.-X.; Xu, C.-H.; Liu, Y.-L.; Li, J.-J.; Hu, X.-D.; Liu, J.; Liu, J.-Y.; Xu, Q. Improved methanol synthesis from CO2 hydrogena-
tion over CuZnAlZr catalysts with precursor pre-activation by formaldehyde. J. Catal. 2019, 379, 147–153.
228. Xiao, S.; Zhang, Y.; Gao, P.; Zhong, L.; Li, X.; Zhang, Z.; Wang, H.; Wei, W.; Sun, Y. Highly efficient Cu-based catalysts via
hydrotalcite-like precursors for CO2 hydrogenation to methanol. Catal. Today 2017, 281, 327–336.
229. Zhang, C.; Yang, H.; Gao, P.; Zhu, H.; Zhong, L.; Wang, H.; Wei, W.; Sun, Y. Preparation and CO2 hydrogenation catalytic
properties of alumina microsphere supported cu-based catalyst by deposition-precipitation method. J. CO2 Util. 2017, 17, 263–
272.
230. Kattel, S.; Ramírez, P.J.; Chen, J.G.; Rodriguez, J.A.; Liu, P. Active sites for CO2 hydrogenation to methanol on Cu/ZnO catalysts.
Science 2017, 355, 1296–1299.
231. Kurtz, M.; Wilmer, H.; Genger, T.; Hinrichsen, O.; Muhler, M. Deactivation of supported copper catalysts for methanol synthe-
sis. Catal. Lett. 2003, 86, 77–80.
232. Kühl, S.; Schumann, J.; Kasatkin, I.; Hävecker, M.; Schlögl, R.; Behrens, M. Ternary and quaternary Cr or Ga-containing ex-ldh
catalysts—Influence of the additional oxides onto the microstructure and activity of Cu/ZnAl2O4 catalysts. Catal. Today 2015,
246, 92–100.
233. Li, M.M.J.; Chen, C.; Ayvalı, T.; Suo, H.; Zheng, J.; Teixeira, I.F.; Ye, L.; Zou, H.; O’Hare, D.; Tsang, S.C.E. CO2 hydrogenation to
methanol over catalysts derived from single cationic layer CuZnGa LDH precursors. ACS Catal. 2018, 8, 4390–4401.
Catalysts 2022, 12, 1484 30 of 32
234. Zheng, H.; Narkhede, N.; Zhang, G.; Zhang, H.; Ma, L.; Yu, S. Highly dispersed cu catalyst based on the layer confinement
effect of Cu/Zn/Ga-ldh for methanol synthesis. Mol. Catal. 2021, 516, 111984.
235. Younas, M.; Loong Kong, L.; Bashir, M.J.K.; Nadeem, H.; Shehzad, A.; Sethupathi, S. Recent advancements, fundamental chal-
lenges, and opportunities in catalytic methanation of CO2. Energy Fuels 2016, 30, 8815–8831.
236. Wei, W.; Jinlong, G. Methanation of carbon dioxide: An overview. Front. Chem. Sci. Eng. 2011, 5, 2–10.
237. Guo, X.; Peng, Z.; Hu, M.; Zuo, C.; Traitangwong, A.; Meeyoo, V.; Li, C.; Zhang, S. Highly active Ni-based catalyst derived from
double hydroxides precursor for low temperature CO2 methanation. Ind. Eng. Chem. Res. 2018, 57, 9102–9111.
238. Pan, Q.; Peng, J.; Sun, T.; Wang, S.; Wang, S. Insight into the reaction route of CO2 methanation: Promotion effect of medium
basic sites. Catal. Commun. 2014, 45, 74–78.
239. He, L.; Lin, Q.; Liu, Y.; Huang, Y. Unique catalysis of ni-al hydrotalcite derived catalyst in CO2 methanation: Cooperative effect
between ni nanoparticles and a basic support. J. Energy Chem. 2014, 23, 587–592.
240. Sun, C.; Świrk, K.; Wierzbicki, D.; Motak, M.; Grzybek, T.; Da Costa, P. On the effect of yttrium promotion on ni-layered double
hydroxides-derived catalysts for hydrogenation of CO2 to methane. Int. J. Hydrog. Energy 2021, 46, 12169–12179.
241. Summa, P.; Samojeden, B.; Motak, M.; Wierzbicki, D.; Alxneit, I.; Świerczek, K.; Da Costa, P. Investigation of Cu promotion
effect on hydrotalcite-based nickel catalyst for CO2 methanation. Catal. Today 2022, 384–386, 133–145.
242. Xiao, X.; Wang, J.; Li, J.; Dai, H.; Jing, F.; Liu, Y.; Chu, W. Enhanced low-temperature catalytic performance in CO2 hydrogena-
tion over Mn-promoted NiMgAl catalysts derived from quaternary hydrotalcite-like compounds. Int. J. Hydrog. Energy 2021,
46, 33107–33119.
243. Aldana, P.A.U.; Ocampo, F.; Kobl, K.; Louis, B.; Thibault-Starzyk, F.; Daturi, M.; Bazin, P.; Thomas, S.; Roger, A.C. Catalytic
CO2 valorization into CH4 on Ni-based ceria-zirconia. Reaction mechanism by operando Ir spectroscopy. Catal. Today 2013, 215,
201–207.
244. Liu, J.; Bing, W.; Xue, X.; Wang, F.; Wang, B.; He, S.; Zhang, Y.; Wei, M. Alkaline-assisted Ni nanocatalysts with largely enhanced
low-temperature activity toward CO2 methanation. Catal. Sci. Technol. 2016, 6, 3976–3983.
245. Guo, X.; Gao, D.; He, H.; Traitangwong, A.; Gong, M.; Meeyoo, V.; Peng, Z.; Li, C. Promotion of CO2 methanation at low tem-
perature over hydrotalcite-derived catalysts-effect of the tunable metal species and basicity. Int. J. Hydrog. Energy 2021, 46, 518–
530.
246. Yin, L.; Chen, X.; Sun, M.; Zhao, B.; Chen, J.; Zhang, Q.; Ning, P. Insight into the role of fe on catalytic performance over the
hydrotalcite-derived Ni-based catalysts for CO2 methanation reaction. Int. J. Hydrog. Energy 2022, 47, 7139–7149.
247. Świrk, K.; Summa, P.; Wierzbicki, D.; Motak, M.; Da Costa, P. Vanadium promoted Ni(Mg,Al)O hydrotalcite-derived catalysts
for CO2 methanation. Int. J. Hydrog. Energy 2021, 46, 17776–17783.
248. Mebrahtu, C.; Krebs, F.; Perathoner, S.; Abate, S.; Centi, G.; Palkovits, R. Hydrotalcite based Ni–Fe/(Mg, Al)Ox catalysts for CO2
methanation—Tailoring fe content for improved co dissociation, basicity, and particle size. Catal. Sci. Technol. 2018, 8, 1016–
1027.
249. He, F.; Zhuang, J.; Lu, B.; Liu, X.; Zhang, J.; Gu, F.; Zhu, M.; Xu, J.; Zhong, Z.; Xu, G.; et al. Ni-based catalysts derived from Ni-
Zr-Al ternary hydrotalcites show outstanding catalytic properties for low-temperature CO2 methanation. Appl. Catal. B 2021,
293, 120218.
250. Zhang, Q.; Xu, R.; Liu, N.; Dai, C.; Yu, G.; Wang, N.; Chen, B. In Situ Ce-doped catalyst derived from NiCeAl-ldhs with enhanced
low-temperature performance for CO2 methanation. Appl. Surf. Sci. 2022, 579, 152204.
251. Lee, W.J.; Li, C.; Prajitno, H.; Yoo, J.; Patel, J.; Yang, Y.; Lim, S. Recent trend in thermal catalytic low temperature CO2 methana-
tion: A critical review. Catal. Today 2021, 368, 2–19.
252. Zhang, L.; Bian, L.; Zhu, Z.; Li, Z. La-promoted Ni/Mg-Al catalysts with highly enhanced low-temperature CO2 methanation
performance. Int. J. Hydrog. Energy 2018, 43, 2197–2206.
253. Ho, P.H.; de Luna, G.S.; Angelucci, S.; Canciani, A.; Jones, W.; Decarolis, D.; Ospitali, F.; Aguado, E.R.; Rodríguez-Castellón, E.;
Fornasari, G.; et al. Understanding structure-activity relationships in highly active la promoted Ni catalysts for CO2 methana-
tion. Appl. Catal. B 2020, 278, 119256.
254. Jiang, F.; Wang, S.; Liu, B.; Liu, J.; Wang, L.; Xiao, Y.; Xu, Y.; Liu, X. Insights into the influence of CeO2 crystal facet on CO2
hydrogenation to methanol over pd/CeO2 catalysts. ACS Catal. 2020, 10, 11493–11509.
255. Guo, X.; He, H.; Traitangwong, A.; Gong, M.; Meeyoo, V.; Li, P.; Li, C.; Peng, Z.; Zhang, S. Ceria imparts superior low temper-
ature activity to nickel catalysts for CO2 methanation. Catal. Sci. Technol. 2019, 9, 5636–5650.
256. Malkar, R.S.; Yadav, G.D. Selectivity engineering in one pot synthesis of raspberry ketone: Crossed aldol condensation of p-
hydroxybenzaldehyde and acetone and hydrogenation over novel Ni/Zn-La mixed oxide. ChemistrySelect 2019, 4, 2140–2152.
257. Novodárszki, G.; Onyestyák, G.; Barthos, R.; Wellisch, Á.F.; Thakur, A.J.; Deka, D.; Valyon, J. Guerbet alkylation of acetone by
ethanol and reduction of product alkylate to alkane over tandem nickel/Mg,Al-hydrotalcite and nickel molybdate/γ-alumina
catalyst systems. React. Kinet. Mech. Catal. 2017, 121, 69–81.
258. Fridrich, B.; Stuart, M.C.A.; Barta, K. Selective coupling of bioderived aliphatic alcohols with acetone using hydrotalcite derived
mg–al porous metal oxide and raney nickel. ACS Sustain. Chem. Eng. 2018, 6, 8468–8475.
259. Onyestyák, G.; Novodárszki, G.; Farkas Wellisch, Á.; Pilbáth, A. Upgraded biofuel from alcohol–acetone feedstocks over a two-
stage flow-through catalytic system. Catal. Sci. Technol. 2016, 6, 4516–4524.
260. Ramos, R.; Tišler, Z.; Kikhtyanin, O.; Kubička, D. Towards understanding the hydrodeoxygenation pathways of furfural–ace-
tone aldol condensation products over supported Pt catalysts. Catal. Sci. Technol. 2016, 6, 1829–1841.
Catalysts 2022, 12, 1484 31 of 32
261. Yang, J.; Li, S.; Li, N.; Wang, W.; Wang, A.; Zhang, T.; Cong, Y.; Wang, X.; Huber, G.W. Synthesis of jet-fuel range cycloalkanes
from the mixtures of cyclopentanone and butanal. Ind. Eng. Chem. Res. 2015, 54, 11825–11837.
262. Onyestyák, G.; Novodárszki, G.; Barthos, R.; Klébert, S.; Wellisch, Á.F.; Pilbáth, A. Acetone alkylation with ethanol over multi-
functional catalysts by a borrowing hydrogen strategy. RSC Adv. 2015, 5, 99502–99509.
263. Al-Wadaani, F.; Kozhevnikova, E.F.; Kozhevnikov, I.V. Pd supported on znii–criii mixed oxide as a catalyst for one-step syn-
thesis of methyl isobutyl ketone. J. Catal. 2008, 257, 199–205.
264. Nikolopoulos, A.A.; Jang, B.W.L.; Spivey, J.J. Acetone condensation and selective hydrogenation to MIBK on Pd and Pt hy-
drotalcite-derived Mg–Al mixed oxide catalysts. Appl. Catal. A 2005, 296, 128–136.
265. Das, N.N.; Das, R. Synthesis, characterization and activation of quaternary layered double hydroxides for the one-pot synthesis
of methyl isobutyl ketone. React. Kinet. Mech. Catal. 2010, 99, 397–408.
266. Hetterley, R.D.; Mackey, R.; Jones, J.T.A.; Khimyak, Y.Z.; Fogg, A.M.; Kozhevnikov, I.V. One-step conversion of acetone to
methyl isobutyl ketone over pd-mixed oxide catalysts prepared from novel layered double hydroxides. J. Catal. 2008, 258, 250–
255.
267. Shen, Y.; Yi, J.; Yan, Y.; Liu, D.; Fan, L.; Li, S. Hydrogenation and condensation of acetone over Ni/MgO–Al2O3 prepared from
hydrotalcite precursors. J. Chem. Eng. Jpn. 2016, 49, 656–662.
268. Basu, S.; Sarkar, J.J.; Pradhan, N.C. Selective synthesis of MIBK via acetone hydrogenation over Cu-Al mixed oxide catalysts.
Catal. Today 2021, 404, 182–189.
269. Li, X.; Sun, J.; Shao, S.; Hu, X.; Cai, Y. Aldol condensation/hydrogenation for jet fuel from biomass-derived ketone platform
compound in one pot. Fuel Process. Technol. 2021, 215, 106768.
270. Shao, S.; Hu, X.; Dong, W.; Li, X.; Zhang, H.; Xiao, R.; Cai, Y. Integrated C–C coupling/hydrogenation of ketones derived from
biomass pyrolysis for aviation fuel over Ni/Mg–Al–O/Ac bifunctional catalysts. J. Clean. Prod. 2021, 282, 124331.
271. Sheng, X.; Li, G.; Wang, W.; Cong, Y.; Wang, X.; Huber, G.W.; Li, N.; Wang, A.; Zhang, T. Dual-bed catalyst system for the direct
synthesis of high density aviation fuel with cyclopentanone from lignocellulose. AIChE J. 2016, 62, 2754–2761.
272. Sheng, X.; Li, N.; Li, G.; Wang, W.; Wang, A.; Cong, Y.; Wang, X.; Zhang, T. Direct synthesis of renewable dodecanol and
dodecane with methyl isobutyl ketone over dual-bed catalyst systems. ChemSusChem 2017, 10, 825–829.
273. Luggren, P.J.; Apesteguía, C.R.; Di Cosimo, J.I. Upgrading of biomass-derived 2-hexanol to liquid transportation fuels on Cu–
Mg–Al mixed oxides. Effect of Cu content. Fuel 2016, 177, 28–38.
274. Zhong, Y.; Zhou, B.; Wang, L. Fe/FeOx embedded in LDH catalyzing C-C bond forming reactions of furfural with alcohols in
the absence of a homogeneous base. Mol. Catal. 2020, 493, 111056.
275. Hernández, W.Y.; De Vlieger, K.; Van Der Voort, P.; Verberckmoes, A. NiCu hydrotalcite-derived mixed oxides as highly
selective and stable catalysts for the synthesis of β-branched bioalcohols by the guerbet reaction. ChemSusChem 2016, 9, 3196–
3205.
276. Li, J.; Lin, L.; Tan, Y.; Wang, S.; Yang, W.; Chen, X.; Luo, W.; Ding, Y.-J. High performing and stable Cu/NiAlOx catalysts for the
continuous catalytic conversion of ethanol into butanol. ChemCatChem 2022, 14, e202200539.
277. Pang, J.; Zheng, M.; He, L.; Li, L.; Pan, X.; Wang, A.; Wang, X.; Zhang, T. Upgrading ethanol to n-butanol over highly dispersed
Ni–MgAlO catalysts. J. Catal. 2016, 344, 184–193.
278. Song, J.; Huang, Z.-F.; Pan, L.; Li, K.; Zhang, X.; Wang, L.; Zou, J.-J. Review on selective hydrogenation of nitroarene by catalytic,
photocatalytic and electrocatalytic reactions. Appl. Catal. B 2018, 227, 386–408.
279. Wang, J.; Du, C.; Wei, Q.; Shen, W. Two-dimensional pd nanosheets with enhanced catalytic activity for selective hydrogenation
of nitrobenzene to aniline. Energy Fuels 2021, 35, 4358–4366.
280. Kowalewski, E.; Krawczyk, M.; Słowik, G.; Kocik, J.; Pieta, I.S.; Chernyayeva, O.; Lisovytskiy, D.; Matus, K.; Śrębowata, A.
Continuous-flow hydrogenation of nitrocyclohexane toward value-added products with CuZnAl hydrotalcite derived materi-
als. Appl. Catal. A 2021, 618, 118134.
281. Corma, A.; Serna, P. Chemoselective hydrogenation of nitro compounds with supported gold catalysts. Science 2006, 313, 332–
334.
282. Tan, Y.; Liu, X.Y.; Zhang, L.; Wang, A.; Li, L.; Pan, X.; Miao, S.; Haruta, M.; Wei, H.; Wang, H. ZnAl-hydrotalcite-supported
Au25 nanoclusters as precatalysts for chemoselective hydrogenation of 3-nitrostyrene. Angew. Chem. 2017, 129, 2753–2757.
283. Tan, Y.; Liu, X.Y.; Li, L.; Kang, L.; Wang, A.; Zhang, T. Effects of divalent metal ions of hydrotalcites on catalytic behavior of
supported gold nanocatalysts for chemoselective hydrogenation of 3-nitrostyrene. J. Catal. 2018, 364, 174–182.
284. Zhao, J.; Yuan, H.; Li, J.; Bing, W.; Yang, W.; Liu, Y.; Chen, J.; Wei, C.; Zhou, L.; Fang, S. Effects of preparation parameters of
nial oxide-supported Au catalysts on nitro compounds chemoselective hydrogenation. ACS Omega 2020, 5, 7011–7017.
285. Lu, Q.; Liu, J.; Ma, L. Recent advances in selective catalytic hydrogenation of nitriles to primary amines. J. Catal. 2021, 404, 475–
492.
286. Tichit, D.; Medina, F.; Durand, R.; Mateo, C.; Coq, B.; Sueiras, J.E.; Salagre, P. Use of ni containing anionic clay minerals as
precursors of catalysts for the hydrogenation of nitriles. In Studies in Surface Science and Catalysis; Blaser, H.U., Baiker, A., Prins,
R., Eds.; Elsevier: Amsterdam, The Netherlands, 1997; Volume 108, pp 297–304.
287. Coq, B.; Tichit, D.; Ribet, S. Co/Ni/Mg/Al layered double hydroxides as precursors of catalysts for the hydrogenation of nitriles:
Hydrogenation of acetonitrile. J. Catal. 2000, 189, 117–128.
288. Cao, Y.; Niu, L.; Wen, X.; Feng, W.; Huo, L.; Bai, G. Novel layered double hydroxide/oxide-coated nickel-based core–shell nano-
composites for benzonitrile selective hydrogenation: An interesting water switch. J. Catal. 2016, 339, 9–13.
Catalysts 2022, 12, 1484 32 of 32
289. Cao, Y.; Zhang, H.; Dong, J.; Ma, Y.; Sun, H.; Niu, L.; Lan, X.; Cao, L.; Bai, G. A stable nickel-based catalyst derived from layered
double hydroxide for selective hydrogenation of benzonitrile. Mol. Catal. 2019, 475, 110452.
290. Sheng, M.; Yamaguchi, S.; Nakata, A.; Yamazoe, S.; Nakajima, K.; Yamasaki, J.; Mizugaki, T.; Mitsudome, T. Hydrotalcite-sup-
ported cobalt phosphide nanorods as a highly active and reusable heterogeneous catalyst for ammonia-free selective hydro-
genation of nitriles to primary amines. ACS Sustain. Chem. Eng. 2021, 9, 11238–11246.
... In LDHs, -OH groups on lamellae provide Brønsted acid-base sites, and metal cations in the lattice create Lewis acid-base sites. LDHs' catalytic activity is due to these acid-base sites, which enable them to perform well in green catalytic reactions, such as condensation, hydrogenation, and biomass conversion [18,19]. 3 of 20 In addition, doping LDHs with transition metal cations introduces additional Lewis acidbase sites that enhance their acid-base and catalytic activities, thus improving their catalytic performance for redox reactions during acid-base catalytic transformations [20]. ...
... In LDHs, -OH groups on lamellae provide Brønsted acid-base sites, and metal cations in the lattice create Lewis acid-base sites. LDHs' catalytic activity is due to these acid-base sites, which enable them to perform well in green catalytic reactions, such as condensation, hydrogenation, and biomass conversion [18,19]. In addition, doping LDHs with transition metal cations introduces additional Lewis acid-base sites that enhance their acid-base and catalytic activities, thus improving their catalytic performance for redox reactions during acid-base catalytic transformations [20]. ...
Article
Full-text available
Layered double hydroxides (LDHs) have emerged as promising catalysts for various acid–base catalytic reactions. Due to their unique structure and regulatable dual acid–base properties, they offer more environmentally friendly and sustainable alternatives to traditional liquid acid and base catalysts. This study introduces the structural composition, preparation methods, and acid–base catalytic properties of LDH-based catalysts. Recent application progress in LDHs and rehydrated LDHs, LDH-based metal nanocatalysts, and LDH-based mixed metal oxide catalysts used as solid acid–base catalysts in acid–base green catalytic conversion is reviewed. The challenges and prospects of LDH-based catalysts as green and sustainable catalysts are summarized and proposed.
... In LDHs, -OH groups on lamellae provide Brønsted acid-base sites, and metal cations in the lattice create Lewis acid-base sites. LDHs' catalytic activity is due to these acid-base sites, which enable them to perform well in green catalytic 3 reactions, such as condensation, hydrogenation, and biomass conversion [16,17]. In addition, doping LDHs with transition metal cations introduces additional Lewis acid-base sites that enhance their acid-base and catalytic activities, thus improving their catalytic performance for redox reactions during acid-base catalytic transformations [18]. ...
Preprint
Full-text available
Layered double hydroxides (LDHs) have emerged as promising catalysts for various acid-base catalytic reactions. Due to their unique structure and regulatable dual acid-base properties, they offer more environmentally friendly and sustainable alternatives to traditional liquid acid and base catalysts. This study introduces the structural composition, preparation methods, and acid-base catalytic properties of LDHs-based catalysts. Recent application progress of LDHs and rehydrated LDHs, LDHs-based metal nanocatalysts, and LDHs-based mixed metal oxide catalysts used as solid acid-base catalysts in acid-base green catalytic conversion is reviewed. The challenges and prospects of LDHs-based catalysts as green and sustainable catalysts are summarized and proposed.
... Hydrotalcites (HTs) can be considered as precursors of basic supports. In the calcined form they have been studied as catalytic precursors or supports for methanation of C0 2 [14]: hydrogenation and hydrogenolysis [15], hydrodesulfurization HDS [16], dry reforming: [17]. HTs were also used as precursors of catalysts in the HDOP: for example, Shailesh et al. [18] studied Pd and Ru supported on a Mg-Al hydrotalcites; the authors reported conversions up to 95% of Jatropha oil. ...
Article
Full-text available
Ni-Mg–Al ternary hydrotalcites (HTs) with molar ratio M²⁺ (Ni²⁺ + Mg2 +)/Al = 3, and different content of Ni, Ni0Mg3, Ni2Mg1, Ni2Mg1, Ni3Mg0 were prepared. The HTs were calcined and then reduced to generate catalysts based on metallic Ni, which, were used to produce green diesel from a mixture of stearic and palmitic acids at 320 °C and 1400 psia of H2. A reaction scheme was proposed; the main products were alkanes n-C17 and n-C15. The prepared catalysts carry out, mainly, reactions of decarboxylation (DCX) and decarbonylation (DCN), and some hydrocracking. The most active catalyst was Ni2Mg1 because it forms small clusters of nickel, apparently stabilized on MgO or NiO-MgO spinel like structures. Graphical Abstract
Article
The conversion of biomass-derived furfuryl alcohol into 1,2-pentanediol, a high-value fine chemical with wide applications, is of high research and commercial value. In this study, Ce-doped CuCeMgAl mixed metal oxides...
Article
Full-text available
The catalytic conversion of ethanol into butanol is one of crucial reactions in biorefinery for the production of renewable fuels and chemicals. The development of highly durable non‐noble metal‐based catalysts is of great significance but poses big challenges. Herein, we prepared a series of NiAl‐hydrotalcite derived Cu/NiAlOx catalysts via the hydrothermal precipitation method. Cu addition to the system shows an enhanced reactivity in the continuous catalytic conversion of ethanol into butanol. The 0.75 %Cu/NiAlOx catalyst afford a sustained ethanol conversion of ∼35 % and butanol selectivity of ∼45 % within 1000 h at 250 °C, 3 MPa N2. Complementary characterization studies reveal that Cu addition into the catalysts could not only provide redox sites for facilitating the dehydrogenation/hydrogenation processes, but also modulate the base/acid property for suppressing the side reactions. The excellent stability might stem from the enhanced interaction between Cu and Ni species, as evidenced by EDX‐mapping and H2‐TPR. We believe this work will help to rational design and develop of high‐performing and durable non‐noble metal catalysts for ethanol coupling to higher alcohols.
Article
Full-text available
Supported Ru catalysts were prepared by using different Ru precursors and examined for the hydrogenation of benzaldehyde to CHM. The catalyst prepared from Ru3(CO)12 precursor and HT support exhibited high yield of CHM. Moderate acidic and basic nature of HT was favourable to control the selectivity for CHM. The physicochemical properties analysis revealed that highly dispersed Ru nanoparticles were effective for the hydrogenation of benzaldehyde. RuCO/HT catalyst was tolerant to different functional groups and was stable until 7th cycle of recycle study.
Article
Full-text available
The catalytic hydrogenolysis of 5-hydroxymethylfurfural (HMF) to 2,5-dimethylfuran (DMF) is a promising route towards sustainable liquid fuels with a high energy density. Herein, a novel CuCoNiAl-containing mixed metal oxide catalyst (CuCoNiAl-MMO) was prepared by calcination a layered double hydroxide (LDH) precursor in N2 at 500 °C, then applied for the catalytic hydrogenolysis of HMF to DMF. The effects of reaction time, reaction temperature and hydrogen pressure on DMF selectivity were investigated. Under relatively mild reaction conditions (180°C, 1.0 MPa H2, 6.0 h), CuCoNiAl-MMO showed both a high initial activity and selectivity for hydrogenolysis of HMF to DMF, with HMF conversion rate of 99.8% and DMF selectivity of 95.3%. Catalysts characterization studies using scanning electron microscopy (SEM), transmission electron microscopy (TEM), X-ray diffraction (XRD) and X-ray photoelectron spectroscopy (XPS) revealed the presence of various metal oxides and metallic copper on the surface of the CuCoNiAl-MMO catalyst, with the presence of mixed metal-oxide-supported metallic Cu nanoparticles being responsible good hydrogenolysis activity of the catalyst for selective DMF synthesis.
Article
The worldwide economic development, population expansion, and technological advancements contribute to a rise in global primary energy consumption. Since fossil fuels now provide around 85% of the energy requirement, a significant quantity of greenhouse gases is released, leading to climate change. To meet the pledges of the Paris agreement, a promising and potential alternative to fossil fuels needs to be commercialised. Therefore, numerous industries recognise hydrogen (H2) as a clean and stable energy source for decarbonisation or de-fossilisation. Around 90% of the world's H2 produced is grey in nature and produced from reforming fossil-based fuels. However, the future of H2 energy lies in its green, blue, and turquoise spectra due to the carbon capture scheme and corresponding clean and sustainable H2 production methodology. The fundamental goal of this research is to learn more about various low-carbon H2 generating systems. In comparison to fossil-based H 2 , green H 2 is a costly option. Blue H2 offers several appealing characteristics; however, the carbon capture utilisation and storage (CCUS) technology are expensive and blue H2 is not carbon-free. The current CCUS technology can only store and catch between 80 and 95% of CO2. Further, it examines worldwide actions related to the H2 development policy. In addition, a debate based on the colour spectrum of H2 was established to classify the purity of H2 generation. Further, the existing obstacles, advancements, and future directions of low-carbon H2 production technologies, including fossil fuel-based and renewable-based H2 , are explored to foster the growth of the low-carbon H2 economy.
Chapter
With the increasing world population and huge demand for fossil fuels, the emission of CO2 into the atmosphere has reached a level of about 417 ppm. Such an enormous amount of greenhouse gas on our planet has been the primary source to cause climate change and environmental devastations, including droughts, floods, wildfires, hurricanes, etc. The depletion of petroleum resources and stringent ecological regulation at present must inspire the whole world to work jointly for the development of an alternative and green fuel to replace fossil fuels. The top agenda to overcome this global concern is the sequestration of carbon from pre-and post-combustion as well as oxyfuel combustion. Thus, this chapter comprehensively reviews the current progress and advancements of CO2 conversion into valuable fuels, including methane, dimethyl ether, methanol, gasoline, and others. However, further intensive research is needed to obtain economic and commercially viable conversion operations.
Article
In this work, a strategy for rational design of high-performance heterogenous supported Ru nanocatalysts over mixed metal oxides for aqueous phase partial hydrodeoxygenation (HDO) of guaiacol in the batch reactor was developed taking advantage of supramolecular structure of Zn-Al layered double hydroxides (LDHs) with different interlayer anions. It was interestingly found that in the case of phosphotungstate intercalated ZnAl-LDH as the support precursor, strong interactions between Ru species and intercalated LDH-derived mixed metal oxide supports could be formed, in addition to the formation of surface reductive ZnWOx component and P-O species. Notably, as-fabricated supported Ru nanocatalyst showed much higher activity in the aqueous phase HDO of guaiacol to produce cyclohexanol at an extraordinarily low Ru/guaiacol molar ratio of 0.0009, along with a higher cyclohexanol yield of ∼86% under reaction conditions (250 °C, 2.0 MPa initial H2 pressure), compared to other two supported Ru ones over mixed metal oxides derived from tungstate- and nitrate-intercalated ZnAl-LDH precursors (cyclohexanol yield of 29.7 and 10.1%, respectively). Studies on structure-property correlations revealed that surface defective and acidic sites facilitated the activation of guaiacol and phenol intermediate, thereby enabling the catalyst with a strong ability to break Cring−OCH3 bond and promote the ring hydrogenation. This finding provides a promising way to tune surface structures of metal-based catalysts for the upgrading of biomass.
Article
Selective hydrogenolysis of lignin is an attractive strategy for valorizing lignin to fuels/value-added chemicals and developing novel and highly active heterogeneous catalysts is crucial. Herein, a highly dispersed Ni-Ru/Al2O3 catalyst prepared by a coprecipitation method using layered double hydroxide as precursor was reported. Various lignin model compounds were used to evaluate the catalyst activity. The results show that all the model compounds can be completely transformed into arenes and phenol below 150 °C using isopropanol as hydrogen source, and the latter was easily reduced to cyclohexanol. Kinetics study reveals that apparent activation energies increase in the order: benzyl phenyl ether (90.6 kJ/mol) < phenethoxybenzene (126.7 kJ/mol) < diphenyl ether (172.2 kJ/mol). 2D HSQC NMR analysis exhibits that most of C–O linkages in lignin were cleaved during hydrogenolysis over Ni-Ru/Al2O3. The results indicate that the bimetallic Ni-Ru catalyst has a high activity for catalyzing hydrogenolysis of lignin and model compounds. The catalyst was characterized with multiple tools. The results suggest that Ni and Ru nanoparticles in the catalyst are highly dispersed and a strong interaction exists between Ni and Ru, which should be responsible for promoting the hydrogenolysis of lignin and model compounds.
Article
A series of layered Cu-Co-Al hydrotalcite (LDHs) catalysts with different Cu/Co molar ratio were synthesized and applied for the hydrogenation of furfuryl alcohol to 1,5-pentanediol. The results indicated that the Cu0.2Co5.8/Al2O3 catalyst displayed an excellent catalytic efficiency compared to monometallic Co/Al2O3 or Cu/Al2O3 catalyst. The yield of 1,5-pentanediol reached 44.7% at 140°C, 4 MPa H2 within 2 h. In-situ XPS, H2-TPR, CO2-TPD analysis revealed that the high efficiency of this catalyst was attributed to the synergistic catalysis between the surface Cu⁰, CoOx and the basicity of metal oxides. Typically, Cu⁰ sites activate dissociated hydrogen, the basic sites anchor furfuryl alcohol by -OH to form an alcoholate species, inducing furfuryl alcohol with C2 oblique adsorption on the catalyst surface. The adjacent CoOx species adsorb and activate C2=C3 group in furfuryl alcohol. Subsequently, C2=C3 bond of furfuryl alcohol was partially hydrogenated, facilitating the cleavage C2-O1 bond to produce 1,5-pentanediol.
Article
Hydrogenation of levulinic acid (LA) to γ-valerolactone (GVL) is an important reaction in biomass upgradation to produce value-added chemicals. In this work, we have successfully prepared well dispersed nanocrystalline Pt-HT catalyst and found that this catalyst is showing excellent performance for the levulinic acid hydrogenation to produce GVL. Pt-HT catalyst showed ∼100% conversion with ∼100% selectivity of GVL at low temperature with external hydrogen in a batch reactor in aqueous medium. Different reaction parameters like hydrogen pressure, temperature, reaction time and loading of Pt were studied in detail. The catalyst was successfully reused up to 7 times without any loss of the catalytic activity. The catalyst was characterized by XRD, SEM, HR-TEM, BET, XPS, TPR, TPD and DFT study. It was observed that the highly dispersed metallic Pt is the active species to achieve the good conversion and selectivity of GVL even at room temperature and the catalyst is showing very high stability and recyclability during hydrogenation, making it very promising truly heterogeneous catalyst for practical γ-valerolactone production via hydrogenation of levulinic acid.