Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

pubs.acs.

org/EF Article

Lignin Catalytic Depolymerization for Guaiacols and Alkylphenols


over Co−Zr/Sepiolite under Supercritical Ethanol
Mingqiang Chen,* Zinan Zhou, Yishuang Wang,* Zheng Fang, Defang Liang, Longyang Li,
Haosheng Xin, Chang Li, Gang Yuan, and Jun Wang
Cite This: Energy Fuels 2024, 38, 4256−4272 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: Lignin catalytic depolymerization (LCD) in supercritical


ethanol for the preparation of liquid fuels and phenolic monomers is
Downloaded via CHONNAM NATL UNIV on March 12, 2024 at 14:23:58 (UTC).

one of the promising strategies for the valorization of lignocellulosic


biomass. This paper reported sepiolite (SEP)-supported Co−Zr
catalysts (Co−Zr/SEP) with different Co/Zr ratios for LCD in
supercritical ethanol. The characterization results revealed that the
cobalt oxides interacted with the SEP framework to create abundant
pores, which increased the accessibility of the active sites with lignin
fragments. Furthermore, the Zr additive promoted the conversion of
Co(OH)2 to Co3O4 to increase Lewis acid sites (LAS) and oxygen
vacancies (VO), which could enhance the adsorption and stabilization of
reactive intermediates. Among the samples, 3Co1Zr/SEP exhibited the
superior surface area and the highest LAS and abundant VO. The
experimental results showed that the Co/Zr ratio significantly affected
the yields of liquid products and phenolic monomers, and 3Co1Zr/SEP achieved the outstanding liquid product yield (60.11%) and
the optimal yields of guaiacols (445.2 mg/g lignin) and alkylphenols (109.8 mg/g lignin) under the optimal reaction conditions.
Additionally, 3Co1Zr/SEP significantly reduced the apparent activation energy for breaking β-O-4 bonds to 71.46 from 89.94 kJ·
mol−1 over SEP by model experiments; furthermore, the possible reaction pathways of LCD over 3Co1Zr/SEP were proposed.

1. INTRODUCTION and 1,4-dioxane coupled with hydrogen pressures (2−4 MPa)


Lignin, as one of the abundant natural renewable resources, has to achieve the lignin depolymerization into liquid fuels and
been extensively researched in the field of fabricating aromatic significantly inhibit the coke formation. Shu et al.21 proposed a
products, biodegradable materials, and liquid fuels due to the solvent system of metal chloride cooperated with methanol to
specificity of its composition and structure.1−3 It has been achieve lignin depolymerization with controllable product
found that lignin consists of coumarin (H), mustard alcohol distribution under a hydrogen pressure of 4 MPa. This catalytic
(S), and pineal (G) groups, which are randomly connected by system suppressed the polymerization of ester structures in
C−C/C−O bonds (4-O-5, α-O-4, β-O-4, β−β, β-1, etc.) to lignin and improved the conversion efficiency of lignin to fuel.
form natural aromatic polymer compounds with complex In addition, some other solvents including n-hexane,
three-dimensional networks.4−8 In recent years, on the diisopropyl ether, ethyl acetate, methanol, and acetone
strength of the structural characteristics of lignin, it has been cooperated with hydrogen pressure (3−5 MPa) were
extensively demonstrated that lignin catalytic depolymerization developed for lignin depolymerization.22−24 Although the
(LCD) for liquid fuels and cooperative production of phenolic HHV of the liquid product or monomer selectivity was clearly
monomers is one of the promising methods for highly efficient improved by employing the complex solvent systems and the
utilization of lignin.9,10
high hydrogen pressure, the high cost of solvent and
However, it had been found that the produced highly
downstream product separation and the urgent need for
reactive intermediates during the LCD process underwent an
irreversible condensation reaction and formed abundant coke,
which restricts the advancement of LCD technology.11,12 It has Received: November 22, 2023
been demonstrated that LCD under a special solvent system Revised: January 30, 2024
and a reducing atmosphere could inhibit coke and increase the Accepted: January 30, 2024
HHV (higher heating value) of liquid products and the Published: February 13, 2024
selectivities of different monomers.13−16 Li et al.17−20
developed an advanced solvent system consisting of methanol

© 2024 American Chemical Society https://doi.org/10.1021/acs.energyfuels.3c04637


4256 Energy Fuels 2024, 38, 4256−4272
Energy & Fuels pubs.acs.org/EF Article

Scheme 1. Preparation of the SEP Precursor (a), Synthetic Process of Co−Zr/SEP (b), and Lignin Depolymerization Process
(c)

delicate catalysts that can efficiently activate hydrogen become phenols, whose selectivity reached to 85.61%. Therefore, it can
a great challenge. be found that the option of suitable active components
Recently, it has been found that depolymerization of lignin particularly metals/metallic oxides and coupling them with
in the ethanol system, especially supercritical ethanol, to unique supports is considered as very important for designing
replace the molecular hydrogen source is among the most and preparing advanced catalysts for LCD.
competent methods for producing liquid fuels and aromatic Non-noble-metal-based catalysts including Ni-, 3 0
products.25−32 It has been demonstrated that supercritical W-,35,36,46,54 Co-,19,40,41 and Mo-based catalysts5,32−34,53 have
ethanol not only could enhance heat transfer and improve the been utilized in lignin conversion owing to their superior
solubility of lignin/intermediates but could also act as a breaking capability for C−C/C−O of lignin and hydro-
nucleophile to break ether bonds in lignin.26−28 Huang et al.29 genation function. Especially, the phase of Co3O4 with a spinel
investigated the LCD using Cu−Mg−Al mixed oxide catalysts structure in Co-based catalysts has multiple functions of Lewis
in supercritical ethanol and confirmed that the hydrogen acid sites (LAS) derived from the Co2+ atoms located in the
produced from ethanol was involved in breaking the lignin C− tetrahedron and Brønsted acid sites (BAS) originated from the
O bonds and the deoxygenation reaction of monomers and Co3+ atoms located in the octahedron.37,38 In addition, the
oligomers. Although the LCD technology under supercritical Co2+ and Co3+ atoms of Co3O4 could be substituted by other
ethanol can inhibit the char formation and realize the atoms to generate more oxygen vacancies (VO).57 It is
hydrogenation of the reaction process, improving the lignin generally accepted that the acidic sites (LAS and BAS) and
depolymerization efficiency and the selectivity of special VO could improve the C−O bond breaking and hydrogenation
monomers is still one of the research emphases of lignin ability.39,40 Qi et al.41 used Ni doping into Co3O4 to tailor the
valorization. It has been widely accepted that the design and NixCo3−xO4 catalysts' Brønsted and Lewis acid centers and
preparation of advanced catalysts is an effective measure to employed the as-prepared catalysts into LCD. It was found
realize this process. Tran et al.31 designed copper supported on that the Ni species optimized the synergistic effect between
the chromium-based metal−organic framework (MOF) (Cu@ acid centers including BAS and LAS and active sites to
MIL-101(Cr)) for LCD in supercritical ethanol and found that promote the lignin cleavage and the excellent selectivity to
the MOF framework increased the Cu particle dispersion and aromatic monomers (up to 90.30 wt %). It can be found that
stabilized the Cu sites that could induce the selective breaking the formation of Vo in Co-based catalysts could activate the
of ether bonds in lignin. This resulted in the concentration of lignin C−O bonds and reduce the dissociation energy of C−O
guaiacols obtaining a yield of 38.5%. In our previous work,33,34 bonds, which led to further deoxygenation and hydrogenation.
the catalysts of Mn-modified Mo/sepiolite (SEP) were It has been demonstrated that addition of ZrO2 with high
synthesized and used for LCD in supercritical ethanol. oxygen storage capacity into metal-based catalysts could
Different from more than 20 alkylation products produced enhance the generation of abundant VO to improve the
by Mo/SEP in supercritical ethanol for LCD, the inclusion of catalytic performance.42−46 Garciá et al.46 reported the
manganese enhanced the proficiency to selectively cleave the catalytic production of γ-valerolactone from furfural over
β-O-4 bonds and repressed the isomerization and alkylation SnZr/Y catalysts with different ratios of Sn and Zr and found
processes of the reactive intermediates, resulting in reduction that the formed Zr−O−Sn increased the active species
of the production of alkylation products and the enrichment of dispersity and the addition of Zr gave rise to more BAS and
4257 https://doi.org/10.1021/acs.energyfuels.3c04637
Energy Fuels 2024, 38, 4256−4272
Energy & Fuels pubs.acs.org/EF Article

LAS. However, the use of Co−Zr-based catalysts for exploiting pink suspension was filtered, washed using deionized water, and dried
LCD has been little studied to our knowledge. at 105 °C for 12 h. The Co/SEP catalyst was obtained by calcining
In addition to this, the selection of excellent support the solid at 500 °C for 4 h under static air with a heating ramp of 4
materials could optimize catalytic activity and increase the °C/min. The content of Co was fixed at 20 wt % in this process.
Zr/SEP was prepared employing the impregnation method.
service life of catalysts. Natural clay sepiolite (SEP) has a
Typically, the predetermined amount of Zr(NO3)4·5H2O was
special nanolayered structure, and its abundant pore structure dissolved in 50 mL of deionized water. Then, 1 g of purified SEP
could anchor and disperse the active metals.33,34 Based on its was added, and the mixture was stirred for 8 h at 60 °C. After that, the
structural characteristics, the SEP as a support or catalytic mixture was dehumidified by an evaporation process at 90 °C and
component had been widely applied into various catalytic dried at 105 °C for 12 h. The Zr/SEP catalyst was obtained by
reactions.47−49 In addition, a series of SEP-supported metal- calcining the solid at 500 °C for 4 h under static air with a heating
based catalysts in our previous reports had been prepared and ramp of 4 °C/min. The content of Zr was fixed at 20 wt % in this
used for LCD.33−35,50 process.
In this paper, for further improving the yield of the phenolic 2.3. Catalyst Characterization. N2 adsorption−desorption
monomer, SEP-supported Co−Zr catalysts (Co−Zr/SEP) analysis was conducted using a BELSORP-mini II surface area
with different mass ratios of Co and Zr were prepared for analyzer (made in Japan) at a temperature of −196 °C. The specific
surface area was calculated employing the Brunauer−Emmet−Teller
LCD in supercritical ethanol. The catalysts that had been
(BET) technique, while the pore volume and pore size distribution
prepared underwent characterization by using various methods were measured via the Barrett−Joyner−Halenda (BJH) method.
such as XRD, HRTEM, FTIR, Raman analysis, XPS, EPR, Py- The actual contents of Si, Al, Co, and Zr from the catalysts were
IR, and NH3-TPD to reveal the effect of Co/Zr mass ratios on calculated by inductively coupled plasma-optical emission spectrom-
the structural properties of as-prepared catalysts. In addition, etry (ICP-OES) on an Agilent ICP-OES 5110 (USA).
elemental analysis, GC-MS, and 2D-HSQC NMR were used to X-ray diffraction was carried out to analyze the crystal phases of the
analyze the liquid-phase products. Model substances with catalysts prepared using a Rigaku Corporation SmartLab (3 kW) X-
characteristic lignin structures were utilized to elucidate the ray diffractometer from Japan. This used Co Kα as the radiation
possible reaction mechanism and pathways for LCD. source, set at 40 kV and 40 mA, with λ = 0.17902 nm and a scanning
step of 0.02° for 2θ = 5−70°.
2. EXPERIMENTAL SECTION FTIR spectra of the samples were acquired using a Nicolet iS50
spectrometer (USA). FTIR spectroscopy was conducted within the
2.1. Materials. Sepiolite (SEP) was purchased from Shanlin Shiyu wavenumber range of 4000−400 cm−1. Prior to measurement, all
Mineral Products Co., Ltd. (Hunan, China). Kraft lignin (Indulin AT) samples were uniformly diluted in spectroscopic-grade KBr powder
was procured from Mead Westvaco (Shanghai, China). The solvents (4−8 wt % in KBr).
and chemical reagents utilized in this study were procured from Raman spectra of samples were recorded by a Renishaw inVia
Sinopharm Chemical Reagent Co., Ltd. (located in Shanghai, China), (UK) instrument with a 532 nm laser (Ar+) at a maximum power of
and all solvents and chemical reagents were of AR (analytical reagent) 100 mW.
purity. Co(NO3)2·6H2O (cobalt nitrate hexahydrate) and Zr(NO3)4· X-ray photoelectron spectroscopy (XPS) was utilized to investigate
5H2O (zirconium nitrate pentahydrate) were employed as the the states of elements in the samples using a Thermo Fisher
precursors of Co and Zr, respectively. 2-Phenoxy-1-phenylethanol ESCALAB 250Xi X-ray photoelectron spectrometer (UK), with Al
(PP) was used in the model experiments. Kα (1486.68 eV) as the photon source.
2.2. Catalyst Preparation. Initially, 20 g of SEP was treated in The electron paramagnetic resonance (EPR) of the samples was
100 mL of 5 mol/L NaOH solution at 60 °C for 2 h. It was then measured by using a Bruker A300 EPR spectrometer (Germany).
transferred to a hydrothermal synthesis reactor and maintained at 180
FEI Tecnai G2 F20 (USA) transmission electron microscopy
°C for 12 h. The solid product obtained was washed with deionized
(TEM), high-resolution transmission electron microscopy
water multiple times before being dried overnight at 150 °C. The
(HRTEM), and TEM-EDX were used to examine the microstructure
recovered solids were subsequently calcined in an air atmosphere at
and distribution of elements of the samples at an accelerating voltage
500 °C with a heating ramp of 4 °C/min over 4 h to obtain the
purified SEP. The preparation procedure can be observed in Scheme of 200 kV.
1a. The reduction process of the catalysts was analyzed using hydrogen
The catalysts comprising Co−Zr/SEP were created using the temperature-programmed reduction (H2-TPR) analysis on an
impregnation−precipitation technique. The diagram detailing the Autochem II 2920 manufactured in the USA. Hydrogen consumption
process is shown in Scheme 1b. Typically, the predetermined amount was measured with a thermal conductivity detector (TCD), and the
of Co(NO3)2·6H2O was dissolved in 50 mL of deionized water, and 1 sample temperature was ramped from 10 to 800 °C at a rate of 10
g of purified SEP was added. The mixture was then stirred for 2 h at °C/min.
room temperature. Then, a certain amount of urea (generally the The acid properties of the sample surface were detected using
molar ratio of Co(NO3)2·6H2O and urea was 0.5) as the precipitant ammonia temperature-programmed desorption (NH3-TPD) on an
was added into above turbid liquid and continuously stirred for 8 h at Autochem II 2920 (USA). Before chemisorption, the samples
90 °C to form the pink suspension. Subsequently, the above pink underwent pretreatment in helium at 200 °C for 1 h. Following
suspension was then mixed with Zr(NO3)4 solution (the predeter- this, the temperature was brought down to 50 °C in helium, and a
mined amount of Zr(NO3)4·5H2O was dissolved in 50 mL of mixture flow of 5% NH3/He was injected until the surface of the
deionized water) and agitated for 8 h at 60 °C. Following this, the samples became saturated. The sample was subjected to a stream of
solution underwent dehydration through an evaporation process at 90 helium at a flow rate of 30 mL min−1 and maintained at 120 °C for an
°C and was left to dry overnight at 105 °C. The resulting solid hour. The desorption process for NH3 was carried out by gradually
material was subjected to calcination at 500 °C (with a heating ramp increasing the temperature to 700 °C at a rate of 10 °C/min.
of 4 °C/min) for 4 h in stagnant air to produce the Co−Zr/SEP The acid site types of the samples were determined using infrared
catalyst. Herein, the metal content of Co+Zr was fixed to 20 wt %, the spectroscopy of pyridine adsorption (Py-IR) on a Nicolet 6700
mass ratios of Co/Zr were 3/1, 1/1, and 1/3, and the corresponding spectrometer from the United States. The catalyst was dried at 120 °C
catalysts were named as 3Co1Zr/SEP, 1Co1Zr/SEP, and 1Co3Zr/ for 12 h and then cooled to room temperature before the introduction
SEP. of pyridine. Pyridine vapor was allowed into the chamber for 24 h to
Co/SEP was prepared employing the precipitation method. As in achieve a saturation adsorption of pyridine. Subsequently, the samples
the preparation process of Co−Zr/SEP described above, the resulting were rinsed using a He stream for 2 h at 150 °C.

4258 https://doi.org/10.1021/acs.energyfuels.3c04637
Energy Fuels 2024, 38, 4256−4272
Energy & Fuels pubs.acs.org/EF Article

Figure 1. N2 adsorption−desorption isotherms of all catalysts.

2.4. Depolymerization Experiments of Lignin. The LCD weight of the monomer


monomer yield(mg/g lignin) = × 100
experiment process is described in Scheme 1c. Initially, a mix of 1.0 g weight of lignin
of lignin, 0.2 g of the catalyst, and 30 mL of ethanol was stirred using (3)
a magnetic stirrer in a 100 mL stainless-steel reactor (Shanghai 2.5. Lignin Model Compound Experiments. The process of
Yanzheng Experimental Instrument Co., Ltd., Shanghai, China). The catalytic conversion of PP as the lignin model compound was
reactor was then purged with N2 to remove air and filled with 0.2 MPa consistent with lignin depolymerization (reaction conditions: 0.1 g of
N2. Subsequently, the reactor was heated to the target temperature the lignin model compound PP, 0.02 g of the catalyst, and 30 mL of
and stirred continuously at 450 rpm for the established time. After ethanol). The conversion and monomer selectivity of PP were
completing the reaction, the reactor was immediately placed in ice calculated by the following equations:
water and cooled to room temperature. The liquid and solid phases conversion of PP(conv. %)
were separated by vacuum filtration to obtain the respective products. mole of the residual substrate
The remaining ethanol in the liquid product was eliminated by using a = (1 ) × 100%
mole of the initial substrate (4)
rotary evaporator at 40 °C. The liquid product was obtained, while
the solid product (char and catalyst) was washed with acetone and mole of the product
monomer selectivity(%) = × 100%
dried at 105 °C for 12 h. The yields of liquid, solid, and monomer mole of the reacted substrate
products were calculated using the following formula: (5)
On the premise of the Arrhenius equation, the reaction rate and the
yield of the liquid product (%) apparent activation energy (Ea) of the conversion reaction of 2-
weight of the liquid product phenoxy-1-phenylethanol (PP) were calculated using the following
= × 100% formulas, assuming it to be a pseudo-first-order reaction:
weight of lignin (1)
ln(1 conv. ) = kt (6)

yield of the solid product (%) Ea


ln k = ln A
yield of the solid product weight of the catalyst RT (7)
= × 100%
weight of lignin where conv. represents the conversion of the model compound, k is
(2) the rate constant, R signifies the constant of molar gas (8.314 J·mol−1·

4259 https://doi.org/10.1021/acs.energyfuels.3c04637
Energy Fuels 2024, 38, 4256−4272
Energy & Fuels pubs.acs.org/EF Article

K−1), T denotes the reaction temperature, and A represents the pre- dispersed on the SEP surface, resulting in creating abundant
exponential factor. regenerated pores. Meanwhile, for Zr/SEP, the addition of Zr
2.6. Analysis of the Depolymerization Product. The clearly reduced the SBET and Vpore to 4.83 m2/g and 0.01 cm3/
elemental contents of carbon, hydrogen, oxygen, nitrogen, and sulfur g, respectively, in comparison with SEP. This might because
in both lignin and liquid products were determined by employing a
Vario EL III elemental analyzer (Germany). The higher heating value
the impregnation process of Zr4+ precursors led to the entrance
(HHV) was obtained according to Dulong’s formula: HHV (MJ/kg) of Zr4+ into slit-pores of SEP. As compared with Co/SEP, it
= 0.3491 × C + 1.1783 × H + 0.1005 × S − 0.1034 × O − 0.0151 × was observed that the SBET and Vpore of Co−Zr/SEP catalysts
N − 0.0211 × ash.10 were obviously decreased with the decrease of Co/Zr mass
The gas chromatography−mass spectrometry (GC-MS) technique ratios. Among Co−Zr/SEP catalysts, 3Co1Zr/SEP had the
was utilized to analyze both the liquid and model conversion highest SBET (198.11 m2/g) and Vpore (0.716 cm3/g). This
products. A Thermo Scientific ISQ 7000 MS detector from the USA further confirmed the above analyses about the effect of Co
and a TG-5SILMS capillary column (30 m × 0.25 mm × 0.25 μm) and Zr species on textural properties. Therefore, it was
were used. To begin with, the oven was kept at a consistent reasonable to conclude that in Co−Zr/SEP catalysts, Co
temperature of 50 °C for 3 min followed by a gradual increase to 280
°C at 10 °C min−1. The end product, the yield of the monomer, was
species mainly led to form irregular mesopores on the SEP
determined using a known amount of acetophenone as an internal surface, while Zr species uniformly distributed in the
standard. mesopores.
A Bruker AVANCE NEO 600 MHz spectrometer (Germany) was The phase and crystal structures of all samples were analyzed
used to record the 2D-HSQC NMR spectra of both lignin and liquid using XRD, and the outcomes are depicted in Figure 2. For
products. Dimethyl sulfoxide-d6 (DMSO-d6) was utilized as a sample
solvent, while the remaining DMSO solvent peak (δC/δH 39.5/2.50)
was employed as an internal reference.

3. RESULTS AND DISCUSSION


3.1. Characterization Analysis of As-Prepared Cata-
lysts. The as-prepared catalysts' textural properties were
discussed by N2 adsorption−desorption, and the results are
presented in Figure 1 and Figure S1. It was found that SEP and
Zr/SEP exhibited typical IV adsorption−desorption isotherms
with H3 hysteresis loops in accordance with the IUPAC
classification, indicating that they had slit-pores formed by the
stacking of sheetlike nanoparticles, which are the typical
morphological features of SEP.35 However, the as-prepared
catalysts containing Co species presented typical IV
adsorption−desorption isotherms with H4 hysteresis loops,
and these loops were shrunk with the increase of the Zr
content, indicating that their pores were irregularly meso-
porous predominantly derived from the cobalt oxides. In
addition, the appearance of “tails” at P/P0 = 0.01 in all as- Figure 2. XRD patterns of all catalysts.
prepared Co-based catalysts indicated the existence of a
minimal amount of micropores and irregular mesopores.51 For SEP, the characteristic diffraction peaks at 2θ = 8.99, 14.09,
further analysis, the as-prepared catalysts' specific surface area 17.87, 24.5, 26.7, 30.06, 42.6, 19.85, 31.88, 34.78, and 37.35°
(SBET) and average pore volume/size (Vpore/Dpore) are listed in correspond to SEP.50 Upon incorporating Co and Zr, it was
Table 1. For Co/SEP, its SBET and Vpore were dramatically apparent that the intensities of the peaks of SEP were
significantly reduced due to the interaction between the Co/Zr
Table 1. Textural Parameters for All Catalysts species and SEP to cover the surface of SEP. Furthermore, for
SBET Vpore Dpore Co Zr Co/SEP, the peaks at 2θ = 31.2, 36.8, and 44.8° are related to
samples (m2/g) (cm3/g) (nm) (wt %) (wt %) Si/Al the (220), (311), and (400) crystal planes of Co3O4, and the
SEP 19.69 0.066 13.02 1.41 peak at 2θ = 37.8° corresponds to the characteristic diffraction
Co/SEP 244.43 0.911 13.23 19.11 1.32 peak of Co(OH)2, which was attributed to the incomplete
3Co1Zr/ 198.11 0.716 13.57 14.29 4.81 1.39 conversion of Co(OH)2 to Co3O4 during the calcination
SEP process. For Zr/SEP, characteristic diffraction peaks of ZrO2
1Co1Zr/ 190.50 0.522 10.48 8.95 9.65 1.35 with a typical mixture of monoclinic and tetragonal phases
SEP
were detected at 2θ = 28.2, 29.9, 31.6, and 50.2°. In addition,
1Co3Zr/ 99.15 0.112 4.52 4.74 13.86 1.39
SEP in Co−Zr/SEP catalysts, only the characteristic peaks of
Zr/SEP 4.83 0.010 10.55 19.13 1.44 Co3O4 were observed without finding the peaks of Co(OH)2
and ZrO2. It was implied that the homogeneous dispersion of
Zr into the mesopores and the Zr additive promoted the
increased to 244.43 m2/g and 0.911 cm3/g, respectively, as conversion of Co(OH)2 into Co3O4 mainly via forming the
compared with those of SEP. It might be attributed to that the interaction between Zr and Co components.
Co2+ precursors that interacted with OH− species during the All samples' metal−oxygen bonds and surface groups were
precipitation to form Co(OH)2 species, and then, they were further investigated by FTIR analysis (Figure S2). The
converted to cobalt oxides during the calcination process. The vibrational bands at 3700−3300 cm−1 were derived from the
formed cobalt oxides interacted with the SEP framework and vibrations of the OH− groups, the band at 1644 cm−1 was
4260 https://doi.org/10.1021/acs.energyfuels.3c04637
Energy Fuels 2024, 38, 4256−4272
Energy & Fuels pubs.acs.org/EF Article

associated with SEP structured water. The vibrational bands at characteristic peaks of Co3O4 at 686 cm−1 (A1g) ascribed to the
1010, 470, and 541 cm−1 were the symmetrical and octahedral sites (CoO6). This demonstrated that the original
asymmetrical vibrations of Si−O bonds of silica in SEP octahedral coordination environment in Co3O4 was altered
tetrahedral sheets.35 When the Co and Zr species were and consequently increased the number of defect sites.
introduced, the vibrational bands of the SEP at 1010 and 470 Notably, since no diffraction peak of Co(OH)2 was observed
cm−1 were blueshifted, which indicated that there were in 3Co1Zr/SEP by XRD analysis, the Raman peak appearing at
interactions between the Co and Zr species and the surface 595 cm−1 might be the CoOx oxides produced by the
of the SEP. For Co−Zr/SEP catalysts, the bands at 571 cm−1 transformation of Co(OH)2 to Co3O4 (under the calcination
correspond to the vibrations of Co3+ atoms at the octahedral process as described in the section Catalyst Preparation). This
sites of Co3O4, and the bands at 664 cm−1 were ascribed to the had also been reported by Gao et al.52
vibrations of Co2+ atoms at the tetrahedral positions of The Co/SEP and 3Co1Zr/SEP were further selected for
Co3O4.47 Meanwhile, these bands were not observed in Co/ HRTEM coupled with elemental mapping to study the
SEP, which is probably attributed to the fact that the Co(OH)2 influence of Zr species on the microscopic morphology and
species as detected by XRD affected the vibrations of Co3+ and crystal composition, and the obtained results are displayed in
Co2+ atoms in Co3O4 species. In addition, except for Zr/SEP, Figures 4 and 5. In accordance with the results of elemental
there was no appearance of bands at 526 and 746 cm−1 mapping in Figure 4a, the Co, Al, Si, and O elements uniformly
attributed to the signals of Zr−O bonds of ZrO2 that appeared distributed in Co/SEP. For 3Co1Zr/SEP as shown in Figure
in the other Zr-containing catalysts. It was suggested that the 5a, the above four elements were uniformly distributed, but the
Zr species did not exist in the form of ZrO2 in Co−Zr/SEP Zr element was partially overlapped. This was mainly caused
catalysts, and the reason for this situation might be attributed by the Zr predominantly distributed into the irregular
to the interaction of ZrO2 with Co species. mesopores formed by the deposition of Co species on the
To further investigate the effect of the addition of Co and Zr SEP surface. In the HRTEM images of Co/SEP displayed in
on the skeletal structure of SEP, Raman characterization was Figure 4b, according to our previous report,47 the fibrillar
performed on all catalysts, and the obtained spectra are phyllosilicate species were clearly observed, which was due to
displayed in Figure 3. It can be found that the intensities of the interaction of Co species with Si species in SEP to form
Co-phyllosilicate structures. Since the formation of Co-
phyllosilicate changed the original structure of SEP, the Al
atoms in SEP were also exposed and distributed near the Co-
phyllosilicate. Therefore, the characteristic lattice spacings of
0.304 and 0.447 nm corresponding to the Al2O3 (022) and
(002) crystal faces were observed as shown in Figure 4c. In
addition, the Co3O4 with the given lattice spacing of 0.286 nm
corresponding to the (220) crystal plane in the neighborhood
of the Co-phyllosilicate in Figure 4d was observed. In the
HRTEM images of 3Co1Zr/SEP, Co-phyllosilicate fibers and
the characteristic crystal planes of Al2O3 (122), (022), and
(002) were also observed (as presented in Figure 5b,c). In
addition, the lattice spacings of 0.202, 0.244, and 0.286 nm
attributed to the characteristic crystal planes of Co3O4 (220),
(311), and (400), respectively, were observed in Figure 5d,e,
which are in line with the results of XRD. It was suggested that
the addition of Zr generated the interaction between Zr and
Co species, which promoted the exposure of Co3O4 crystal
planes and provided more reactive sites for LCD. In addition
Figure 3. Raman patterns of all catalysts.
to this, the lattice spacing of 0.219 nm corresponding to the
(201) crystal plane of ZrO2 was also detected.
SEP Raman peaks at 252, 451, and 698 cm−1 decreased or H2-TPR measurements were performed to further study the
disappeared with the incorporation of Co and Zr. This interaction of Co with Zr and SEP. As shown in Figure S3, the
demonstrated that the formation of the interaction of Co and reduction profile of Co/SEP exhibited the broad peak at 320
Zr species with SEP weakens the Raman signal of SEP. For °C, which corresponded to the reduction of Co(OH)2 and Co
Co/SEP, the Raman bands at 482, 522, 616, and 686 cm−1 oxides, and the peak at about 730 °C was attributed to the
correspond to the E2g, F22g, F2g, and A1g phonon modes of the reduction of Co-phyllosilicate.47,55 According to the re-
Co3O4 phase, respectively.41 In addition, the peak at around ports,44,56 there was almost no occurrence of the ZrO2
595 cm−1 corresponds to the mixture of Co(OH)2 and Co3O4. reduction peak in the H2-TPR profile. Consequently, the
The Co(OH)2 species in Co/SEP was not fully converted to weak reduction peak at 600 °C in the Zr/SEP TPR profile
Co3O4 generally attributed to Co(OH)2 present in the outer mainly corresponded to the reduction of ZrO2−x species that
layer of the catalyst, which was easily determined by the were produced by the interaction between SEP and the Zr
Raman laser. For Zr/SEP, the Raman peak at 263 cm−1 is species but not the ZrO2 reduction. For Co−Zr/SEP catalysts,
specific to the tetragonal phase of ZrO2, and the Raman peaks the peaks in the 250−450 °C region corresponded to the
at 480 and 634 cm−1 were attributed to the ZrO2 monoclinic reduction of Co3O4 to Co0 (Co3O4 → CoO → Co0), and the
phase. In Co−Zr/SEP, the characteristic bands attributed to peaks in the high-temperature region (>600 °C) correspond to
the phonon modes of the Co3O4 were all detected, while the the reduction of Co-phyllosilicate. Obviously, the reduction
addition of Zr species resulted in the blueshift of the Raman temperature of Co-phyllosilicate was significantly decreased
4261 https://doi.org/10.1021/acs.energyfuels.3c04637
Energy Fuels 2024, 38, 4256−4272
Energy & Fuels pubs.acs.org/EF Article

Figure 4. Corresponding element mapping (a) and HRTEM images (b−d) of Co/SEP.

Figure 5. Corresponding element mapping (a) and HRTEM images (b−e) of 3Co1Zr/SEP.

with the addition of Zr, which suggested that the interaction reduced with the decrease of Co/Zr ratios. Based on the results
between Co species and Zr species decreased the Co- of XRD and Raman analysis, this might be attributed to the
phyllosilicate species. This further demonstrated that the combined action of the Zr additive promoting the trans-
addition of Zr species resulted in more conversion of Co formation of Co(OH)2 into CoOx species and the decrease of
species to Co3O4 species rather than the formation of Co- the Co content in the catalyst. Additionally, in Co−Zr/SEP
phyllosilicate, which was consistent with the XRD and Raman catalysts, the relative contents of Co3+ and Co2+ of Co3O4
results. species increased with the decrease of the Co/Zr ratio,
To explore the surface chemical states of the different suggesting that the inclusion of Zr species facilitated the
elements and the interactions among the various components, creation of Co3O4, which was consistent with the results of
XPS characterization was conducted for all catalysts, and the Raman analysis and H2-TPR.
obtained XPS spectra of Co 2p, Zr 3d, and O 1s are shown in Figure 6b exhibits the Zr 3d XPS spectra for all catalysts. In
Figure 6. It was seen that the Co 2p XPS spectra were split into the control group of Zr/SEP, the peaks at about 184.2 and
two major peaks at about 781.4 and 797.1 eV accompanied 181.9 eV were related to Zr3+ species according to the reports
with satellite (sat.) peaks at about 786.5 and 803.4 eV, which in refs 44 and 56. Generally, Zr(NO3)4·5H2O as the precursor
are attributed to the Co 2p1/2 and Co 2p3/2 originated from the salt of the Zr additive could be completely converted to ZrO2
split of spin orbits of the Co 2p orbital in Figure 6a. under 4 h of calcination at 500 °C in an air atmosphere, which
Furthermore, following deconvolution treatment, the Co 2p was the necessary process for catalyst preparation in this paper.
XPS spectrum could be divided into three doublets. The Therefore, the appearance of Zr3+ species in Zr/SEP was
coupled peaks at 779.9 and 795.2 eV were correlated to Co3+ attributed to the interaction of Zr species with the SEP
of Co3O4 species, and the binary peaks at 781.2 and 797.5 eV framework, which confirmed the existence of electron transfer
were attributed to Co2+ of Co3O4 species, while the conjugated between the SEP framework and Zr. For 3Co1Zr/SEP, the two
peaks at 782.5 and 798.5 eV corresponded to a mixture of main peaks at about 185.0 and 182.8 eV were related to Zr
Co(OH)2 and CoOx species (2 < x < 3).41,52 It was found that 3d3/2 and Zr 3d5/2 orbits, demonstrating the existence of Zr4+
Co(OH)2 and CoOx in Co/SEP were the major existential species according to the reports in refs 44 and 56.
states of the Co species. Compared with Co/SEP, the peak Furthermore, the peaks of Zr 3d3/2 and Zr 3d5/2 in other
area of the mixture of Co(OH)2 and CoOx species gradually catalysts shifted straightly to lower binding energies with the
4262 https://doi.org/10.1021/acs.energyfuels.3c04637
Energy Fuels 2024, 38, 4256−4272
Energy & Fuels pubs.acs.org/EF Article

Figure 6. XPS of (a) Co 2p, (b) Zr 3d, and (c) O 1s energy regions of catalysts. (d) EPR spectra of catalysts.

decrease of the Co/Zr ratio, which suggested that valence state species not approaching the SEP surface; therefore, only Zr4+
of Zr decreased gradually with the Zr addition. The correlation species were found. With the reduction of the Co content, the
between the change in the valence state of Zr species and the SEP surface was not completely covered, and then, some Zr
Co/Zr ratio can be clarified here. It should be pointed out that species interacted with SEP leading to the Zr 3d XPS peak
in Co−Zr/SEP catalysts, the Co species were first supported moving into lower binding energies such as in 1Co1Zr/SEP
on SEP then introduced with Zr species during catalyst and 1Co3Zr/SEP.
preparation. In 3Co1Zr/SEP, the sufficient Co species first Figure 6c shows the XPS spectra of O 1s of all samples. After
contacted and covered the SEP surface and resulted in the Zr undergoing deconvolution treatment, the XPS spectra of O 1s
4263 https://doi.org/10.1021/acs.energyfuels.3c04637
Energy Fuels 2024, 38, 4256−4272
Energy & Fuels pubs.acs.org/EF Article

Figure 7. (a) NH3-TPD profile and (b) Py-IR spectra of SEP, Co/SEP, and 3Co1Zr/SEP.

could be resolved into four distinct peaks: 530.3 (lattice categorized into three temperature regions, which correspond
oxygen), 531.1 (Vo, oxygen vacancy), 531.8 (surface hydroxyl to weak, medium, and strong acid sites. Moreover, the HN3-
oxygen), and 532.4 eV (adhering water). The formation of TPD profiles demonstrated that the medium acidic sites of the
adhering water on the catalyst surface was basically attributed SEP surface were predominant, and the total acidity was 0.31
to its SBET, and the amount of adsorbed water was positively mmol NH3/g. For Co/SEP, the surface acid sites increased
correlated with the SBET of the catalyst. Therefore, the adhering significantly in terms of weak acid sites and medium acid sites
water content of various catalysts as presented in Figure 6c in comparison with SEP. Generally, the rise in weak acidic sites
exhibited a positive correlation with their specific surface areas was attributed to the introduction of Co active species that
(listed in Table 1). The peak at 531.8 eV represented the promoted the weak absorption of NH3 molecules.55 Mean-
surface hydroxyl oxygen of the catalyst, and it is known that the while, the augment of medium acid sites was ascribed to the
surface hydroxyl groups are correlated with the amount of increase of −OH groups originated from Co(OH)2 that
BAS.33,35 In addition, Figure 6c displays the surface relative resulted in the medium absorption of NH3 molecules on the
proportion of various O species, it was found that the VO surface of the catalyst.39,41 With the addition of Zr into Co/
content of 3Co1Zr/SEP was increased as compared with Co/ SEP, it was observed that the distribution of weak and medium
SEP, which is generally attributed to the conversion of acid sites indicated almost no changes in 3Co1Zr/SEP
Co(OH)2 to CoOx. With the increase of the Zr content, the Zr compared with Co/SEP. However, the total acidity decreased
species promoted the conversion of Co species to Co3O4, and to 0.47 mmol NH3/g for 3Co1Zr/SEP from 0.62 mmol NH3/g
thus, the VO content of 1Co1Zr/SEP further decreased. It was for Co/SEP. This was mainly because the Zr additive
noteworthy that the VO content of 1Co3Zr/SEP was promoted the conversion of Co(OH)2 species to CoOx as
significantly increased, but its CoOx content was very low confirmed by XRD. Another interesting finding was that the
(as confirmed by Co 2p XPS). This suggested that the strong acid sites were detected in 3Co1Zr/SEP, which was
increased VO did not originate from the CoOx species but from mainly assigned to absorption of NH3 molecules on the
the Zr species.58 coordination-unsaturated Zrδ+ or electron-deficient metal
In addition, electron paramagnetic resonance (EPR) was oxides (CoOx) based on the report of Ranaware et al.59
utilized to validate the influence of Zr addition on VO Py-IR as presented in Figure 7b recognized the Brønsted
concentration. As a result of the generation of VO, the cations
(BAS) and Lewis acid sites (LAS) and determined their
became undercoordinated, leading to a decrease in the electron
relative amounts summarized in Table 2. The vibrational bands
density of the system. This caused the electrons to become
at 1447−1450 and 1599−1610 cm−1 were attributed to the
localized in the center of the cations, which then captured
LAS, the vibrational bands at 1548 and 1650 cm−1 were
them and acquired magnetism;51 the EPR signal could
attributed to the BAS, and the band at 1490 cm−1 was
represent the concentration of VO. The EPR signal with g =
2.003 corresponded to VO as demonstrated in Figure 6d. The
3Co1Zr/SEP signal intensity was higher than that of Co/SEP, Table 2. Acid Site Types and Density of Catalysts
and the signal intensity of 1Co1Zr/SEP was less than that of Determined by NH3-TPD and Py-IR
3Co1Zr/SEP, which confirmed the conclusion that the VO was
acid sites
related to the CoOx content. Furthermore, 1Co3Zr/SEP had (μmol Py/g)
the highest signal intensity, which further supported the above
catalyst total acidity (mmol NH3/g) Lewis Brønsted L/B
XPS discussion result.
NH3-TPD was employed to study the amount of surface SEP 0.31 66.5 2.5 26.6
acid sites of as-prepared catalysts, and the obtained profiles are Co/SEP 0.62 26.4 1.9 13.9
displayed in Figure 7a. The NH3 desorption peaks could be 3Co1Zr/SEP 0.47 216.2 4.3 50.3

4264 https://doi.org/10.1021/acs.energyfuels.3c04637
Energy Fuels 2024, 38, 4256−4272
Energy & Fuels pubs.acs.org/EF Article

Figure 8. Liquid product and solid product yield over different catalysts (a), selectivity of different monomers over different LP (b), and monomer
yield over different LP (c). Reaction conditions: 1 g of lignin, 0.2 g of the catalyst, 30 mL of ethanol, 0.2 MPa N2 initial pressure, 290 °C, 4 h, 450
rpm.

attributed to the mixture of BAS and LAS.41,50 For SEP, the over 3Co1Zr/SEP promoting the cleavage of the lignin C−C
peaks at 1447−1450 and 1599−1610 cm−1 were significantly and C−O bonds and their derived intermediates. Additionally,
stronger than those at 1548 and 1650 cm−1, which indicated the recyclability of 3Co1Zr/SEP was investigated, and the
that the amount of LAS on the SEP surface was obviously results are shown in Figure S4. The liquid yield decreased after
higher than that of BAS. Compared to SEP, the L/B ratio of each cycle, and the liquid product yield decreased to 41.59%
Co/SEP was decreased due to the presence of partially after five cycles, indicating that the active sites of the catalyst
unconverted Co(OH)2 on the surface of Co/SEP. In addition, were changed. Figure S4b,c indicates that the species of the
the amount of LAS and BAS was reduced for Co/SEP catalysts were changed, in which the CoOx active phase almost
compared to SEP. This was due to the low transmittance of disappeared and transformed to Co3O4. In addition, the
fuscous Co/SEP during testing, which resulted in much lower decrease in the amount of acid from the spent catalysts was
amounts of LAS and BAS than the actual amounts.34 The also one of the reasons for the deactivation of the catalysts.
elevated ratio of L/B for 3Co1Zr/SEP compared with Co/SEP The GC-MS analysis coupled with an internal standard
was attributed to the incorporation of Zr, which promoted the method was conducted for further evaluating the composition
conversion of Co(OH)2 and reduced the amount of −OH on of the monomer in the liquid products. For convenience, the
the surface. liquid products (LP) that were obtained by using SEP, Co/
3.2. LCD Experiments of Co−Zr/SEP Catalysts. SEP, 3Co1Zr/SEP, 1Co1Zr/SEP, 1Co3Zr/SEP, and Zr/SEP
3.2.1. Catalyst Performance and Liquid Product Character- were named as LP-1−6, respectively. The obtained GC-MS
ization. In order to investigate the LCD performance of as- spectra are shown in Figure S5, and the specific information on
prepared catalysts, the reaction was performed for 4 h at 290 monomers is listed in Table S1. As a whole, the monomers in
°C, and the yields of liquid and solid products are displayed in these liquid products reached to 33, and these compounds
Figure 8. With SEP as a catalyst, the yield of liquid products could be divided into three categories, namely, guaiacols (red),
was only 27.44%, but the yielded solid product reached to alkylphenols (orange), and others (black). Furthermore, the
71.76%. After the individual addition of Co and Zr, the yields selectivities and yields of the three categories are exhibited in
of liquid products over Co/SEP and Zr/SEP were increased to Figure 8b,c. The guaiacol selectivity in LP-1 reached to 88.61%
46.71 and 29.62%, respectively. Meanwhile, the corresponding along with a yield of 243.1 mg/g lignin. Our previous work had
yields of solid products were slightly reduced to 52.09 and demonstrated that the acidic sites exposed on the SEP surface
66.28%, respectively. This is mainly attributed to the could promote the lignin C−C and C−O bond break-
introduction of the Co and Zr-containing species enabling age.33−35,50 It was found that the selectivity of guaiacols in
the activation of ethanol to produce more in situ hydrogens, the LP-2 and LP-6 slightly decreased to 78.33 and 89.32%,
which enabled the hydrogenation of reactive intermediates and respectively. Meanwhile, the selectivity of alkylphenols in LP-2
reduced the production of char. It had been confirmed that increased to 15.05%, while that in LP-6 decreased to 8.19%.
Co-based catalysts could promote the performance of aqueous- These results demonstrated that Co species exhibited a more
phase reforming of ethanol hydrogen production.61 For Co− significant promotion effect on LCD in comparison with Zr
Zr/SEP, it was revealed that only 3Co1Zr/SEP showed the additives. It was observed that the selectivity of guaiacols in
best catalytic performance, in which the yield of liquid LP-3 decreased to 74.06%, but the selectivity of alkylphenols
products reached to 60.11%, while the solid product yield obviously increased to 18.27 as compared with LP-2 and LP-6.
decreased to 38.49%. It was mainly attributed to the higher Meanwhile, as compared with LP-3, the selectivity of guaiacols
SBET, the highly dispersed metal sites, and abundant acidic sites of LP-4 decreased to 70.29%, while the selectivity of
4265 https://doi.org/10.1021/acs.energyfuels.3c04637
Energy Fuels 2024, 38, 4256−4272
Energy & Fuels pubs.acs.org/EF Article

Table 3. Elemental Composition of Lignin and the Liquid Product


elemental contents (wt %)
samples C H O N S ash H/C O/C HHV (MJ/kg)
lignin 57.84 4.63 22.98 0.89 2.28 2.12 0.08 0.40 23.44
LP-2 71.29 7.72 18.87 0.60 0.26 0.17 0.11 0.26 32.05
LP-3 71.94 8.01 17.96 0.62 0.22 0.14 0.11 0.25 32.71

Figure 9. FTIR patterns (a) and 1H NMR spectra (b) of lignin and LP-3.

alkylphenols increased to 22.85%. However, it was noteworthy the catalyst, and Zr species did not promote LCD significantly
that the selectivity of guaiacols of LP-5 increased to 85.31% and inhibited the promotion of Co species for LCD.
while that of alkylphenols decreased to 11.23% with the further In addition, elemental analysis of the lignin and the liquid
increase of the Zr content. These results confirmed that the product under the three reaction conditions was performed to
addition of Zr could promote the alkylation reaction during determine the elemental content of the samples. The higher
LCD, but the superabundant Zr content did not facilitate the heating value (HHV) was calculated and is summarized in
alkylation reaction. Table 3. Compared to lignin, the value of H/C of the LP
It should be pointed out that the yields of guaiacols and samples increased, while the value of O/C decreased, which
alkylphenols in LP-3 all reached to the highest values, which indicated that effective depolymerization of lignin was achieved
were 445.2 and 109.8 mg/g lignin, respectively, as presented in in the presence of the catalyst. Compared to LP-2, LP-3
Figure 8c. This suggested that there was a synergistic effect showed a lower O content and higher C and H contents, which
between Co and Zr additives. Based on above discussion, it was related to the increased products of alkylation. Compared
could be concluded that Co species could promote the to lignin, the HHV of LP-3 obtained by 3Co1Zr/SEP
cleavage of lignin C−C and C−O bonds and had facilitating increased significantly to 32.71 from 23.44 MJ/kg of lignin.
To further investigate the change in structural information
effects on in situ hydrogen production. In addition, the
from lignin to the liquid product, FTIR, 1H NMR, and 2D-
addition of Zr resulted in more LAS and VO for the catalyst, in
HSQC NMR analyses were conducted on lignin and the LP-3.
which LAS could promote the cleavage of C−C and C−O
Figure 9a illustrates the FTIR spectra of lignin and LP-3, and
bonds, and VO could adsorb and stabilize the reactive the specific characteristic vibrations are given in Table S2. The
intermediates and inhibit the condensation reaction. In bands at 2960, 2920, and 2875 cm−1 correspond to the
summary, the synergistic effect between Co and Zr promoted stretching vibrations of methyl, methylene, and methyne
the generation of more active intermediates, which led to the groups, respectively, and the band at 1460 cm−1 was attributed
increase of subsequent hydrogenation and alkylation reactions, to the bending vibrations of the methylene functional groups.33
resulting in the optimal LCD performance for 3Co1Zr/SEP. Furthermore, the band at 1510 cm−1 corresponds to the
VO was also one of the reasons for the highest yield of skeleton vibration of the aromatic ring.33,60 Meanwhile, the
guaiacols (445.2 mg/g lignin), where VO adsorbed and bands at 1120 cm−1 are attributed to the C−O bond stretching
stabilized the reactive intermediates in the LCD process, vibrations in the syringyl unit.33 The bands at 1085 and 1030
inhibited the occurrence of polycondensation reactions, and cm−1 are correlated to the primary and secondary alcohol
facilitated further hydrogenation or alkylation reactions of the groups stretching vibrations, respectively.33,60 Compared to
intermediates.40,41,51 When 1Co3Zr/SEP was used as the lignin, LP-3 showed stronger vibrational bands at 2960, 2920,
catalyst, which was almost the same as the product 2875, and 1460 cm−1. The weakened signals in the LP-3 band
composition of LP-6, the main reason for the decrease in the at 1510 cm−1 might be due to the breakage of the connecting
alkylation reaction at this time was that Zr species dominated bond of the structural unit of the lignin, and the products were
4266 https://doi.org/10.1021/acs.energyfuels.3c04637
Energy Fuels 2024, 38, 4256−4272
Energy & Fuels pubs.acs.org/EF Article

Figure 10. 2D-HSQC NMR spectra of lignin before the reaction (a,c) and LP-3 after the reaction (b,d).

more concentrated. The disappearance of the band at 1120 [δC/δH 90−135/6.0−8.0] correspond to the aliphatic side-
cm−1 in the lignin further suggested the breakage of the chain region and the aromatic region, respectively. For lignin,
connecting bond of the lignin structural unit. Furthermore, the the signals of Aα (δC/δH 73.2/4.76, orange), Aβ (δC/δH 84.2/
enhancement of the band signals located at 1085 and 1030 4.30, orange), and Aγ (δC/δH 59.8/3.80, orange) correspond-
cm−1 demonstrated an increase of primary and secondary ing to the β-O-4 bonds of lignin could be seen. Additionally,
alcohol groups in LP-3. the observed signals of Bγ (brown), Cα, Cβ (pink), and Fγ
Similarly, the 1H NMR spectra of lignin and LP-3 are shown (green) were attributed to C−O bonds in the side-chain region
in Figure 9b and divided into three regions: the methyl and of lignin. For LP-3, the signals of Aα, Aβ, Aγ above-mentioned
methylene region, methoxy region, and aromatic region.51 The
disappeared, which indicated that the β-O-4 and side-chain C−
enhanced signal in the methyl and methylene regions of LP-3,
O bonds were broken. In addition, the increased signal in the
the disappearance of part of the signal in the methoxy and
Csp3-H region [δC/δH 10−50/0−2.5] for LP-3 as presented in
aromatic region of lignin, and the more concentrated signal in
the methoxy and aromatic region of LP-3 were observed. The Figure S7 was attributed to the hydrogenation that allowed the
results discussed above all indicated that the C−C and C−O substitution of oxygen by hydrogen to form an alkylated
bonds used to connect the lignin basic units were sufficiently product when the C−O bond was broken, in line with the
disrupted during the LCD process. results of GC-MS and FTIR. The liquid product had more
2D-HSQC NMR was further employed to investigate the concentrated signals in the aromatic ring region compared to
structural changes of lignin and LP-3, and the obtained results lignin, suggesting that a portion of the aromatic ring structure
are exhibited in Figure 10 and Figure S7. According to a had disappeared, and the product was more concentrated as
previous report,51 the regions of [δC/δH 50−90/2.5−5.5] and the guaiacyl unit structure. There is also a disappearance of
4267 https://doi.org/10.1021/acs.energyfuels.3c04637
Energy Fuels 2024, 38, 4256−4272
Energy & Fuels pubs.acs.org/EF Article

Figure 11. Effect of reaction temperature (a) and reaction time (b) over the 3Co1Zr/SEP catalyst.

other signals in the region, indicating the excellent ability of the


prepared 3Co1Zr/SEP to break C−O bonds.
3.2.2. Effect of Reaction Time and Temperature on LCD.
Commonly, LCD is strongly influenced by the temperature
and the time of the reaction. The influence of reaction
temperatures on LCD over 3Co1Zr/SEP for 4 h was
investigated, and the results are presented in Figure 11a. The
temperature was raised from 230 to 290 °C. This brought
about the increase in the production of LP from 28.34 to
60.11%, while the yield of solid products dropped from 70.66
to 38.49%. It was suggested that the elevated temperature
facilitated the broken chemical bonds of the lignin structure
and the production of activated hydrogen in the ethanol
system, while the active intermediates generated from lignin
degradation further reacted with the activated hydrogen to
obtain the small-molecule monomer products. When the
temperature was raised to 310 °C, the LP yield decreased to
52.68% and the solid product yield increased to 42.12%. This Figure 12. PP conversion and product selectivity on the 3Co1Zr/SEP
is generally attributed to the excessive high temperature that catalyst as a function of time. Reaction conditions: 0.1 g of PP, 0.02 g
accelerated the polycondensation reaction of the intermediate of the catalyst, 30 mL of ethanol, 50 mg of n-dodecane as an internal
fragments. In addition, lignin-derived macromolecular frag- standard, 0.2 MPa N2 initial pressure, 240 °C, 450 rpm.
ments were also carbonized at excessive temperatures resulting
in the increase of the solid product. when the reaction lasted for 150 min, and the monomers that
The influence of reaction time on LCD over 3Co1Zr/SEP at converted from PP mainly contained phenol, phenethyl
290 °C was investigated, and the outcomes are presented in alcohol, and phenylacetaldehyde according to the results of
Figure 11b. It was found that the yields of the LP exhibited a GC-MS as presented in Figure S10. The other byproduct
clearly increased tendency with the reaction run from 1 to 12 benzaldehyde was generated by breakage of the Cα−Cβ bond
h, while the solid product yield presented the gradual or a condensation reaction. As the reaction time increased, the
reduction. The yield of the LP increased to 78.83%, and the monomer product of phenylacetaldehyde was further hydro-
solid product decreased to 20.23% after 12 h of reaction. In genated to form phenethyl alcohol; consequently, the
addition, the GC-MS results of the LP obtained after different selectivity of phenylacetaldehyde gradually reduced with a
reaction times are presented in Figure S9. This confirmed that concomitant increase of phenethyl alcohol selectivity. There-
the alkylation products significantly increased with reaction fore, according to the experimental outcomes and previous
times. These results demonstrated that the elongation of studies, the β-O-4 linkage in PP was broken in the following
reaction time could facilitate the lignin depolymerization steps: generally, the H+ derived from ethanol activation over
reaction, while the guaiacols could continue to undergo Co metallic sites would attack the Cα−OH to promote the
alkylation to produce alkylphenols. dehydration reaction to remove the hydroxyl groups of PP;
3.3. The Bond-Breaking Mechanism of β-O-4 and then, the reactive Cα* species would rearrange in the electron-
Kinetic Analysis. To investigate the bond-breaking ability of deficient environment to generate the (2-phenoxyvinyl)-
3Co1Zr/SEP on lignin, the model compound PP (2-phenoxy- benzene intermediate, which instantaneously hydrolyzed into
1-phenylethanol), which has the typical linkage bond (β-O-4) phenylacetaldehyde and phenol; then, phenylacetaldehyde
of lignin, was chosen for depolymerization experiments. As underwent the hydrogenation reaction to form phenethyl
shown in Figure 12, the conversion of PP reached 99.01% alcohol.62 Subsequently, the other pathway involved the
4268 https://doi.org/10.1021/acs.energyfuels.3c04637
Energy Fuels 2024, 38, 4256−4272
Energy & Fuels pubs.acs.org/EF Article

Figure 13. (a) SEP and (b) 3Co1Zr/SEP catalyst for PP conversion as a function of reaction temperature and reaction time. (c) Arrhenius plots for
SEP and the 3Co1Zr/SEP catalyst for PP conversion.

Scheme 2. Possible Reaction Pathway for LCD

breaking of the Cα−Cβ under the influence of LAS and Co the slopes varied with the increase of the temperature,
metal sites to produce benzaldehyde and anisole, and the indicating that there was a close relationship between the
anisole was hydrogenated to crack the C−O in methoxy to reaction rate and the temperature. Subsequently, the Arrhenius
produce phenol. formula was used to calculate the apparent activation energies
Furthermore, Figure 13 displays the impact of catalysts on (Ea) for PP conversions catalyzed by the SEP and 3Co1Zr/
the thermodynamic characteristics associated with the cleavage SEP. The Ea for PP conversion over SEP (89.94 kJ·mol−1) was
of the β-O-4 bond in PP. Figure 13a,b shows the well first- higher than that over 3Co1Zr/SEP (71.46 kJ·mol−1). It was
order linear relationship between the reaction time and ln(1 − indicated that the apparent activation energy for β-O-4 bond
conv.) of PP over SEP and 3Co1Zr/SEP. It was observed that breaking could be significantly reduced by using 3Co1Zr/SEP.
4269 https://doi.org/10.1021/acs.energyfuels.3c04637
Energy Fuels 2024, 38, 4256−4272
Energy & Fuels pubs.acs.org/EF Article

3.4. Proposing the Pathway of 3Co1Zr/SEP for LCD.


Generally, lignin was initially depolymerized into macro-
■ AUTHOR INFORMATION
Corresponding Authors
molecular fragments under the pyrolysis system, and then, the
breaking of the macromolecular fragments' C−C and C−O Mingqiang Chen − School of Chemical Engineering, Anhui
University of Science and Technology, 232001 Huainan, P.R.
bonds was realized under catalytic action to obtain highly
China; orcid.org/0000-0002-1607-8780;
reactive intermediates.29,32−35 Experimental and character-
Email: mqchen@aust.edu.cn, hnitchen@163.com
ization results showed that Co/SEP possessed the abundant Yishuang Wang − School of Chemical Engineering, Anhui
pores that facilitated the contact of lignin fragments with the University of Science and Technology, 232001 Huainan, P.R.
active sites. Additionally, the active Co species promoted the China; Email: yswang@aust.edu.cn, ys_wang1@163.com
breakage of lignin C−C and C−O bonds or their derived
macromolecular fragments and simultaneously activated Authors
ethanol to form H* and CH3CH2* species, subsequently Zinan Zhou − School of Chemical Engineering, Anhui
enhancing the hydrogenation and alkylation to obtain University of Science and Technology, 232001 Huainan, P.R.
guaiacols and alkylphenols. However, since the Co species in China
Co/SEP existed mainly as Co(OH)2, leading to the low Zheng Fang − School of Chemical Engineering, Anhui
number of LAS and catalyst active sites for breaking C−C and University of Science and Technology, 232001 Huainan, P.R.
C−O bonds, therefore, the guaiacol and alkylphenol yields China
were lower over Co/SEP. For 3Co1Zr/SEP, it not only had Defang Liang − School of Chemical Engineering, Anhui
the abundant pores, but the incorporation of Zr promoted the University of Science and Technology, 232001 Huainan, P.R.
transformation of metal hydroxide Co(OH)2 to Co3O4. China
Furthermore, it had been found that that the addition of Zr Longyang Li − School of Chemical Engineering, Anhui
facilitated the formation of a higher amount of LAS and VO, University of Science and Technology, 232001 Huainan, P.R.
which could effectually promote the breakage of lignin C−C China
and C−O bonds or their derived macromolecular fragments Haosheng Xin − School of Chemical Engineering, Anhui
and the activation of ethanol to further improve the University of Science and Technology, 232001 Huainan, P.R.
hydrogenation and alkylation reaction to obtain more guaiacols China
and alkylphenols during LCD. Consequently, under the Chang Li − School of Chemical Engineering, Anhui University
synergistic effect between the superior surface area and the of Science and Technology, 232001 Huainan, P.R. China;
highest LAS and abundant VO, 3Co1Zr/SEP exhibited the orcid.org/0000-0001-9857-1765
highest yields of guaiacols and alkylphenols. Therefore, Gang Yuan − School of Chemical Engineering, Anhui
University of Science and Technology, 232001 Huainan, P.R.
combining the results of the above analysis and discussion,
China
the reaction pathway for the LCD catalyzed by 3Co1Zr/SEP is
Jun Wang − School of Chemical Engineering, Anhui University
proposed in Scheme 2. of Science and Technology, 232001 Huainan, P.R. China
Complete contact information is available at:
4. CONCLUSIONS
https://pubs.acs.org/10.1021/acs.energyfuels.3c04637
The influence of SEP-supported Co−Zr catalysts with varying
Co/Zr ratios on LCD was examined in this paper. Character- Notes
ization results demonstrated that the addition of Co brought The authors declare no competing financial interest.
an abundant pore structure to the catalyst, which facilitated the M.C. performed supervision, resources and funding acquis-
contact between the active site and lignin fragments. Moreover, ition, project administration, and review and editing of the
the addition of Zr species contributed to the conversion of manuscript. Z.Z. performed investigation, data curation,
Co(OH)2 in Co−Zr/SEP and increased LAS and VO. writing of the original draft, visualization, validation, and
3Co1Zr/SEP possessed the well pore structure and plenty of editing of the manuscript. Y.W. performed supervision,
LAS and VO and then exhibited the excellent LCD perform- resources and funding acquisition, visualization, project
ance, in which the optimal yields of guaiacols (445.2 mg/g administration, and review and editing of the manuscript.
lignin) and alkylphenols (109.8 mg/g lignin) were achieved Z.F. performed visualization and software analysis. D.L.
under the optimal reaction conditions. In addition, it was performed data curation and visualization. L.L. performed
demonstrated that 3Co1Zr/SEP significantly reduced the software analysis. H.X. performed data curation and editing of
apparent activation energy for breaking β-O-4 bonds to the manuscript. C.L. performed review and editing of the
71.46 from 89.84 kJ·mol−1 over SEP. manuscript. G.Y. performed formal analysis. J.W. performed
visualization.
■ ASSOCIATED CONTENT
* Supporting Information
sı ■ ACKNOWLEDGMENTS
The Supporting Information is available free of charge at The authors are thankful for financial support from the funds
https://pubs.acs.org/doi/10.1021/acs.energyfuels.3c04637. of the National Natural Science Foundation of China
(22378003, 51906001, and 51876001), the Natural Science
Additional characterization results for catalysts, detailed Research Project of Universities in Anhui Province
information on GC-MS results for liquid product, and (KJ2020ZD31), and the Scientific Research Foundation for
recyclability experiment details and characterization data High-Level Talents of Anhui University of Science and
(PDF) Technology. In addition, we are also thankful for the support
4270 https://doi.org/10.1021/acs.energyfuels.3c04637
Energy Fuels 2024, 38, 4256−4272
Energy & Fuels pubs.acs.org/EF Article

of the characterization analysis technology given by the test (19) Dou, X.; Li, W.; Zhu, C. Catalytic hydrotreatment of Kraft
center of Anhui University of Science and Technology. lignin into liquid fuels over porous ZnCoOx nanoplates. Fuel 2021,
283, No. 118801.

■ REFERENCES
(1) Schutyser, W.; Renders, T.; Van den Bosch, S.; Koelewijn, S. F.;
(20) Zhu, L.; Li, W.; Zhang, H.; Zhang, X.; Jin, J.; Wu, M. Bimetallic
ruthenium- and zinc-doped beta zeolite for efficiently depolymerizing
Kraft lignin. Fuel 2023, 349, No. 128766.
Beckham, G. T.; Sels, B. F. Chemicals from lignin: an interplay of (21) Shu, R.; Xu, Y.; Ma, L.; Zhang, Q.; Wang, C.; Chen, Y.
lignocellulose fractionation, depolymerisation, and upgrading. Chem. Controllable production of guaiacols and phenols from lignin
Soc. Rev. 2018, 47, 852−908. depolymerization using Pd/C catalyst cooperated with metal chloride.
(2) Zhou, Z.; Liu, D.; Zhao, X. Conversion of lignocellulose to Chemical Engineering Journal 2018, 338, 457−464.
biofuels and chemicals via sugar platform: An updated review on (22) Ren, T.; You, S.; Zhang, M.; Wang, Y.; Qi, W.; Su, R.; He, Z.
chemistry and mechanisms of acid hydrolysis of lignocellulose. Improved conversion efficiency of Lignin-to-Fuel conversion by
Renewable and Sustainable Energy Reviews 2021, 146, No. 111169. limiting catalyst deactivation. Chem. Eng. J. 2021, 410, No. 128170.
(3) Zhang, C.; Shen, X.; Jin, Y.; Cheng, J.; Cai, C.; Wang, F. (23) Raikwar, D.; Van Aelst, K.; Vangeel, T.; Corderi, S.; Van Aelst,
Catalytic Strategies and Mechanism Analysis Orbiting the Center of J.; Van den Bosch, S.; Servaes, K.; Vanbroekhoven, K.; Elst, K.; Sels,
Critical Intermediates in Lignin Depolymerization. Chem. Rev. 2023, B. F. Elucidating the effect of the physicochemical properties of
123, 4510−4601. organosolv lignins on its solubility and reductive catalytic depolyme-
(4) Gao, Y.; Ma, H.; Rao, Y.; Lv, K.; Shu, F.; Long, J. Selective rization. Chemical Engineering Journal 2023, 461, No. 141999.
hydrogenolysis of lignin in the presence of Ni3Fe1 alloy supported on (24) Wu, Y.; Dang, Q.; Wu, T.; Lei, T.; Wang, K.; Luo, Z. Efficient
zirconium phosphate. Chem. Eng. Sci. 2023, 271, No. 118570. Lignin Depolymerization Process for Phenolic Products with Lignin-
(5) Liu, Q.; Sang, Y.; Bai, Y.; Wu, K.; Ma, Z.; Chen, M.; Ma, Y.; Based Catalysts and Mixed Solvents. Energy Fuels 2023, 37, 5206−
Chen, H.; Li, Y. Catalytic conversion of Kraft lignin into platform 5219.
chemicals in supercritical ethanol over a Mo(OCH2CH3)x/NaCl (25) Limarta, S. O.; Kim, H.; Ha, J.-M.; Park, Y.-K.; Jae, J. High-
catalyst. Catal. Today 2023, 408, 204−210. quality and phenolic monomer-rich bio-oil production from lignin in
(6) Lu, X.; Gu, X. Efficient lignin conversion over Ni/(Fe/Zn/Co/ supercritical ethanol over synergistic Ru and Mg-Zr-oxide catalysts.
Mo/Cu)−WO3/Al2O3 for selectively yielding alkyl phenols. Catalysis Chemical Engineering Journal 2020, 396, No. 125175.
Science & Technology 2023, 13, 468−478. (26) Huang, X.; Korányi, T. I.; Boot, M. D.; Hensen, E. J. M.
(7) Petridis, L.; Smith, J. C. Molecular-level driving forces in Catalytic Depolymerization of Lignin in Supercritical Ethanol.
lignocellulosic biomass deconstruction for bioenergy. Nature Reviews ChemSusChem 2014, 7, 2276−2288.
Chemistry 2018, 2, 382−389. (27) Zhou, M.; Sharma, B. K.; Liu, P.; Ye, J.; Xu, J.; Jiang, J.-C.
(8) Subbotina, E.; Rukkijakan, T.; Marquez-Medina, M. D.; Yu, X.; Catalytic in Situ Hydrogenolysis of Lignin in Supercritical Ethanol:
Johnsson, M.; Samec, J. S. M. Oxidative cleavage of C-C bonds in Effect of Phenol, Catalysts, and Reaction Temperature. ACS
lignin. Nat. Chem. 2021, 13, 1118−1125. Sustainable Chem. Eng. 2018, 6, 6867−6875.
(9) Zhang, C.; Wang, F. Catalytic Lignin Depolymerization to (28) Zhang, M.; Wu, Y.; Han, X.; Zeng, Y.; Xu, C. C. Upgrading
Aromatic Chemicals. Acc. Chem. Res. 2020, 53, 470−484. pyrolysis oil by catalytic hydrodeoxygenation reaction in supercritical
(10) Prashanth, P. F.; Gurrala, L.; Mohan, R. V.; Sarvanakumar, K.;
ethanol with different hydrogen sources. Chemical Engineering Journal
Vinu, R. Microwave-assisted torrefaction and pyrolysis of rice straw
2022, 446, No. 136952.
pellets for bioenergy. IET Renewable Power Gener. 2022, 16, 2964−
(29) Huang, X.; Atay, C.; Korányi, T. I.; Boot, M. D.; Hensen, E. J.
2977.
M. Role of Cu-Mg-Al Mixed Oxide Catalysts in Lignin Depolymeriza-
(11) Anderson, E. M.; Stone, M. L.; Katahira, R.; Reed, M.;
tion in Supercritical Ethanol. ACS Catal. 2015, 5, 7359−7370.
Muchero, W.; Ramirez, K. J.; Beckham, G. T.; Román-Leshkov, Y.
(30) Du, B.; Liu, C.; Wang, X.; Han, Y.; Guo, Y.; Li, H.; Zhou, J.
Differences in S/G ratio in natural poplar variants do not predict
Renewable lignin-based carbon nanofiber as Ni catalyst support for
catalytic depolymerization monomer yields. Nat. Commun. 2019, 10,
2033. depolymerization of lignin to phenols in supercritical ethanol/water.
(12) Achour, A.; Bernin, D.; Creaser, D.; Olsson, L. Evaluation of Renewable Energy 2020, 147, 1331−1339.
kraft and hydrolysis lignin hydroconversion over unsupported NiMoS (31) Tran, M. H.; Phan, D.-P.; Nguyen, T. H.; Kim, H. B.; Kim, J.;
catalyst. Chemical Engineering Journal 2023, 453, No. 139829. Park, E. D.; Lee, E. Y. Catalytic hydrogenolysis of alkali lignin in
(13) Long, J.; Lou, W.; Wang, L.; Yin, B.; Li, X. [C4H8SO3Hmim]- supercritical ethanol over copper monometallic catalyst supported on
HSO4 as an efficient catalyst for direct liquefaction of bagasse lignin: a chromium-based metal-organic framework for the efficient
Decomposition properties of the inner structural units. Chem. Eng. Sci. production of aromatic monomers. Bioresour. Technol. 2021, 342,
2015, 122, 24−33. No. 125941.
(14) Chen, J.; Wang, D.; Lu, X.; Guo, H.; Xiu, P.; Qin, Y.; Xu, C.; (32) Yan, F.; Ma, R.; Ma, X.; Cui, K.; Wu, K.; Chen, M.; Li, Y.
Gu, X. Effect of Cobalt(II) on Acid-Modified Attapulgite-Supported Ethanolysis of Kraft lignin to platform chemicals on a MoC1‑x/Cu-
Catalysts on the Depolymerization of Alkali Lignin. Ind. Eng. Chem. MgAlOz catalyst. Applied Catalysis B: Environmental 2017, 202, 305−
Res. 2022, 61, 1675−1683. 313.
(15) Karnitski, A.; Choi, J. W.; Suh, D. J.; Yoo, C. J.; Lee, H.; Kim, (33) Chen, M.; Dai, W.; Wang, Y.; Tang, Z.; Li, H.; Li, C.; Yang, Z.;
K. H.; Kim, C. S.; Kim, K.; Ha, J. M. Roles of metal and acid sites in Wang, J. Selective catalytic depolymerization of lignin to guaiacols
the reductive depolymerization of concentrated lignin over supported over Mo-Mn/sepiolite in supercritical ethanol. Fuel 2023, 333,
Pd catalysts. Catal. Today 2023, 411−412. No. 126365.
(16) Li, Y.; Cai, Z.; Liao, M.; Long, J.; Zhao, W.; Chen, Y.; Li, X. (34) Wang, Y.; Tang, Z.; Chen, M.; Zhang, J.; Shi, J.; Wang, C.;
Catalytic depolymerization of organosolv sugarcane bagasse lignin in Yang, Z.; Wang, J. Effect of Mo content in Mo/Sepiolite catalyst on
cooperative ionic liquid pairs. Catal. Today 2017, 298, 168−174. catalytic depolymerization of Kraft lignin under supercritical ethanol.
(17) Dou, X.; Li, W.; Zhu, C.; Jiang, X. Catalytic waste Kraft lignin Energy Conversion and Management 2020, 222, No. 113227.
hydrodeoxygenation to liquid fuels over a hollow Ni-Fe catalyst. (35) Chen, M.; Shi, J.; Wang, Y.; Tang, Z.; Yang, Z.; Wang, J.;
Applied Catalysis B: Environmental 2021, 287, No. 119975. Zhang, H. Conversion of Kraft lignin to phenol monomers and liquid
(18) Dou, X.; Jiang, X.; Li, W.; Zhu, C.; Liu, Q.; Lu, Q.; Zheng, X.; fuel over trimetallic catalyst W-Ni-Mo/sepiolite under supercritical
Chang, H.-m.; Jameel, H. Highly efficient conversion of Kraft lignin ethanol. Fuel 2021, 303, No. 121332.
into liquid fuels with a Co-Zn-beta zeolite catalyst. Applied Catalysis (36) Ji, J.; Guo, H.; Li, C.; Qi, Z.; Zhang, B.; Dai, T.; Jiang, M.; Ren,
B: Environmental 2020, 268, No. 118429. C.; Wang, A.; Zhang, T. Tungsten-Based Bimetallic Catalysts for

4271 https://doi.org/10.1021/acs.energyfuels.3c04637
Energy Fuels 2024, 38, 4256−4272
Energy & Fuels pubs.acs.org/EF Article

Selective Cleavage of Lignin C-O Bonds. ChemCatChem. 2018, 10, Co(OH)2 and Co3O4 to ward Boosting Electrochemical Charge
415−421. Storage. Adv. Funct. Mater. 2021, 32, 2108644.
(37) Zhao, M.; Deng, J.; Liu, J.; Li, Y.; Liu, J.; Duan, Z.; Xiong, J.; (53) Huang, N.; Qu, Z.; Dong, C.; Qin, Y.; Duan, X. Superior
Zhao, Z.; Wei, Y.; Song, W.; Sun, Y. Roles of Surface-Active Oxygen performance of α@β-MnO2 for the toluene oxidation: Active
Species on 3DOM Cobalt-Based Spinel Catalysts MxCo3‑xO4 (M = Zn interface and oxygen vacancy. Applied Catalysis A: General 2018,
and Ni) for NOx-Assisted Soot Oxidation. ACS Catal. 2019, 9, 7548− 560, 195−205.
7567. (54) Wang, J.; Wang, W.; Wang, J.; Xue, K.; Peng, Y.; Yan, Y.; Wang,
(38) Rautiainen, S.; Di Francesco, D.; Katea, S. N.; Westin, G.; Y.; Wang, H.; Wu, Y. The generation of lattice oxygen defects
Tungasmita, D. N.; Samec, J. S. M. Lignin Valorization by Cobalt- enhanced by β particles: Layered microsphere-like Bi2WO6 as a
Catalyzed Fractionation of Lignocellulose to Yield Monophenolic template leads to Bix@Bi2‑xWOn for the efficient removal of
Compounds. ChemSusChem 2019, 12, 404−408. oxytetracycline. Chemical Engineering Journal 2021, 416, No. 129197.
(39) Zhang, C.; Zhang, X.; Wu, J.; Zhu, L.; Wang, S. Hydro- (55) Wang, C.; Wang, Y.; Chen, M.; Liang, D.; Cheng, W.; Li, C.;
deoxygenation of lignin-derived phenolics to cycloalkanes over Ni-Co Yang, Z.; Wang, J. Hydrogen production from tar steam reforming
alloy coupled with oxophilic NbO. Applied Energy 2022, 328, over hydrangea-like Co-phyllosilicate catalyst derived from Co/
No. 120199. Sepiolite. Int. J. Hydrogen Energy 2023, 48, 2542−2557.
(40) Xiang, S.; Dong, L.; Wang, Z.; Han, X.; Guo, Y.; Liu, X.; Gong, (56) Cui, G.; Zhang, X.; Wang, H.; Li, Z.; Wang, W.; Yu, Q.; Zheng,
X.-Q.; Wang, Y. Co@CoO: An efficient catalyst for the depolymeriza- L.; Wang, Y.; Zhu, J.; Wei, M. ZrO2‑x modified Cu nanocatalysts with
tion and upgrading of lignocellulose to alkylcyclohexanols with synergistic catalysis towards carbon-oxygen bond hydrogenation.
cellulose intact. Journal of Energy Chemistry 2023, 77, 191−199. Applied Catalysis B: Environmental 2021, 280, No. 119406.
(41) Qi, Y.; Xiao, X.; Mei, Y.; Xiong, L.; Chen, L.; Lin, X.; Lin, Z.; (57) Xiao, Z.; Wang, Y.; Huang, Y.-C.; Wei, Z.; Dong, C.-L.; Ma, J.;
Shen, S.; Li, Y.; Wang, S. Filling the oxygen vacancies in Co3O4 with
Sun, S.; Han, B.; Yang, D.; Qin, Y.; Qiu, X. Modulation of Bro̷ nsted
phosphorus: an ultra-efficient electrocatalyst for overall water
and Lewis Acid Centers for NixCo3‑xO4 Spinel Catalysts: Towards
splitting. Energy Environ. Sci. 2017, 10, 2563−2569.
Efficient Catalytic Conversion of Lignin. Adv. Funct. Mater. 2022, 32,
(58) Ji, P.; Feng, X.; Veroneau, S. S.; Song, Y.; Lin, W. Trivalent
2111615. Zirconium and Hafnium Metal-Organic Frameworks for Catalytic 1,4-
(42) Hongmanorom, P.; Ashok, J.; Das, S.; Dewangan, N.; Bian, Z.; Dearomative Additions of Pyridines and Quinolines. J. Am. Chem. Soc.
Mitchell, G.; Xi, S.; Borgna, A.; Kawi, S. Zr-Ce-incorporated Ni/SBA- 2017, 139, 15600−15603.
15 catalyst for high-temperature water gas shift reaction: Methane (59) Ranaware, V.; Kurniawan, R. G.; Verma, D.; Kwak, S. K.; Ryu,
suppression by incorporated Zr and Ce. J. Catal. 2020, 387, 47−61. B. C.; Kang, J. W.; Kim, J. Solvent-mediated selectivity control of
(43) Xiao, T.; Liu, X.; Xu, G.; Zhang, Y. Phase tuning of ZrO2 furfural hydrogenation over a N-doped carbon-nanotube-supported
supported cobalt catalysts for hydrodeoxygenation of 5-hydroxyme- Co/CoOx catalyst. Applied Catalysis B: Environmental 2022, 318,
thylfurfural to 2,5-dimethylfuran under mild conditions. Applied No. 121838.
Catalysis B: Environmental 2021, 295, No. 120270. (60) Luo, Z.; Kong, J.; Ma, B.; Wang, Z.; Huang, J.; Zhao, C.
(44) Dostagir, N. H. M. D.; Rattanawan, R.; Gao, M.; Ota, J.; Liquefaction and Hydrodeoxygenation of Polymeric Lignin Using a
Hasegawa, J.-y.; Asakura, K.; Fukouka, A.; Shrotri, A. Co Single Atoms Hierarchical Ni Microreactor Catalyst. ACS Sustainable Chem. Eng.
in ZrO2 with Inherent Oxygen Vacancies for Selective Hydrogenation 2020, 8, 2158−2166.
of CO2 to CO. ACS Catal. 2021, 11, 9450−9461. (61) Davidson, S. D.; Sun, J.; Hong, Y.; Karim, A. M.; Datye, A. K.;
(45) Zhou, Y.; Liu, L.; Li, G.; Hu, C. Insights into the Influence of Wang, Y. The effect of ZnO addition on Co/C catalyst for vapor and
ZrO2 Crystal Structures on Methyl Laurate Hydrogenation over Co/ aqueous phase reforming of ethanol. Catal. Today 2014, 233, 38−45.
ZrO2 Catalysts. ACS Catal. 2021, 11, 7099−7113. (62) Mei, X.; Liu, H.; Wu, H.; Wu, W.; Zheng, B.; Zhang, K.; Xu, C.;
(46) García, A.; Sánchez-Tovar, R.; Miguel, P. J.; Montejano-Nares, Xu, J.; He, M.; Han, B. Efficient Cleavage of Lignin Model
E.; Ivars-Barceló, F.; Cecilia, J. A.; Torres-Olea, B.; Solsona, B. Compounds into Phenols and Aldehydes over NiOOH Catalyst.
Catalytic production of γ-valerolactone, a biofuel precursor, from ChemCatChem. 2022, 14, No. e202200554.
furfural in one-pot: Synergistic effect between Zr and Sn. Fuel 2023,
352, No. 129045.
(47) Wang, Y.; Wang, C.; Chen, M.; Hu, J.; Tang, Z.; Liang, D.;
Cheng, W.; Yang, Z.; Wang, J.; Zhang, H. Influence of CoAl2O4
spinel and Co-phyllosilicate structures derived from Co/sepiolite
catalysts on steam reforming of bio-oil for hydrogen production. Fuel
2020, 279, No. 118449.
(48) Huang, Z.; Zhang, J.; Li, S.; Yuan, G.; Li, F.; Zeng, Y.; Han, L.;
Jia, Q.; Zhang, H.; Zhang, S. Joule-heatable bird-nest-bioinspired/
carbon nanotubes-modified sepiolite porous ceramics: An efficient,
sturdy, and continuous strategy for oil recovery. Journal of Hazardous
Materials 2021, 417, No. 125979.
(49) Dong, N.; Ye, Q.; Chen, M.; Cheng, S.; Kang, T.; Dai, H.
Sodium-treated sepiolite-supported transition metal (Cu, Fe, Ni, Mn,
or Co) catalysts for HCHO oxidation. Chinese Journal of Catalysis
2020, 41, 1734−1744.
(50) Chen, M.; Tang, Z.; Wang, Y.; Shi, J.; Li, C.; Yang, Z.; Wang, J.
Catalytic depolymerization of Kraft lignin to liquid fuels and guaiacol
over phosphorus modified Mo/Sepiolite catalyst. Chemical Engineering
Journal 2022, 427, No. 131761.
(51) Chen, M.; Li, H.; Wang, Y.; Tang, Z.; Dai, W.; Li, C.; Yang, Z.;
Wang, J. Lignin depolymerization for aromatic compounds over Ni-
Ce/biochar catalyst under aqueous-phase glycerol. Applied Energy
2023, 332, No. 120489.
(52) Gao, P.; Zeng, Y.; Tang, P.; Wang, Z.; Yang, J.; Hu, A.; Liu, J.
Understanding the Synergistic Effects and Structural Evolution of

4272 https://doi.org/10.1021/acs.energyfuels.3c04637
Energy Fuels 2024, 38, 4256−4272

You might also like