You are on page 1of 40

Catalytic transfer hydrogenolysis of switchgrass

lignin with ethanol using spinel-type mixed-metal

oxide catalysts affords control of the oxidation state

of isolated aromatic products

James A. Godwin,1 Jonah P. Babusci,1 Nichole M. Wonderling,2 Jeffrey R. Shallenberger,2

Kendhl Seabright,3 David P. Harper,3 Stephen C. Chmely1*

1
Department of Agricultural and Biological Engineering, The Pennsylvania State University,

University Park, PA 16802 USA.

2
Materials Research Institute, The Pennsylvania State University, University Park, PA 16802

USA.

3
Center for Renewable Carbon, University of Tennessee Institute of Agriculture, Knoxville, TN

37996 USA.

KEYWORDS: lignin, biorefining, catalytic transfer hydrogenation, catalytic transfer

hydrogenolysis, spinel, metal oxide.

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
1
Abstract:

Chemical reductions of lignin are useful to remove oxygen and create product slates that can

function as renewable platform molecules for new fuels and chemicals. Catalytic transfer

hydrogenolysis (CTH) is an underexplored method to reduce lignin that obviates the use of

dangerous and non-renewable hydrogen gas. While noble metals are used extensively as catalysts

for transfer hydrogenation, sustainability remains a major challenge to their deployment. In this

work, we synthesized mixed-metal oxides of earth-abundant Co and Ni and characterized the

catalysts using powder x-ray diffraction (XRD). Catalyst reactivity for CTH of acetophenone was

also assessed. Among the catalysts tested, spinel NiCo2O4 demonstrated the highest conversion of

acetophenone (75%) and the highest selectivity for ethylbenzene (90%); thus, we applied it to

valorize switchgrass lignin, extracted under mild operating conditions by cosolvent enhanced

lignocellulosic fractionation (CELF). The catalytically depolymerized lignin showed an increase

in selectively deoxygenated monomeric compounds. As demonstrated using 2D-NMR

spectroscopy, the lignin displayed highly reduced aliphatic carbons resulting from the catalyst-

mediated reduction reaction at the Cα sites. These results are critical to the further development of

the lignin-first biorefinery, as they demonstrate the use of sustainable catalyst materials and mild

transformation conditions to generate and refine a suite of new bioproducts.

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
2
INTRODUCTION

Lignocellulosic biomass has garnered attention as a renewable carbon feedstock because of its

abundance and potential to replace petroleum as a source of fuels and chemicals. Lignin is a multi-

alkyl-aromatic heteropolymer and a primary component of lignocellulosic biomass.1 It holds great

potential to be directly incorporated into a variety of consumer products or otherwise chemically

processed in a manner analogous to petrochemical refining.2 Traditional biomass processing

technologies, including pulping for papermaking and newer cellulosic ethanol refining, focus on

isolating pure cellulose and maximizing pulp yield by removing lignin, which is considered a

source of biomass recalcitrance.3,4 Technologies used to isolate cellulose, including Kraft pulping,

acidic pretreatments, and even early organosolv fractionation schemes, typically sacrifice lignin

structural quality for high cellulose yield.

However, because of the unique structural features of lignin, including its abundance of aromatic

groups and chemically reactive handles (e.g., hydroxyl and carboxylate groups), new fractionation

technologies have focused on depolymerizing and stabilizing lignin to ensure high purity and a

product slate with limited structural heterogeneity. The resultant molecular product platform

would incur minimal separation costs and maximize utility as a source of new bio-privileged

molecules.5 This type of lignin-first biorefining6 is alternatively referred to as catalytic upstream

biorefining,7,8 early-stage catalytic conversion of lignin,9–11 or reductive catalytic fractionation

(RCF),12 each with subtle but important differences in reaction conditions. Ultimately, these

technologies are designed to isolate high-yield and high-purity lignin for conversion into new

industrially relevant products.13,14

Hydrogenation and hydrodeoxygenation reactions are a central focus of lignin valorization

schemes because the catalytic cleavage of ether bonds in lignin necessarily requires hydrogen.

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
3
However, hydrogen is highly flammable and challenging to use at scale, primarily because of its

limited solubility in solvents relevant to biomass conversion,15which necessitates high operating

pressures. A recent technoeconomic analysis of a biorefinery using RCF showed that high-pressure

reactors required for traditional hydrogenations account for nearly 60% of the installed capital

costs of the biorefinery.16 Moreover, despite the significant research efforts dedicated to producing

renewable hydrogen,17 steam reforming of non-renewable natural gas continues to supply 95% of

global hydrogen demand,18 offsetting the potential sustainability gains of biorefining processes.16

Transfer hydrogenation is emerging as a more sustainable alternative to deoxygenate and

depolymerize lignin.19–23 Formic acid, alcohols, alkanes, ammonia borane, hydrazine, and even

biomass fractions can serve as hydrogen donors in methods to achieve transfer hydrogenation and

circumvent the use of molecular hydrogen.19 Alcohol solvents are, perhaps, the most intriguing

hydrogen donors; they are readily synthesized from the polysaccharide fractions of biomass and

can function simultaneously as the reducing agent and reaction solvent. Several existing examples

of catalytic transfer hydrogenation/hydrogenolysis (CTH) of lignin employ ethanol or 2-propanol

with a variety of catalysts, including those based on Ni,24–28 Cu,29–31 Ru,32–37 Cs,38 bimetallic

species,39–43 and noble metals.44–47 These and other catalysts and systems for lignin transfer

hydrogenation and hydrogenolysis have been thoroughly reviewed elsewhere. 22,48–52

Spinel-type mixed-metal oxide catalysts are a particularly exciting class of catalyst materials for

CTH. Their chemical reactivity can be tuned by altering the ratios of metals in the material, and

they are air-stable metal oxides that do not require a reducing environment for activation.

Furthermore, mixtures containing iron, cobalt, and nickel are ferromagnetic, meaning the catalysts

are readily magnetically separable from batch reaction mixtures.53 Although these materials have

been deployed for lab-scale catalytic hydrogenation of furan-based molecules (derivable from

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
4
biomass polysaccharides),21 their utility for CTH of biomass-derived lignin remains largely

unexplored, despite their apparent usefulness for such a process.

Controlling the oxidation state of the alpha-C in lignin (Scheme 1) can bestow industrially-

valuable material properties. For example, maximally oxidized Cα (i.e., a ketone) yields highly

conjugated systems capable of absorbing UV light, lending light-blocking capabilities.54 On the

other hand, a Cα hydroxyl group provides a chemical platform to access new small molecules55 or

polymer compounds56 through judicious chemical modification. The Cα oxidation state is sensitive

to lignin isolation procedures, whereby oxidizing acid (e.g., H2SO4) can oxidize the naturally

occurring alpha-hydroxy group to a ketone. Therefore, exerting control over the Cα oxidation state

can be used to explore the chemical and photo-properties of lignin, potentially introducing a

scheme to enhance lignin’s industrial utility.

Scheme 1. Labeling scheme and potential industrial applications of lignin.

Accordingly, we evaluated spinel-type mixed-metal oxides of cobalt and nickel as lignin CTH

catalysts and present the findings here. Using acetophenone, we demonstrated differences in

reactivity and selectivity at the meso (benzylic) carbon based on the atomic composition of the as-

tested catalysts. As the mixed Co-Ni species showed enhanced ability to reduce aliphatic carbons,

this catalyst was applied to switchgrass lignin isolated via the co-solvent enhanced lignocellulosic

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
5
fractionation (CELF) process.57–59 Our findings from gel permeation chromatography (GPC), gas
1
chromatography-mass spectrometry (GC/MS), and H-13C heteronuclear single-quantum

coherence (HSQC) nuclear magnetic resonance (NMR) spectroscopy indicate the catalyst species

can selectively deoxygenate biomass-derived lignin using ethanol via CTH. We comment on the

relative sustainability of spinel-type mixed-metal oxide catalysts and state-of-the-art supported

metal catalysts and consider these catalysts’ role in a lignocellulosic biorefinery that employs

reductive catalytic fractionation.

EXPERIMENTAL SECTION

Biomass and other reagents

Alamo switchgrass (Panicum virgatum) was obtained from Ernst Conservation Seeds. The

plants were harvested after senescence from a plot in Meadville, northwestern PA, USA. The

biomass material was size reduced to 1 mm using a knife mill (Model 4 Wiley mill, Arthur H.

Thomas Company, Philadelphia, PA USA). All other chemical reagents used were of analytical

grade with at least 95% purity unless otherwise specified, and they were used as received.

Quantification of Lignin in Switchgrass

The lignin content of the switchgrass was determined using the acetyl bromide soluble lignin

(ABSL) assay protocol as described elsewhere.60 In brief, 5 g of milled switchgrass was used for

the lignin quantification, as described in the protocol, and the percentage of ABSL was then

computed using the absorbance and extinction coefficient of switchgrass reported from previous

studies.61

Biomass fractionation and isolation of lignin

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
6
Lignin extraction from switchgrass was achieved using cosolvent enhanced lignocellulosic

fractionation (CELF process) as described previously.57–59 Briefly, a 1.0-L Parr reactor with

Inconel wetted parts was charged with 360 mL deionized water containing 3.6 mL concentrated

H2SO4 (98%, 18.4 M), 360 mL tetrahydrofuran (THF), and 80 g switchgrass. An aluminum block

heater was pre-heated above the desired reaction temperature and subsequently applied to the

reactor to quickly heat it to the desired temperature, which was controlled externally by

simultaneously monitoring the temperature of the reactor skin and the internal reactor contents.

This allowed for a rapid heat-up time of the room-temperature reactor contents to reaction

temperature in 20 min. The CELF process was conducted at 150 °C for 20 min (measured after the

reactor reached the final temperature) with constant stirring at 200 rpm. The reactor was

subsequently cooled in an ice bath and the contents were collected and filtered through a glass

fiber filter with a pore size of 0.5 micron. The lignin-rich liquid phase was neutralized with

ammonium hydroxide and concentrated with a rotary evaporator to precipitate the dissolved lignin.

The remaining liquor was decanted, and the lignin precipitate was washed with deionized water

and diethyl ether and dried under vacuum at 40 °C for 20 h.

Catalyst preparation

A urea-assisted synthesis procedure was adapted to synthesize various Co and Ni metal oxides

using CoCl2•6H2O and NiCl2•6H2O as precursors.62 For mono-metallic oxides, a

poly(tetrafluoroethylene)-lined (PTFE) hydrothermal autoclave (Parr Model 4744A, Parr

Instrument Company, Moline, IL USA) was charged with a 36-mL aqueous solution containing

25 g of the desired metal chloride hydrate and 11 g urea (7:3 total metal salt : urea by weight) and

heated to 180 °C for 16 h. After the reaction was complete, the resulting product was removed

from the reactor and washed with deionized water and ethanol and dried under vacuum at 40 °C

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
7
for 16 h. Before catalytic reactions, this powder was calcined for 2 h at 500 °C under flowing air

to produce the metal oxide. The procedure for mixed-metal oxides was identical except the ratios

of trivalent to divalent metals (e.g., Co:Ni) was 2:1.

Catalyst characterization

X-ray diffraction patterns were collected at 40 kV and 40 mA on a 240 mm

radius Panalytical Empyrean® theta-theta X-ray diffractometer equipped with a line source [Co

K-α 1-2 (1.789010/1.792900 Å)] X-ray tube. The instrument is equipped with a flat programmable

Z-stage and used for phase identification. Data was collected with a step size of 0.0167° from 10-

85° 2-theta. The incident optics consisted of a Bragg-Brentano HD® Co optic fitted with 0.04

rad Soller slits, a 10 mm beam mask, 1/8° and 1/2° divergence, and anti-scatter slit, respectively.

The diffracted optics included a X’Celerator® detector with a 2.1223 active length in scanning

line mode with a 1/4° programmable anti-scatter slit and 0.04 rad Soller slits. Phase identification

was carried out using Jade software (version 8.5) from Materials Data Inc. (MDI) and the

International Centre for Diffraction Data (ICDD) PDF4 database.

Major elemental concentrations were determined using a Cameca SXFive Electron Probe Micro-

Analyzer (EPMA). The instrument is equipped with five wavelength dispersive spectrometers and

a LaB6 electron source. An accelerating voltage of 20 keV and a beam current of 20 nA were used

with a defocused 3 µm beam. The produced X-ray intensities were subject to a PAP (phi-rho-z)

matrix correction algorithm as described by Pouchou and Pichoir (1987) and reported by Christien

et al.,63 and converted to concentrations by comparison to metal oxide standards. The concentration

data for the major elements were collected on pressed power samples. Multiple point analyses (5

for CoO and 10 for CoNiO spinel) were averaged with associated errors to obtain the material

compositions.

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
8
X-ray photoelectron spectroscopy (XPS) experiments were performed using a Physical

Electronics VersaProbe III instrument equipped with a monochromatic Al kα x-ray source (hν =

1,486.6 eV) and a concentric hemispherical analyzer. Charge neutralization was performed using

both low energy electrons (<5 eV) and argon ions. The binding energy axis was calibrated using

sputter cleaned Cu (Cu 2p3/2 = 932.62 eV, Cu 3p3/2 = 75.1 eV) and Au foils (Au 4f7/2 = 83.96 eV).

Peaks were charge referenced to CHx band in the carbon 1s spectra at 284.8 eV. Measurements

were made at a takeoff angle of 45° with respect to the sample surface plane. This resulted in a

typical sampling depth of 3-6 nm (95% of the signal originated from this depth or shallower).

Quantification was done using instrumental relative sensitivity factors (RSFs) that account for the

x-ray cross section and inelastic mean free path of the electrons. On homogeneous samples major

elements (>5 atom%) tend to have standard deviations of <3% while minor elements can be

significantly higher. The analysis size was ~200µm in diameter.

Catalytic transfer hydrogenation/hydrogenolysis (CTH) of acetophenone

A stock solution containing 0.1 g acetophenone, 0.1 g durene, and 9.8 g anhydrous ethanol was

prepared (10 g total mass). Then, a stainless steel microreactor consisting of a 0.5-in MNPT

stainless steel pipe fitting (Swagelok, P/N: SS-8-HN) and a pair of 0.5-in FNPT stainless steel pipe

caps (Swagelok, P/N: SS-8-CP) was charged with 0.01 g catalyst of interest and 1.0 mL of the

stock solution. The resulting suspension was diluted with 6 mL anhydrous ethanol, and the reactor

was sealed with Teflon tape. The entire reactor was submerged in a molten salt bath (50:50

NaNO2:KNO3) and heated to 320 °C for 2 h. The reactor was subsequently quenched in an ice-

water bath and allowed to equilibrate over 10 min, after which time it was carefully opened to

remove an aliquot of the reaction mixture for analysis by GC/MS. Caution! Acetaldehyde, the

product of dehydrogenation of ethanol, has a boiling point of 20 °C. Opening the reactor directly

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
9
after removal from the ice water bath helps to prevent rapid escape of gaseous acetaldehyde. The

reactions were performed in triplicate.

CTH of CELF lignin

The CTH of CELF lignin was carried out in a manner like that described above, with some

adaptations of a procedure reported previously.64 Briefly, a stainless steel microreactor (Swagelok,

P/N: SS-16-HLN-3.00) and a pair of 1.0-in FNPT stainless steel pipe caps (Swagelok, P/N: SS-

16-CP) was charged with 0.600 g CELF lignin, 0.060 g catalyst of interest (10% (w/w)), and 10

mL anhydrous ethanol. The microreactor was sealed and heated in the manner described above to

320 °C for 2 h. Subsequently, the reactor was quenched in an ice-water bath, opened, and decanted

through a glass fiber filter with a pore size of 0.5 micron. The lignin-rich filtrate was washed with

water and, subsequently, dichloromethane in a separatory funnel. The collected organic fractions

were dried over anhydrous sodium sulfate and evaporated to dryness to afford a brown lignin oil,

which was further characterized as described below. The reactions were performed in triplicate.

Characterization of products

All catalytic products (acetophenone and lignin) were characterized by GC/MS using an Agilent

7890A GC coupled with an Agilent 5975 MS (Agilent Technologies, Santa Clara, CA USA).

Chromatographic separation of a 1.0 μL injection aliquot was achieved on a 30-m Rxi-5MS GC

column (Restek, Bellefonte, PA USA) with a 0.25 mm ID and a 0.25 μm film. Helium was used

as the carrier gas with a flow rate of 1.0 mL/minute with the following GC oven parameters: hold

at 50 °C for 2.5 min, followed by a ramp of 10 °C/min to 275 °C, with a final hold of 5 min. The

mass spectrometer detector was operated in electron ionization mode with the following

parameters: transfer line at 275 °C, ion sources at 230 °C, full scan of m/z 30-550.

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
10
Lignin molecular weight (MW) analyses were performed using a Tosoh EcoSEC gel permeation

chromatography (GPC) system with a refractive index (RI) detector equipped with a flow reference

cell (Tosoh Bioscience, LLC, King of Prussia, PA USA). Lignin samples were modified by

acetobromination to enhance their solubility in THF using a procedure described elsewhere.65 The

flow rates of the instrument and reference cell were set to 0.35 mL/min and the analysis was

performed at 40 °C using 2 Tosoh TSKgel SuperMultipore HZ-M (4.6 x 150 mm; 4 μm) columns

and a TSKgel SuperMultipore HZ-M guard column. The instrument was calibrated using

polystyrene standards (580 Da – 2.379 MDa).

Band-selective gradient heteronuclear (1H-13C) single quantum coherence (HSQC) NMR

spectroscopy was performed using a Bruker AV III 600 equipped with a 5 mm TCI z-axis gradient

helium-cooled cryoprobe operating at 600.130 MHz for 1H and 150.903 MHz for 13C. Samples

were prepared using ~20 mg lignin dissolved in 0.5 mL dimethylsulfoxide-d6 (DMSO-d6). Data

collection and processing were performed as described elsewhere.66

RESULTS AND DISCUSSION

Catalyst synthesis and characterization

Metal oxide catalysts were synthesized using a urea-assisted synthesis.62 This method is useful

because it smoothly transforms the water-soluble metal salt precursors to metal carbonates by

thermal decomposition of urea to ammonium carbonate, which serves to raise the pH of the

aqueous solution, thereby precipitating the precursors as metal carbonate materials. Then, these

materials can be collected and transformed into the related metal oxides by calcining in flowing

air. In fact, the morphology of the resulting metal oxide particles can be controlled by using various

surfactants,67,68 although we made no attempts to exert such control for these experiments. In

addition, we used this process for the mono-metal oxides to synthesize bimetallic structures of

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
11
cobalt and nickel. Overall, this process could be relatively environmentally benign as it employs

aqueous solutions. Some additional wastewater treatment to remove nitrogen may be necessary,

which would negatively impact the overall process sustainability. These studies are ongoing in our

laboratory.

The atomic structure of the resulting metal oxides were characterized using powder x-ray

diffraction (Figure 1).

Figure 1. X-ray diffraction (XRD) pattern for synthesized (a) NiO and related ICDD card, (b)

NiCo2O4 and related ICDD cards, and (c) Co3O4 and related ICDD card.

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
12
The mono-metal nickel and cobalt oxide samples (Figure 1a and 1c) did not show any crystalline

impurities, nor were any indicators of amorphous material present. As expected,69 the cobalt oxide

crystallized as the mixed-valence Co3O4 spinel structure, with no evidence of CoO or Co2O3

present.

Powder XRD is insufficient to discriminate between the mixed-metal spinel NiCo2O4 and the

mono-metallic parent Co3O4 (Figure 1b and c). However, we used EPMA to determine the atomic

ratio of the mixed-metal spinel. These results show the atomic percentage of cobalt and nickel to

be 35 ± 6% and 15 ± 5% (w/w), respectively. This corresponds to a Co:Ni ratio of approximately

2:1 as expected for NiCo2O4. In addition, the shape and position of the Ni 2p spectrum of NiCo2O4

(Figure S1 in the Supporting Information) was similar to two other Ni-based spinels (NiFe2O4 and

NiCr2O4).70 The Co:Ni ratio was less than the expected 2:1, but XPS is extremely surface sensitive,

and there was evidence of hydroxides in the O 1s spectra. We interpret these observations as

evidence of a mixture of NiCo2O4 and nickel hydroxide at the surface of the powder.

We used the whole-pattern-fitting module in Jade (pseudo-Voigt fitting algorithm and third order

polynomial fitted background) to determine the amounts of spinel and excess NiO. We determined

that the sample contained 97.4% (w/w) NiCo2O4 and 2.6% (w/w) excess crystalline NiO. We

caution the reader that this amount falls within the standard error of the mean of measurements

using EPMA above (5-6%). Based on all these data, although we are unable to unequivocally

determine the amount of excess NiO present, the existence of a small amount of it in the NiCo2O4

sample is a reasonable conclusion.

Catalyst screening for transfer hydrogenation/hydrogenolysis

We then moved to screen our putative catalyst species for catalytic transfer

hydrogenation/hydrogenolysis of acetophenone using ethanol. While we are ultimately interested

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
13
in CTH reactions of lignin, and acetophenone is not a particularly realistic lignin model compound,

we chose acetophenone because it is a straightforward model to demonstrate whether our catalyst

species could moderate transformations of polar double bonds (i.e., the C=O bond in

acetophenone) in the presence of non-polar double bonds (i.e., the aromatic ring in acetophenone).

For example, Gilhaume and coworkers demonstrated reduction of meso carbons in lignin model

compounds was possible using Pt/C in ethanol/water mixtures.46 Both carbonyl and alcohol groups

could be reduced to alkanes using a Pt-containing catalyst. Such reactivity is crucial to controlling

the oxidation state at the meso (benzylic) carbon in lignin, which we and others have identified as

important for controlling the chemical and physical properties of isolated lignin with respect to its

use in industrially relevant products.

We initially hypothesized a CTH product slate based on sequential transfer hydrogenation of

acetophenone to 1-phenylethanol followed by transfer hydrogenolysis of 1-phenylethanol to

ethylbenzene. We also hypothesized that 1-phenylethanol could undergo a dehydration reaction to

form styrene. This product slate and related reaction scheme are depicted in Scheme 2 below.

Scheme 2. Hypothetical reaction scheme for CTH of acetophenone using metal oxide catalysts

and ethanol.

The hypotheses were tested by conducting the CTH reactions in stainless steel microreactors at

320 °C for 2 h. Ethanol was used as the solvent and reductant (Scheme 2) because it can be easily

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
14
isolated from biomass; in an industrial biorefinery, isolating cellulosic ethanol and upgraded small

molecules from lignin as coproducts represents an economically and ecologically favorable

industrial process.71–73 GC/MS was used to monitor the concentrations of the molecules shown in

Scheme 2, using durene as an internal standard and constructing a response factor curve for each

molecule. Each catalyst species was tested as was a blank reaction with no catalyst present. The

results of these experiments are summarized in Table 1.

Table 1. Catalyst screening and reactivity test.

Catalysts Xa (%) Y; Sb (%) Y; S (%) Y; S (%)

Blank 50.8(2)c 51.3(8); 101(2) NDd ND


NiO 76.9(1) 1.8(1); 2.3(1) ND 23.1(9); 30(1)
Co3O4 42.2(3) 8.3(2); 19.7(5) ND 29(3); 69(7)
NiCo2O4 74.5(4) 7.0(1); 9.4(1) ND 66.6(4); 89.4(7)
a
X = conversion
b
Y = yield, S = selectivity
c
Number in parentheses shows the standard error of the mean in the preceding digit.
d
ND = not detected

In the absence of catalyst, roughly half of the acetophenone present was converted to products.

This is perhaps unsurprising given the harsh reaction conditions (320 °C for 2 h); yet, we were

intrigued by the result that, in the absence of catalyst, the reaction is 100% selective for reduction

of acetophenone to 1-phenylethanol. No other products were detectable in this reaction, including

the hypothesized ethylbenzene or styrene.

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
15
When using NiO, the conversion of acetophenone rises to 77%. However, the selectivity for 1-

phenylethanol decreases substantially. We were also able to detect ethylbenzene, albeit with

relatively low selectivity (30%). Although not quantified, both butylbenzene and 1-

ethoxyethylbenzene (the ethyl ether of 1-phenylethanol) were detected. The formation of

butylbenzene occurs via aldol condensation as seen in other Meerwein-Ponndorf-Verley (MPV)

reductions of aldehydes and ketones to alcohols,39 whereas dehydrative etherification could afford

1-ethoxyethylbenzene. The presence of these molecules could account for the poor carbon

recovery shown for this reaction.

In the case of Co3O4, the conversion of acetophenone (42%) was comparatively low, but the

selectivity for 1-phenylethanol (69%) was higher than that of NiO. We again noted the presence

of butylbenzene and 1-ethoxyethylbenzene. However, their formation was evidently suppressed as

more of the desired 1-phenylethanol and ethylbenzene were formed, despite the lower overall

conversion of acetophenone.

When using NiCo2O4, conversion of acetophenone is comparatively high (75%), with high

selectivity for formation of ethylbenzene (89%). While we are unsure of the exact reason for the

enhanced conversion and selectivity when using the spinel as compared to the monometallic

species, there are likely structural differences in the active sites of these materials that warrant

further inquiry. These investigations are ongoing in our laboratory.

Styrene was undetectable in any of the reactions we performed. This was surprising because the

alcohol should be dehydrated with relative ease to form the alkene. Xu and coworkers

demonstrated this reactivity on birch lignin using supported Ni(0) catalysts.74 They demonstrated

a Cα=Cβ bond formation in monomers isolated from birch lignin via transfer hydrogenolysis in

methanol. The ensuing C=C bond was further hydrogenated to alkane using Ni/C. For the case

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
16
reported here, we speculate that the relative reaction kinetics preclude styrene accumulation in our

system. Specifically, as shown in Scheme 2, the rate constant for dehydration (kdehydration) may be

much lower than the rate constant for hydrogenation of styrene (khydrogenation). In this case, styrene

would be formed slowly and quickly consumed to form ethylbenzene. We are currently exploring

this reactivity for our system using isotopic labeling experiments.

CTH of CELF Lignin

Turning our attention to CTH of lignin using the Co-Ni spinel catalyst, we isolated lignin using

the CELF process from switchgrass. The switchgrass contained 23% (w/w) lignin as shown using

the ABSL protocol,60 which is in agreement with previous results.75 From a feedstock of 80 g, 5.4

g of high purity lignin on dry weight basis (29% yield) was obtained.

A series of experiments analogous to those for our model compound above were performed

using our as-isolated CELF lignin as the substrate with the NiCo2O4 spinel catalyst in ethanol at

320 °C for 2 h. The molecular weight of the lignin samples before and after reaction was

determined using gel permeation chromatography, and the results of these measurements are

shown in Table 2. Apparent molecular weight distributions are displayed in the Supporting

Information (Figure S2).

Table 2. Average Molecular Weight Distribution for CELF Lignin before and after CTH

Sample Mw (g/mol) Mn (g/mol) Ð

CELF lignin (Starting material) 3604 1851 1.947


Control reaction 1533 866 1.77
Reaction with NiCo2O4 2128 984 2.16

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
17
The as-isolated CELF lignin had a weight-average molecular weight (Mw) of 3604 g/mol, a

number-average molecular weight (Mn) of 1851 g/mol, and a dispersion (Ð, sometimes called

polydispersity, PDI) of 1.947. These values are within the range of molecular weights and Ð

reported previously for CELF-derived lignins for the reaction conditions explored here.76,77 The

control reaction (320 °C for 2 h in the absence of catalyst) and the catalytic reaction produced

lignin samples that displayed nominally smaller molecular weights with a tighter dispersion for

the control reaction. Our overall conclusion is that the reaction conditions and the presence of the

catalyst produce lignin fractions with lower molecular weight distributions. This result agrees with

our previous results using mixed-metal B-containing catalysts.39

A semi-quantitative analysis of the product slate from both the uncatalyzed and catalyzed

reactions on CELF lignin was performed using GC/MS. 1,2,4,5-tetramethylbenzene (durene) was

used as an internal standard, and the peak areas of 10 compounds were normalized to track how

their concentrations changed under reaction conditions with and without catalyst. Each experiment

was performed three times. A nonparametric Mann-Whitney U test was applied to determine the

relative statistical certainty of the noted changes of the product slate using the achieved confidence

level. For example, a confidence level > 95% is mathematically equivalent to a p-value less than

α = 0.05 and indicates, with at least 95% certainty, that the amount of compound either increased

or decreased in the presence of catalyst. The results of these experiments are displayed in Figure

2.

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
18
Figure 2. Plot of the means (3 replicates) of normalized peak area of 10 compounds tracked in the

product slate of the catalytic reaction using CELF lignin. Peak areas were normalized against an

internal standard of durene. For each compound 1-10, the left peak is the normalized peak area for

the uncatalyzed reaction, and the right peak is that for the catalytic reaction. The probability that

the median values are statistically different with an achieved confidence of 90, 80, and 30% (as

verified by a Mann-Whitney U test using Minitab) are shaded dark gray, light gray, and white,

respectively. Error bars represent the standard error of the mean.

The tracked product compounds can be roughly divided into H- (1-2), G- (3-8), and S-type (9-

10) monomers. We attempted to track the change in concentration of each of these as catalyst was

applied to the reaction. Given our interest in the oxidation state of the meso (benzylic) carbon, we

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
19
chose compounds that both do (8-10) and do not (1-6) contain a ketone (or aldehyde) at the α-

position (see Scheme 1 above). In addition, we also tracked eugenol (7) since it contains a reducible

double bond.

Overall, the concentrations of compounds such as 2 and 4-6 increased with a high degree of

statistical certainty (α = 0.1). The NiCo2O4 catalyst appears to enhance the formation of these

compounds that contain unoxidized meso carbons (i.e., methyl, ethyl, and propyl side chains). This

agrees with the model compound studies described above, which show that acetophenone is

converted to ethylbenzene with near 90% selectivity.

In addition, the concentration of compounds such as 9 decreased with a high-to-moderate degree

of statistical certainty (α = 0.2). Based on the model compound studies, we expected a decrease in

concentration of compounds containing an aldehyde (or ketone, see below). In those cases, the

oxygen-containing moiety would be reduced to an aliphatic side chain (analogous to the reduction

of acetophenone to ethylbenzene).

In contrast, the concentration of acetovanillone (8) remained constant per the low statistical

certainty displayed (α = 0.7). Our expectation was that the amount of acetovanillone would

decrease in the presence of the catalyst since it contains a reducible ketone moiety and is

structurally like acetophenone. The reason for this deviation from our hypothesis is unclear; the

reactivity of this substrate with the Co-Ni spinel catalyst may be different due to the presence of

electronic-modulating groups on the aromatic ring.

Interestingly, the concentration of eugenol (7) was reduced in the presence of the catalyst with

high-to-moderate statistical certainty (α = 0.2). If the catalyst was able to moderate C=C bond

reduction reactions, this could explain why styrene was not a detectable product in the model

compound reactions (see above). If styrene was formed, it could be immediately reduced to

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
20
ethylbenzene based on the relative reaction kinetics (Scheme 2). In any event, these two results

suggest there is some selectivity that can be leveraged for this catalyst. We acknowledge the

limited statistical power of our analysis in this determination and are continuing to investigate

these phenomena.

Although GC/MS is a useful tool to separate and identify mixtures of compounds, using it to

unambiguously describe the effects of a catalytic reaction with a complex product slate containing

several oligomeric (or polymeric) compounds is challenging. The resulting chromatogram may

have dozens of peaks, some overlapping, and the limited volatility of some products may make

them undetectable in the mass spectral detector. Alternatively, 2D NMR spectroscopy has been

successfully used to characterize lignin, since it will provide structural information on all the

soluble components in the sample. As above, 2D 1H-13C HSQC NMR spectroscopy was used to

characterize the product slates of the reactions (Figure 3).

Figure 3. Two-dimensional 1H-13C HSQC NMR spectra of untreated CELF lignin (A) and

catalytically treated CELF lignin using NiCo2O4 catalyst (B). Peaks are marked with triangular

flags and structural information is color-coded as shown on the right.

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
21
2D NMR spectroscopy was used to determine the chemical shift values of the aromatic

hydrogens and carbons, which are sensitive to the oxidation state of the meso (benzylic) carbon.

Chemical shift values were assigned based on the seminal work on 2D HSQC spectroscopy of

biorefinery switchgrass lignin by Bozell and coworkers66 Their study includes a thorough

assignment of peaks arising from changes in the oxidation state at the alpha-carbon in the so-called

“aromatic region” of lignin (colored labels in Figure 3).

As-isolated CELF lignin contains myriad oxidized moieties (Figure 3a), including S-, G-, and

H-type units with carbonyl groups at the alpha position and hydroxycinnamic (caffeic, CA) and

ferulic acid (FA). The molecular symmetry-dependent chemical shift values for the 2,6 (for H- and

S-type moieties) and the 2, 5, and 6 (for G-type moieties) H–C groups are highly sensitive to the

oxidation state of the alpha-carbon, and these moieties resonate far downfield in the proton

dimension of the related standard S-, G-, and H- signals. For example, the peak for oxS2,6 (yellow)

is clearly distinguishable from other S2,6 signals (labeled in gray).

Unsurprisingly, the peaks associated with oxS2,6, oxH2,6, and CA/FA are undetectable in the NMR

spectrum associated with catalytically treated lignin (Figure 3b). This agrees with our model

compound studies and the GC/MS results above, which indicate that the NiCo2O4 can reduce

oxidized alpha-carbon moieties to the associated aliphatic side chain. The region associated with

G-type aromatic H–C groups is of lower resolution, given the molecular asymmetry of a guaiacyl-
ox
type molecule (Figure 3, blue color). The presence of G groups is less easily resolvable.

However, our GC/MS results suggest that acetovanillone is relatively unchanged with catalytic

reaction, which further suggests there could be some electronic differences with this substrate-

catalyst combinations containing G-type moieties, warranting further investigation.

Sustainability of Cobalt

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
22
No modern catalyst study would be complete without considering the long-term sustainability

issues of the catalyst materials. Transition metals are widely regarded as effective catalysts; their

relative abundance within the Earth’s crust could enable their long-term adoption for use as

industrial catalyst materials. We selected Ni- and Co-based catalysts for this study because their

estimated concentrations within the Earth’s crust are around 90 ppm and 30 ppm, respectively. In

contrast, metals commonly used for hydrogenolysis catalysts, such as Pd and Ru, have estimated

concentrations of 6.3 parts per billion and 1 ppb, respectively.78 However, relative abundance is

not the only aspect to consider when evaluating the sustainability of catalyst materials.

Most lifecycle assessment studies performed for cobalt-based catalysts have focused on the

environmental impact of cobalt mining (environmental lifecycle assessment, e-LCA). Social

lifecycle assessment (s-LCA) is gaining interest as a complementary method, because, despite

cobalt’s high earth abundance,79 the social impacts of metal extraction draws sharp criticism from

some stakeholders. These social impacts include exposure to metals through consumption of edible

products around the mine fields, reduced air quality from the dust particles generated during

mining activities, exploitation of mine workers and mining-adjacent communities, and use of child

labor, among other environmental and social justice issues.80

Studies related to cobalt mine fields in the Democratic Republic of Congo (DRC) is essential,

as the DRC holds over half of the world cobalt reserves. To responsibly source cobalt, one study

proposed implementing a system that considers artisanal and small-scale mining (ASM) and the

more established downstream sectors, which can help improve quality of data needed for s-LCA.81

The United Nations Environment Program (UNEP) proposed guidelines for studying the social

impact through the life cycle of products effectively, including impacts on workers, the local

community, and value chain actors such as consumers, children, and society as a whole.82,83 In

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
23
addition to promoting uniform evaluations of the social and socio-economic impacts of products

throughout their lifecycle, the UNEP guidelines will help companies make informed decisions,

improve their social performance, and meet the expectations of socially responsible consumers.84

While cobalt’s effectiveness in materials for chemical catalysis and as critical components for

emerging technologies such as electric vehicles is well-documented, the need to consider its social

impact in future sustainability studies cannot be overstated.

CONCLUSION

We have demonstrated the effectiveness of NiCo2O4 as a catalyst for transfer hydrogenolysis of

CELF lignin using biomass-derivable ethanol as the solvent and reductant. Using acetophenone to

probe this reactivity, we obtained 75% conversion with ethylbenzene selectivity nearing 90%.

Moreover, we observed a statistically significant increase in deoxygenated compounds in CELF

lignin after catalytic transfer hydrogenolysis with NiCo2O4. GC/MS and NMR spectroscopy

results confirm the reduction occurs at the alpha carbon, which could unlock lignin’s potential for

conversion to platform molecules for new biopriviledged materials. Further studies to characterize

the mechanism of this reaction as well as the lifecycle implications of using cobalt for lignin

valorization are ongoing in our laboratory.

AUTHOR INFORMATION

Corresponding Author

* E-mail: sc411@psu.edu

ORCID

Stephen C. Chmely: 0000-0002-2637-9974

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
24
Notes

The authors declare no competing financial interest.

Author Contributions

The manuscript was written through contributions of all authors. All authors approve of the final

version of the manuscript.

Funding Sources

This work was supported by funding from USDA-NIFA Project PEN04671 and Accession number

1017582.

SUPPORTING INFORMATION

Supporting information: High-resolution Ni 2p and Co 2p XPS spectra for all catalyst species

and GPC molecular weight distribution curves for lignin samples.

ACKNOWLEDGMENT

The authors wish to acknowledge the Penn State Materials Characterization Lab for use of the

Malvern Panalytical Empyrean diffractometer, the Cameca SXFive Electron Probe Micro-

Analyzer (EMPA), and the Physical Electronics VersaProbe III x-ray photoelectron spectrometer

(XPS). We would also like to acknowledge the support of Dr. Christy George and Dr. Tapas Mal

for their help with collecting 2D-HSQC NMR data and Dr. Christy Payne for her helpful

discussions about the manuscript. Finally, we wish to acknowledge the support of Prof. Charles

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
25
Anderson, Dr. Niroshan Siva, and Hannah Klatte for their help with the ABSL assay data

collection.

REFERENCES

(1) Boerjan, W.; Ralph, J.; Baucher, M. Lignin Biosynthesis. Annu. Rev. Plant Biol. 2003, 54,

519–546. https://doi.org/10.1146/annurev.arplant.54.031902.134938.

(2) Strassberger, Z.; Tanase, S.; Rothenberg, G. The Pros and Cons of Lignin Valorisation in

an Integrated Biorefinery. RSC Adv. 2014, 4 (48), 25310–25318.

https://doi.org/10.1039/c4ra04747h.

(3) Wang, D.; Lee, S. H.; Kim, J.; Park, C. B. “Waste to Wealth”: Lignin as a Renewable

Building Block for Energy Harvesting/Storage and Environmental Remediation. ChemSusChem

2020, 13 (11), 2807–2827. https://doi.org/10.1002/cssc.202000394.

(4) Xu, J.; Li, C.; Dai, L.; Xu, C.; Zhong, Y.; Yu, F.; Si, C. Biomass Fractionation and Lignin

Fractionation towards Lignin Valorization. ChemSusChem 2020, 13 (17), 4284–4295.

https://doi.org/10.1002/cssc.202001491.

(5) Shanks, B. H.; Keeling, P. L. Bioprivileged Molecules: Creating Value from Biomass.

Green Chem. 2017, 19 (14), 3177–3185. https://doi.org/10.1039/c7gc00296c.

(6) Abu-Omar, M. M.; Barta, K.; Beckham, G. T.; Luterbacher, J. S.; Ralph, J.; Rinaldi, R.;

Román-Leshkov, Y.; Samec, J. S. M.; Sels, B. F.; Wang, F. Guidelines for Performing Lignin-

First Biorefining. Energy Environ. Sci. 2021, 14 (1), 262–292.

https://doi.org/10.1039/d0ee02870c.

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
26
(7) Ferrini, P.; Chesi, C.; Parkin, N.; Rinaldi, R. Effect of Methanol in Controlling

Defunctionalization of the Propyl Side Chain of Phenolics from Catalytic Upstream Biorefining.

Faraday Discuss. 2017, 202, 403–413. https://doi.org/10.1039/c7fd00069c.

(8) Ferrini, P.; Rezende, C. A.; Rinaldi, R. Catalytic Upstream Biorefining through Hydrogen

Transfer Reactions: Understanding the Process from the Pulp Perspective. ChemSusChem 2016, 9

(22), 3171–3180. https://doi.org/10.1002/cssc.201601121.

(9) Graça, I.; Woodward, R. T.; Kennema, M.; Rinaldi, R. Formation and Fate of Carboxylic

Acids in the Lignin-First Biorefining of Lignocellulose via H-Transfer Catalyzed by Raney Ni.

ACS Sustain. Chem. Eng. 2018, 6 (10), 13408–13419.

https://doi.org/10.1021/acssuschemeng.8b03190.

(10) Kennema, M.; De Castro, I. B. D.; Meemken, F.; Rinaldi, R. Liquid-Phase H-Transfer from

2-Propanol to Phenol on Raney Ni: Surface Processes and Inhibition. ACS Catal. 2017, 7 (4),

2437–2445. https://doi.org/10.1021/acscatal.6b03201.

(11) Chesi, C.; de Castro, I. B. D.; Clough, M. T.; Ferrini, P.; Rinaldi, R. The Influence of

Hemicellulose Sugars on Product Distribution of Early-Stage Conversion of Lignin Oligomers

Catalysed by Raney Nickel. ChemCatChem 2016, 8 (12), 2079–2088.

https://doi.org/10.1002/cctc.201600235.

(12) Renders, T.; Van den Bossche, G.; Vangeel, T.; Van Aelst, K.; Sels, B. Reductive Catalytic

Fractionation: State of the Art of the Lignin-First Biorefinery. Curr. Opin. Biotechnol. 2019, 56,

193–201. https://doi.org/10.1016/j.copbio.2018.12.005.

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
27
(13) Hayes, G.; Laurel, M.; MacKinnon, D.; Zhao, T.; Houck, H. A.; Becer, C. R. Polymers

without Petrochemicals: Sustainable Routes to Conventional Monomers. Chem. Rev. 2022.

https://doi.org/10.1021/acs.chemrev.2c00354.

(14) Sun, Z.; Fridrich, B.; De Santi, A.; Elangovan, S.; Barta, K. Bright Side of Lignin

Depolymerization: Toward New Platform Chemicals. Chem. Rev. 2018, 118 (2), 614–678.

https://doi.org/10.1021/acs.chemrev.7b00588.

(15) Brunner, E. Solubility of Hydrogen in 10 Organic Solvents at 298.15, 323.15, and 373.15

K. 1985, 269–273. https://doi.org/https://pubs.acs.org/doi/epdf/10.1021/je00041a010.

(16) Bartling, A. W.; Stone, M. L.; Hanes, R. J.; Bhatt, A.; Zhang, Y.; Biddy, M. J.; Davis, R.;

Kruger, J. S.; Thornburg, N. E.; Luterbacher, J. S.; Rinaldi, R.; Samec, J. S. M.; Sels, B. F.; Román-

Leshkov, Y.; Beckham, G. T. Techno-Economic Analysis and Life Cycle Assessment of a

Biorefinery Utilizing Reductive Catalytic Fractionation. Energy Environ. Sci. 2021, 14 (8), 4147–

4168. https://doi.org/10.1039/d1ee01642c.

(17) Kim, S. H.; Kumar, G.; Chen, W. H.; Khanal, S. K. Renewable Hydrogen Production from

Biomass and Wastes (ReBioH2-2020). Bioresour. Technol. 2021, 331.

https://doi.org/10.1016/j.biortech.2021.125024.

(18) US DOE. Enabling A Low-Carbon Economy. Washington, DC 20585 2020, 24.

https://doi.org/https://www.energy.gov/sites/prod/files/2020/07/f76/USDOE_FE_Hydrogen_Stra

tegy_July2020.pdf.

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
28
(19) Zhao, B.; Hu, Y.; Gao, J.; Zhao, G.; Ray, M. B.; Xu, C. C. Recent Advances in

Hydroliquefaction of Biomass for Bio-Oil Production Using in Situ Hydrogen Donors. Ind. Eng.

Chem. Res. 2020, 59 (39), 16987–17007. https://doi.org/10.1021/acs.iecr.0c01649.

(20) Wang, D.; Astruc, D. The Golden Age of Transfer Hydrogenation. Chem. Rev. 2015, 115

(13), 6621–6686. https://doi.org/10.1021/acs.chemrev.5b00203.

(21) Gilkey, M. J.; Xu, B. Heterogeneous Catalytic Transfer Hydrogenation as an Effective

Pathway in Biomass Upgrading. ACS Catal. 2016, 6 (3), 1420–1436.

https://doi.org/10.1021/acscatal.5b02171.

(22) Jin, W.; Pastor-Pérez, L.; Shen, D. K.; Sepúlveda-Escribano, A.; Gu, S.; Ramirez Reina,

T. Catalytic Upgrading of Biomass Model Compounds: Novel Approaches and Lessons Learnt

from Traditional Hydrodeoxygenation – a Review. ChemCatChem 2019, 11 (3), 924–960.

https://doi.org/10.1002/cctc.201801722.

(23) Ferlin, F.; Valentini, F.; Marrocchi, A.; Vaccaro, L. Catalytic Biomass Upgrading

Exploiting Liquid Organic Hydrogen Carriers (LOHCs). ACS Sustain. Chem. Eng. 2021, 9 (29),

9604–9624. https://doi.org/10.1021/acssuschemeng.1c03247.

(24) Jiang, M.; Chen, X.; Wang, L.; Liang, J.; Wei, X.; Nong, W. Anchoring Single Ni Atoms

on CeO2 Nanospheres as an Efficient Catalyst for the Hydrogenolysis of Lignin to Aromatic

Monomers. Fuel 2022, 324 (PA), 124499. https://doi.org/10.1016/j.fuel.2022.124499.

(25) Philippov, A. A.; Nesterov, N. N.; Pakharukova, V. P.; Martyanov, O. N. High-Loaded Ni-

Based Catalysts Obtained via Supercritical Antisolvent Coprecipitation in Transfer Hydrogenation

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
29
of Anisole: Influence of the Support. Appl. Catal. A Gen. 2022, 643 (May), 118792.

https://doi.org/10.1016/j.apcata.2022.118792.

(26) Li, H.; Liu, M.; Zou, W.; Lv, Y.; Liu, Y.; Chen, L. Selective Hydrodeoxygenation of Lignin

and Its Derivatives without Initial Reaction Pressure Using MOF-Derived Carbon-Supported

Nickel Composites. ACS Sustain. Chem. Eng. 2022, 10 (17), 5430–5440.

https://doi.org/10.1021/acssuschemeng.1c08144.

(27) Chen, C.; Liu, P.; Zhou, M.; Sharma, B. K.; Jiang, J. Selective Hydrogenation of Phenol to

Cyclohexanol over Ni/CNT in the Absence of External Hydrogen. Energies 2020, 13 (4), 1–12.

https://doi.org/10.3390/en13040846.

(28) Philippov, A. A.; Chibiryaev, A. M.; Martyanov, O. N. Raney® Nickel-Catalyzed

Hydrodeoxygenation and Dearomatization under Transfer Hydrogenation Conditions—Reaction

Pathways of Non-Phenolic Compounds. Catal. Today 2020, 355 (December 2018), 35–42.

https://doi.org/10.1016/j.cattod.2019.05.033.

(29) Huang, Y. B.; Zhang, J. L.; Zhang, X.; Luan, X.; Chen, H. Z.; Hu, B.; Zhao, L.; Wu, Y. L.;

Lu, Q. Catalytic Depolymerization of Lignin via Transfer Hydrogenation Strategy over Skeletal

CuZnAl Catalyst. Fuel Process. Technol. 2022, 237 (April), 107448.

https://doi.org/10.1016/j.fuproc.2022.107448.

(30) Chen, K.; Sang, J.; Wang, Z.; Ibrahim, U. K.; Xia, W.; Guo, A.; Zhang, J.; Hou, D.

Production of Low-Oxygenated Bio-Fuels (Hydrocarbons and Polymethylphenols) from

Lignocellulose by a Two-Stage Strategy with Non-Noble Metal Catalysts. Fuel 2021, 286 (P2),

119401. https://doi.org/10.1016/j.fuel.2020.119401.

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
30
(31) Chui, M.; Metzker, G.; Bernt, C. M.; Tran, A. T.; Burtoloso, A. C. B.; Ford, P. C. Probing

the Lignin Disassembly Pathways with Modified Catalysts Based on Cu-Doped Porous Metal

Oxides. ACS Sustain. Chem. Eng. 2017, 5 (4), 3158–3169.

https://doi.org/10.1021/acssuschemeng.6b02954.

(32) Hossain, M. A.; Phung, T. K.; Rahaman, M. S.; Tulaphol, S.; Jasinski, J. B.;

Sathitsuksanoh, N. Catalytic Cleavage of the Β-O-4 Aryl Ether Bonds of Lignin Model

Compounds by Ru/C Catalyst. Appl. Catal. A Gen. 2019, 582 (November 2018), 117100.

https://doi.org/10.1016/j.apcata.2019.05.034.

(33) Khan, T. S.; Singh, D.; Samal, P. P.; Krishnamurty, S.; Dhepe, P. L. Mechanistic

Investigations on the Catalytic Transfer Hydrogenation of Lignin-Derived Monomers over Ru

Catalysts: Theoretical and Kinetic Studies. ACS Sustain. Chem. Eng. 2021, 9 (42), 14040–14050.

https://doi.org/10.1021/acssuschemeng.1c02942.

(34) Singh, D.; Dhepe, P. L. An Efficient Catalytic Transfer Hydrogenation-

Hydrodeoxygenation of Lignin Derived Monomers: Investigating Catalyst Properties-Activity

Correlation. Catal. Commun. 2021, 149 (November 2020), 106220.

https://doi.org/10.1016/j.catcom.2020.106220.

(35) Li, T.; Lin, H.; Ouyang, X.; Qiu, X.; Wan, Z.; Ruan, T. Impact of Nitrogen Species and

Content on the Catalytic Activity to C–O Bond Cleavage of Lignin over N-Doped Carbon

Supported Ru-Based Catalyst. Fuel 2020, 278 (April), 118324.

https://doi.org/10.1016/j.fuel.2020.118324.

(36) Wu, H.; Song, J.; Xie, C.; Wu, C.; Chen, C.; Han, B. Efficient and Mild Transfer

Hydrogenolytic Cleavage of Aromatic Ether Bonds in Lignin-Derived Compounds over Ru/C.

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
31
ACS Sustain. Chem. Eng. 2018, 6 (3), 2872–2877.

https://doi.org/10.1021/acssuschemeng.7b02993.

(37) Kristianto, I.; Limarta, S. O.; Lee, H.; Ha, J. M.; Suh, D. J.; Jae, J. Effective

Depolymerization of Concentrated Acid Hydrolysis Lignin Using a Carbon-Supported Ruthenium

Catalyst in Ethanol/Formic Acid Media. Bioresour. Technol. 2017, 234, 424–431.

https://doi.org/10.1016/j.biortech.2017.03.070.

(38) Xuezhong Wu a, b, Wenqian Jiao a, Bing‐Zheng Li b, Yanming Li b, Yahong Zhang a,

Quanrui Wang c, Y. T. a. Decomposition of a Β‐O‐4 Lignin Model Compound over Solid Cs‐

substituted Polyoxometalates in Anhydrous Ethanol: Acidity or Redox Property Dependence?

Chinese J. Catal. 2017, 38 (7), 1216–1228. https://doi.org/10.1016/S1872.

(39) Regmi, Y. N.; Mann, J. K.; McBride, J. R.; Tao, J.; Barnes, C. E.; Labbé, N.; Chmely, S.

C. Catalytic Transfer Hydrogenolysis of Organosolv Lignin Using B-Containing FeNi Alloyed

Catalysts. Catal. Today 2018, 302, 190–195. https://doi.org/10.1016/j.cattod.2017.05.051.

(40) Awan, I. Z.; Beltrami, G.; Bonincontro, D.; Gimello, O.; Cacciaguerra, T.; Tanchoux, N.;

Martucci, A.; Albonetti, S.; Cavani, F.; Di Renzo, F. Copper-Nickel Mixed Oxide Catalysts from

Layered Double Hydroxides for the Hydrogen-Transfer Valorisation of Lignin in Organosolv

Pulping. Appl. Catal. A Gen. 2021, 609 (November 2020), 117929.

https://doi.org/10.1016/j.apcata.2020.117929.

(41) Espro, C.; Gumina, B.; Paone, E.; Mauriello, F. Upgrading Lignocellulosic Biomasses:

Hydrogenolysis of Platform Derived Molecules Promoted by Heterogeneous Pd-Fe Catalysts.

Catalysts 2017, 7 (3). https://doi.org/10.3390/catal7030078.

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
32
(42) Mauriello, F.; Ariga-Miwa, H.; Paone, E.; Pietropaolo, R.; Takakusagi, S.; Asakura, K.

Transfer Hydrogenolysis of Aromatic Ethers Promoted by the Bimetallic Pd/Co Catalyst. Catal.

Today 2020, 357 (November 2018), 511–517. https://doi.org/10.1016/j.cattod.2019.06.071.

(43) Mauriello, F.; Paone, E.; Pietropaolo, R.; Balu, A. M.; Luque, R. Catalytic Transfer

Hydrogenolysis of Lignin-Derived Aromatic Ethers Promoted by Bimetallic Pd/Ni Systems. ACS

Sustain. Chem. Eng. 2018, 6 (7), 9269–9276. https://doi.org/10.1021/acssuschemeng.8b01593.

(44) Guan, W.; Chen, X.; Zhang, J.; Hu, H.; Liang, C. Catalytic Transfer Hydrogenolysis of

Lignin α-O-4 Model Compound 4-(Benzyloxy)Phenol and Lignin over Pt/HNbWO6/CNTs

Catalyst. Renew. Energy 2020, 156, 249–259. https://doi.org/10.1016/j.renene.2020.04.078.

(45) Fraga, G.; Yin, Y.; Konarova, M.; Hasan, M. D.; Laycock, B.; Yuan, Q.; Batalha, N.; Pratt,

S. Hydrocarbon Hydrogen Carriers for Catalytic Transfer Hydrogenation of Guaiacol. Int. J.

Hydrogen Energy 2020, 45 (51), 27381–27391. https://doi.org/10.1016/j.ijhydene.2020.07.136.

(46) Besse, X.; Schuurman, Y.; Guilhaume, N. Reactivity of Lignin Model Compounds through

Hydrogen Transfer Catalysis in Ethanol/Water Mixtures. Appl. Catal. B Environ. 2017, 209, 265–

272. https://doi.org/10.1016/j.apcatb.2017.03.013.

(47) Smith, C. A.; Brandi, F.; Al-Naji, M.; Guterman, R. Resin-Supported Iridium Complex for

Low-Temperature Vanillin Hydrogenation Using Formic Acid in Water. RSC Adv. 2021, 11 (26),

15835–15840. https://doi.org/10.1039/d1ra01460a.

(48) Sharma, V.; Getahun, T.; Verma, M.; Villa, A.; Gupta, N. Carbon Based Catalysts for the

Hydrodeoxygenation of Lignin and Related Molecules: A Powerful Tool for the Generation of

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
33
Non-Petroleum Chemical Products Including Hydrocarbons. Renew. Sustain. Energy Rev. 2020,

133 (August), 110280. https://doi.org/10.1016/j.rser.2020.110280.

(49) Espro, C.; Gumina, B.; Szumelda, T.; Paone, E.; Mauriello, F. Catalytic Transfer

Hydrogenolysis as an Effective Tool for the Reductive Upgrading of Cellulose, Hemicellulose,

Lignin, and Their Derived Molecules. Catalysts 2018, 8 (8). https://doi.org/10.3390/catal8080313.

(50) Zhang, J. Catalytic Transfer Hydrogenolysis as an Efficient Route in Cleavage of Lignin

and Model Compounds. Green Energy Environ. 2018, 3 (4), 328–334.

https://doi.org/10.1016/j.gee.2018.08.001.

(51) Thallada, B.; Kumar, A.; Jindal, M.; Maharana, S. Lignin Biorefinery: New Horizons in

Catalytic Hydrodeoxygenation for the Production of Chemicals. Energy and Fuels 2021, 35 (21),

16965–16994. https://doi.org/10.1021/acs.energyfuels.1c01651.

(52) Tang, D.; Huang, X.; Tang, W.; Jin, Y. Lignin-to-Chemicals: Application of Catalytic

Hydrogenolysis of Lignin to Produce Phenols and Terephthalic Acid via Metal-Based Catalysts.

Int. J. Biol. Macromol. 2021, 190 (September), 72–85.

https://doi.org/10.1016/j.ijbiomac.2021.08.188.

(53) Advani, J. H.; More, G. S.; Srivastava, R. Spinel-Based Catalysts for the Biomass

Valorisation of Platform Molecules via Oxidative and Reductive Transformations. Green Chem.

2022, 3574–3604. https://doi.org/10.1039/d1gc04592j.

(54) Sadeghifar, H.; Ragauskas, A. Lignin as a UV Light Blocker-a Review. Polymers (Basel).

2020, 12 (5), 1–10. https://doi.org/10.3390/POLYM12051134.

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
34
(55) Badazhkova, V. D.; Savela, R.; Leino, R. Selective Modification of Hydroxyl Groups in

Lignin Model Compounds by Ruthenium-Catalyzed Transfer Hydrogenation. Dalt. Trans. 2022,

51 (17), 6587–6596. https://doi.org/10.1039/d2dt00267a.

(56) Li, Q.; Naik, M. T.; Lin, H. S.; Hu, C.; Serem, W. K.; Liu, L.; Karki, P.; Zhou, F.; Yuan,

J. S. Tuning Hydroxyl Groups for Quality Carbon Fiber of Lignin. Carbon N. Y. 2018, 139, 500–

511. https://doi.org/10.1016/j.carbon.2018.07.015.

(57) Cai, C. M.; Zhang, T.; Kumar, R.; Wyman, C. E. THF Co-Solvent Enhances Hydrocarbon

Fuel Precursor Yields from Lignocellulosic Biomass. Green Chem. 2013, 15 (11), 3140–3145.

https://doi.org/10.1039/c3gc41214h.

(58) Nguyen, T. Y.; Cai, C. M.; Kumar, R.; Wyman, C. E. Co-Solvent Pretreatment Reduces

Costly Enzyme Requirements for High Sugar and Ethanol Yields from Lignocellulosic Biomass.

ChemSusChem 2015, 8 (10), 1716–1725. https://doi.org/10.1002/cssc.201403045.

(59) Nguyen, T. Y.; Cai, C. M.; Osman, O.; Kumar, R.; Wyman, C. E. CELF Pretreatment of

Corn Stover Boosts Ethanol Titers and Yields from High Solids SSF with Low Enzyme Loadings.

Green Chem. 2016, 18 (6), 1581–1589. https://doi.org/10.1039/c5gc01977j.

(60) Barnes, W.; Anderson, C. Acetyl Bromide Soluble Lignin (ABSL) Assay for Total Lignin

Quantification from Plant Biomass. Bio-Protocol 2017, 7 (5), 1–11.

https://doi.org/10.21769/bioprotoc.2149.

(61) Foster, C. E.; Martin, T. M.; Pauly, M. Comprehensive Compositional Analysis of Plant

Cell Walls (Lignocellulosic Biomass) Part I: Lignin. J. Vis. Exp. 2010, No. 37, 5–8.

https://doi.org/10.3791/1745.

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
35
(62) Ci, S.; Wen, Z.; Qian, Y.; Mao, S.; Cui, S.; Chen, J. NiO-Microflower Formed by

Nanowire-Weaving Nanosheets with Interconnected Ni-Network Decoration as Supercapacitor

Electrode. Sci. Rep. 2015, 5 (1), 1–12. https://doi.org/10.1038/srep11919.

(63) Christien, F.; Ferchaud, E.; Nowakowski, P.; Allart, M. The Use of Electron Probe

MicroAnalysis to Determine the Thickness of Thin Films in Materials Science. X-Ray Spectrosc.

2012, No. February. https://doi.org/10.5772/28877.

(64) Subbotina, E.; Rukkijakan, T.; Marquez-Medina, M. D.; Yu, X.; Johnsson, M.; Samec, J.

S. M. Oxidative Cleavage of C–C Bonds in Lignin. Nat. Chem. 2021, 13 (11), 1118–1125.

https://doi.org/10.1038/s41557-021-00783-2.

(65) Anderson Guerra, L. A. L. and D. S. A. Isolation and Characterization of Lignins from

Eucalyptus Grandis Hill Ex Maiden and Eucalyptus Globulus Labill. by Enzymatic Mild

Acidolysis (EMAL). 2008, 62 (Issue 1), 24–30.

https://doi.org/https://doi.org/10.1515/HF.2008.004.

(66) Bozell, J. J.; O’Lenick, C. J.; Warwick, S. Biomass Fractionation for the Biorefinery:

Heteronuclear Multiple Quantum Coherence-Nuclear Magnetic Resonance Investigation of Lignin

Isolated from Solvent Fractionation of Switchgrass. J. Agric. Food Chem. 2011, 59 (17), 9232–

9242. https://doi.org/10.1021/jf201850b.

(67) Hwang, J.; Ejsmont, A.; Freund, R.; Goscianska, J.; Schmidt, B. V. K. J.; Wuttke, S.

Controlling the Morphology of Metal-Organic Frameworks and Porous Carbon Materials: Metal

Oxides as Primary Architecture-Directing Agents. Chem. Soc. Rev. 2020, 49 (11), 3348–3422.

https://doi.org/10.1039/c9cs00871c.

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
36
(68) Freund, R.; Lanza, A. E.; Canossa, S.; Gemmi, M.; Goscianska, J.; Cauda, V.; Oschatz,

M.; Wuttke, S. Understanding the Chemistry of Metal Oxide to Metal − Organic Framework

Reactions for Morphology Control. 2023. https://doi.org/10.1021/acs.chemmater.2c02946.

(69) N.N. GREENWOOD, A. EARNSHAW, B.-H. Cobalt, Rhodium and Iridium. Chem. Elem.

(Second Ed. 1987, 1, 1113–1143. https://doi.org/10.1016/B978-0-7506-3365-9.50032-8.

(70) Biesinger, M. C.; Payne, B. P.; Grosvenor, A. P.; Lau, L. W. M.; Gerson, A. R.; Smart, R.

S. C. Resolving Surface Chemical States in XPS Analysis of First Row Transition Metals, Oxides

and Hydroxides: Cr, Mn, Fe, Co and Ni. Appl. Surf. Sci. 2011, 257 (7), 2717–2730.

https://doi.org/10.1016/j.apsusc.2010.10.051.

(71) Padilha, C. E. de A.; Nogueira, C. da C.; Alencar, B. R. A.; de Abreu, Í. B. S.; Dutra, E.

D.; Ruiz, J. A. C.; Souza, D. F. de S.; dos Santos, E. S. Production and Application of Lignin-

Based Chemicals and Materials in the Cellulosic Ethanol Production: An Overview on Lignin

Closed-Loop Biorefinery Approaches. Waste and Biomass Valorization 2021, 12 (12), 6309–6337.

https://doi.org/10.1007/s12649-021-01455-5.

(72) Liu, C. G.; Xiao, Y.; Xia, X. X.; Zhao, X. Q.; Peng, L.; Srinophakun, P.; Bai, F. W.

Cellulosic Ethanol Production: Progress, Challenges and Strategies for Solutions. Biotechnol. Adv.

2019, 37 (3), 491–504. https://doi.org/10.1016/j.biotechadv.2019.03.002.

(73) Viikari, L.; Vehmaanperä, J.; Koivula, A. Lignocellulosic Ethanol: From Science to

Industry. Biomass and Bioenergy 2012, 46, 13–24.

https://doi.org/10.1016/j.biombioe.2012.05.008.

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
37
(74) Song, Q.; Wang, F.; Cai, J.; Wang, Y.; Zhang, J.; Yu, W.; Xu, J. Lignin Depolymerization

(LDP) in Alcohol over Nickel-Based Catalysts via a Fragmentation-Hydrogenolysis Process.

Energy Environ. Sci. 2013, 6 (3), 994–1007. https://doi.org/10.1039/c2ee23741e.

(75) Becker, J.; Wittmann, C. A Field of Dreams: Lignin Valorization into Chemicals,

Materials, Fuels, and Health-Care Products. Biotechnol. Adv. 2019, 37 (6), 107360.

https://doi.org/10.1016/j.biotechadv.2019.02.016.

(76) Zhao, Z. M.; Meng, X.; Scheidemantle, B.; Pu, Y.; Liu, Z. H.; Li, B. Z.; Wyman, C. E.;

Cai, C. M.; Ragauskas, A. J. Cosolvent Enhanced Lignocellulosic Fractionation Tailoring Lignin

Chemistry and Enhancing Lignin Bioconversion. Bioresour. Technol. 2022, 347 (September

2021), 126367. https://doi.org/10.1016/j.biortech.2021.126367.

(77) Meng, X.; Parikh, A.; Seemala, B.; Kumar, R.; Pu, Y.; Christopher, P.; Wyman, C. E.; Cai,

C. M.; Ragauskas, A. J. Chemical Transformations of Poplar Lignin during Cosolvent Enhanced

Lignocellulosic Fractionation Process. ACS Sustain. Chem. Eng. 2018, 6 (7), 8711–8718.

https://doi.org/10.1021/acssuschemeng.8b01028.

(78) Jones, C. J. D- and F-Block Chemistry. J. Chem. Educ. 2001, 78 (11), 1468.

https://doi.org/10.1021/ed078p1468.

(79) Farjana, S. H.; Huda, N.; Mahmud, M. A. P. Life Cycle Assessment of Cobalt Extraction

Process. J. Sustain. Min. 2019, 18 (3), 150–161. https://doi.org/10.1016/j.jsm.2019.03.002.

(80) Squadrone, S.; Burioli, E.; Monaco, G.; Koya, M. K.; Prearo, M.; Gennero, S.; Dominici,

A.; Abete, M. C. Human Exposure to Metals Due to Consumption of Fish from an Artificial Lake

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
38
Basin Close to an Active Mining Area in Katanga (D.R. Congo). Sci. Total Environ. 2016, 568,

679–684. https://doi.org/10.1016/j.scitotenv.2016.02.167.

(81) Mancini, L.; Eslava, N. A.; Traverso, M.; Mathieux, F. Assessing Impacts of Responsible

Sourcing Initiatives for Cobalt: Insights from a Case Study. Resour. Policy 2021, 71 (October

2020), 102015. https://doi.org/10.1016/j.resourpol.2021.102015.

(82) Traverso, M.; Valdivia, S.; Luthin, A.; Roche, L.; Arcese, G.; Neugebauer, S.; Petti, L.;

D’Eusanio, M.; Tragnoone, B. M.; Mankaa, R.; Hanafi, J.; Benoît Norris, C. Zagmani, A.

Methodological Sheets for Subcategories in Social Life Cycle Assessment (S-LCA). 2021, 150.

(83) Benoît-Norris, C.; Vickery-Niederman, G.; Valdivia, S.; Franze, J.; Traverso, M.; Ciroth,

A.; Mazijn, B. Introducing the UNEP/SETAC Methodological Sheets for Subcategories of Social

LCA. Int. J. Life Cycle Assess. 2011, 16 (7), 682–690. https://doi.org/10.1007/s11367-011-0301-

y.

(84) Catherine Benoît, UQAM/CIRAIG, and B. M. UNEP-SETAC: GUIDELINES FOR SCLA

of Product; Catherine Benoît, UQAM/CIRAIG, and Bernard Mazijn, G. U., Ed.; UNEP/SETAC

Life Cycle Initiative at UNEP, CIRAIG, FAQDD and the Belgium Federal Public Planning Service

Sustainable Development: Belgium, 2009.

https://doi.org/https://www.unep.org/resources/report/guidelines-social-life-cycle-assessment-

products.

TOC Graphic

https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
39
https://doi.org/10.26434/chemrxiv-2023-qdbkr-v3 ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
40

You might also like