Download as pdf or txt
Download as pdf or txt
You are on page 1of 25

Article

A New Method for Solid Acid Catalyst Evaluation for


Cellulose Hydrolysis
Maksim Tyufekchiev 1 , Jordan Finzel 1 , Ziyang Zhang 1 , Wenwen Yao 2 , Stephanie Sontgerath 1 ,
Christopher Skangos 1 , Pu Duan 3 , Klaus Schmidt-Rohr 3 and Michael T. Timko 1, *

1 Department of Chemical Engineering, Worcester Polytechnic Institute, 100 Institute Road,


Worcester, MA 01609, USA; maksimvt@wpi.edu (M.T.); jpfinzel@wpi.edu (J.F.); zzhang10@wpi.edu (Z.Z.);
sasontgerath@wpi.edu (S.S.); caskangos@wpi.edu (C.S.)
2 Department of Civil and Environmental Engineering, Worcester Polytechnic Institute, 100 Institute Road,
Worcester, MA 01609, USA; wyao@wpi.edu
3 Department of Chemistry, Brandeis University, 415 South Street, Waltham, MA 02453, USA;
pduan91@brandeis.edu (P.D.); srohr@brandeis.edu (K.S.-R.)
* Correspondence: mttimko@wpi.edu

Abstract: A systematic and structure-agnostic method for identifying heterogeneous activity of solid
acids for catalyzing cellulose hydrolysis is presented. The basis of the method is preparation of
a supernatant liquid by exposing the solid acid to reaction conditions and subsequent use of the
supernatant liquid as a cellulose hydrolysis catalyst to determine the effects of in situ generated
homogeneous acid species. The method was applied to representative solid acid catalysts, including
 polymer-based, carbonaceous, inorganic, and bifunctional materials. In all cases, supernatant liquids

produced from these catalysts exhibited catalytic activity for cellulose hydrolysis. Direct comparison
Citation: Tyufekchiev, M.; Finzel, J.;
of the activity of the solid acid catalysts and their supernatants could not provide unambiguous de-
Zhang, Z.; Yao, W.; Sontgerath, S.;
tection of heterogeneous catalysis. A reaction pathway kinetic model was used to evaluate potential
Skangos, C.; Duan, P.;
false-negative interpretation of the supernatant liquid test and to differentiate heterogeneous from
Schmidt-Rohr, K.; Timko, M.T. A New
homogeneous effects on cellulose hydrolysis. Lastly, differences in the supernatant liquids obtained
Method for Solid Acid Catalyst
Evaluation for Cellulose Hydrolysis.
in the presence and absence of cellulose were evaluated to understand possibility of false-positive
Sustain. Chem. 2021, 2, 645–669. interpretation, using structural evidence from the used catalysts to gain a fresh understanding of
https://doi.org/10.3390/ reactant–catalyst interactions. While many solid acid catalysts have been proposed for cellulose
suschem2040036 hydrolysis, to our knowledge, this is the first effort to attempt to differentiate the effects of heteroge-
neous and homogeneous activities. The resulting supernatant liquid method should be used in all
Academic Editor: Raffaele Cucciniello future attempts to design and develop solid acids for cellulose hydrolysis.

Received: 11 October 2021 Keywords: biomass; cellulose; cellulose hydrolysis; solid acid; solid acid catalyst
Accepted: 11 November 2021
Published: 15 November 2021

Publisher’s Note: MDPI stays neutral 1. Introduction


with regard to jurisdictional claims in
As a major constituent of lignocellulosic biomass, cellulose is the planet’s most abun-
published maps and institutional affil-
iations.
dant source of renewable carbon [1]. Efficient conversion of cellulose to its constituent
monomer, glucose, or other small molecules is the gateway to production of cost-effective
second-generation biofuels and biochemicals [2,3]. Unfortunately, cellulose is resistant to
depolymerization [4], and converting it into soluble products requires energy-intensive
pretreatments [5–10], harsh conditions [11–16], excess enzymes [17–19], or expensive sol-
Copyright: © 2021 by the authors.
vents [20,21] that contribute disproportionately to biofuel production costs [2,22].
Licensee MDPI, Basel, Switzerland.
Dilute acid pretreatment followed by enzyme hydrolysis is the most commercially
This article is an open access article
advanced method of converting cellulose present in biomass into glucose [22]. Dilute acid
distributed under the terms and
conditions of the Creative Commons
pretreatment consists of hydrolyzing and removing the hemicellulose that protects the
Attribution (CC BY) license (https://
underlying cellulose, while enzymes depolymerize the remaining cellulose to produce
creativecommons.org/licenses/by/
glucose [23–26]. Unfortunately, sequential dilute acid pretreatment and enzyme hydrolysis
4.0/).

Sustain. Chem. 2021, 2, 645–669. https://doi.org/10.3390/suschem2040036 https://www.mdpi.com/journal/suschem


Sustain. Chem. 2021, 2 646

of cellulose for production of ethanol remain prohibitively expensive compared with


petroleum fuels [22].
In the past several years, solid acids have been proposed as alternatives to homoge-
neous acids and enzymes for catalyzing cellulose hydrolysis [27,28]. Unlike homogeneous
acids and enzymes, solid acids can theoretically be recycled with potential for correspond-
ing reductions in cellulose conversion costs. However, in typical reaction conditions
(>120 ◦ C, >10 h), both the catalyst and the cellulose substrate are solids, and the mecha-
nism of solid–solid catalytic action has not been elucidated, hindering rational catalyst
development efforts [29].
Traditionally, catalyst design emphasizes development of structure–activity relation-
ships [30], and this approach has also been adopted for design of solid acid catalysts for
cellulose depolymerization [27]. Structure–activity data have then been combined with
carbohydrate adsorption data to propose a mechanism that describes depolymerization as
a two-step process involving solid–solid binding, followed by catalytic hydrolysis [31–35].
Bifunctional catalysts, possessing both carbohydrate-binding and acid groups, have been
suggested as especially effective for cellulose hydrolysis [31,35–37]. While the bifunctional
catalyst hypothesis may have a precedent [38,39], it suffers from the need for extensive
structural catalyst characterization, ideally under reaction conditions or at least after
use [40–43].
On the other hand, the field of solid acid-catalyzed cellulose hydrolysis may be im-
pacted by confirmation bias where negative results are not published, thereby erroneously
promulgating the bifunctional binding hypothesis [44]. Interestingly, some studies indicate
that some of the most promising bifunctional solid acids release homogeneous acid species
under cellulose hydrolysis conditions [40,41,43]. The result is parallel hydrolysis by solid
acids (as desired) and homogeneous acids (which is undesired). Distinguishing between
homogeneous or heterogeneous reaction promotion can be difficult, especially since homo-
geneous acids released during reaction have the potential to completely overwhelm the
sought-after effect of the solid acid.
While the impact of homogeneous acid on cellulose hydrolysis has been acknowl-
edged [40,45], the prevalence of the phenomenon for different catalyst types is not known,
and no approach has yet been established to quantify the effect or even identify when it is
present [46]. Despite the evidence of homogeneous acid hydrolysis [40], studies continue
to propose new solid materials for catalyzing cellulose depolymerization, usually with
incomplete characterization of the effects of homogeneous acids released under reaction
conditions [42,47–50]. Accordingly, this raises the important question: how can homoge-
neous contributions to cellulose hydrolysis activity be differentiated from heterogeneous
ones? A method for differentiating homogeneous from heterogeneous activity could be ap-
plied universally to new catalysts during the early stages of development, thereby helping
to allocate resources to only the most promising materials. Ideally, differentiation should
not depend on structural analysis, which can be costly and ultimately inconclusive given
that even the release of what could be considered trace species can result in catalytic effects
without impacting catalyst structure [45], instead consisting of a generalized approach that
any lab can perform.
Accordingly, we aimed to develop a structure-agnostic approach that can be used to
(1) identify catalysts with cellulose hydrolysis activity and minimal or negligible release of
homogeneous acids during reaction, (2) account for the hydrolysis activity of the homoge-
neous acids when they are present, and (3) be applied regardless of reaction conditions.
Once the effects of the homogeneous acid are known, then the remaining hydrolysis activ-
ity, if any, can be attributed to the solid acid. Unlike structure–property methodologies,
a structure-agnostic approach does not rely on exhaustive structural characterization and
can, therefore, avoid false structure–activity correlations made by characterizing a catalyst
incompletely or under conditions other than those relevant to reaction [35,40]. A structure-
agnostic procedure can be used as a screening tool for identification of promising materials.
Sustain. Chem. 2021, 2 647

Only after a promising candidate with heterogeneous activity has been identified is its
detailed characterization initiated, thus saving time and effort.
A structure agnostic catalyst evaluation methodology must satisfy several criteria
to be useful to the community. First, and most obviously, it needs to reliably separate
the effects of the homogeneous and solid acids on the observed cellulose hydrolysis and
require a minimum of additional experiments to do so. Second, the methodology should
use techniques that are accessible to the community, contrasting with characterization
techniques, many of which are prohibitively expensive or time-consuming and not available
in every lab. Conversion (or yield) measurements are the obvious choice, since all labs
working in this area will have the capability of measuring cellulose conversion and/or
glucose yields. Third, the method should be broadly applicable to all catalyst types and
reaction conditions.
The structure-agnostic approach for differentiating homogeneous and heterogeneous
activity must, therefore, be compatible with existing experimental methods. Different
methodologies have been applied for the catalytic evaluation of solid acids, as shown in
Scheme 1. The most widely used approach (Scheme 1a) is mixing cellulose and the solid
catalysts in a batch reactor, heating to reaction temperature, and carrying out the reaction
for a specified period. Clearly, this method cannot distinguish between heterogeneous and
homogeneous activity since both types of catalysts will be present in the reaction mixture,
as the solid catalyst degrades to generate soluble products.
Generation of soluble catalytic species during cellulose hydrolysis can be suppressed
by first exposing the solid catalyst to hydrothermal conditions more severe than those used
for the reaction itself [33,51]. This strategy is shown as Scheme 1b. The assumption is
that the harsh pretreatment will remove thermally and/or hydrothermally labile catalyst
components from the solid material [33,51], and the resulting solid is more hydrothermally
stable than the original. In this respect, the pretreatment can be viewed as a final step in
catalyst synthesis, with its objective being to improve catalyst stability. The disadvantage
to this method is that the pretreated catalyst may still release soluble acids under reaction
conditions, meaning that the potential contribution from homogeneous species is not
eliminated, and that structural characterization is still required.
A permutation of the catalyst pretreatment approach is repetitive recovery and reuse
of the catalyst [49,50,52,53]. Each use may remove a fraction of the potential soluble acids,
leaving most of them on the catalyst, resulting in a slow decrease in catalyst activity.
The catalyst must still be characterized before and after use, making the approach no
more definitive than simply using the catalyst a single time and characterizing it. Both
the hydrothermal pretreatment and catalyst reuse approach are prone to “false positives”,
i.e., falsely attributing meaningful heterogeneous activity to materials that do not possess
it, due to the unjustified assumption that the effects of homogeneous species have been
suppressed or eliminated.
A third option is shown in Scheme 1c [40,45]. Here, the focus shifts from characterizing
the catalyst itself to understanding the supernatant liquid obtained after hydrolysis. First,
the apparent activity of the solid catalyst is determined by measuring cellulose conversion
and/or soluble product yield as usual under a set of predetermined reaction conditions.
In a separate set of experiments, the fresh solid acid is treated under the same hydrothermal
conditions used for hydrolysis, now in the absence of cellulose. The supernatant liquid
from this test is then recovered and can be analyzed for putative catalytic species (e.g., H+ )
before it is used as the liquid phase for cellulose hydrolysis under the same conditions as
before, this time in the absence of solid acid catalyst. Cellulose conversion and/or soluble
product yields are then measured after the supernatant liquid test for comparison with
results obtained using the solid acid.
Sustain. Chem.
Sustain. Chem. 2021,
2021, 22, FOR PEER REVIEW 6484

Scheme 1. Experimental approaches for analyzing the apparent cellulose hydrolysis activity of the solid acid catalyst.
Scheme 1. Experimental approaches for analyzing the apparent cellulose hydrolysis activity of the solid acid catalyst. In
In (a), a typical activity test consists of mixing cellulose, solid acid catalyst, and water, after which the temperature is
(a), a typical activity test consists of mixing cellulose, solid acid catalyst, and water, after which the temperature is elevated
elevated to carry
to carry out out hydrolysis.
hydrolysis. In (b), theInsolid
(b), acid
the solid acidiscatalyst
catalyst is pretreated
pretreated at hydrothermal
at hydrothermal conditionsconditions
that exceedthat
the exceed
severitythe
of
severity of those used for cellulose hydrolysis; after pretreatment, the catalyst is separated and used for cellulose
those used for cellulose hydrolysis; after pretreatment, the catalyst is separated and used for cellulose hydrolysis test. In hydrolysis
test. In (c),
(c), the the catalyst
catalyst is treated
is treated at the at the same
same hydrothermal
hydrothermal conditions
conditions used used for hydrolysis,
for hydrolysis, butthe
but in in absence
the absence of cellulose;
of cellulose; the
supernatant
the supernatantis separated
is separatedand and
usedused
for cellulose hydrolysis
for cellulose to determine
hydrolysis the activity
to determine of homogeneous
the activity of homogeneousacid species gener-
acid species
ated. Red color
generated. of theofliquid
Red color indicates
the liquid potential
indicates presence
potential of homogeneous
presence of homogeneouscatalysts generated
catalysts by degradation
generated of theofsolid
by degradation the
acid, acid,
solid whilewhile
blue is pure
blue water.
is pure water.

In this work, we evaluate the use of the supernatant liquid obtained from hydrother-
treatment of
mal treatment of solid
solid acid
acidcatalysts
catalystsas asthe
thebasis
basisofofaastructure-agnostic
structure-agnosticmethod
methodfor foridenti-
iden-
tifyingheterogeneous
fying heterogeneous cellulose
cellulose hydrolysis
hydrolysis activity.
activity. A series
A series of representative
of representative catalysts
catalysts was
selected, and their
was selected, activity
and their for cellulose
activity hydrolysis
for cellulose was was
hydrolysis measured at standard
measured at standardconditions
condi-
chosen to achieve
tions chosen approximately
to achieve 5–20%5–20%
approximately product yield. Solid
product yield.acids
Solidwere
acidsplaced in liquid
were placed in
water under the same reaction conditions as used for hydrolysis, but in
liquid water under the same reaction conditions as used for hydrolysis, but in the absence the absence of
cellulose; thethe
of cellulose; supernatant
supernatant liquids from
liquids these
from reactions
these werewere
reactions recovered, filtered
recovered, to remove
filtered to re-
catalyst particles,
move catalyst and then
particles, andused
thenas the as
used reaction medium
the reaction for hydrolysis
medium of a new
for hydrolysis of abatch
new
of cellulose
batch at the same
of cellulose at theconditions as the other
same conditions as thetests.
otherAtests.
reaction engineering
A reaction time study
engineering time
combined with modeling
study combined of the homogeneous
with modeling acid activity
of the homogeneous acid was performed
activity for a subset
was performed forofa
these catalysts as an approach to determine whether “false negative” interpretations
subset of these catalysts as an approach to determine whether “false negative” interpreta- can be
identified. Lastly, the effect of reactant–catalyst interactions on the release of homogeneous
tions can be identified. Lastly, the effect of reactant–catalyst interactions on the release of
Sustain. Chem. 2021, 2 649

catalysts was examined for representative catalysts. The results from this study can be
used by the community in future work to develop solid acid catalysts with heterogeneous
activity for cellulose hydrolysis and address the culture of reporting potentially interesting
but possible misinterpreted findings obtained from new materials.

2. Materials and Methods


2.1. Materials
Avicel PH-101 cellulose, Amberlyst-15, H2 SO4 (98%), 0.1 M HCl, 0.1 M NaOH, cel-
lobiose, glucose, xylose, hydroxymethyl furfural (HMF), levulinic acid, and formic acid
were purchased from Sigma-Aldrich (Burlington, MA, USA). Sulfated zirconia, Norit,
and HZSM-5 (Si/Al = 38) were purchased from MEL Chemicals (Flemington, NJ, USA),
Cabot Corporation, and ACS Material (Pasadena, CA, USA), respectively. 13 C-enriched
glucose was purchased from Sigma-Aldrich.

2.2. Catalysts Preparation


2.2.1. Sulfonated Humins (SH)
Xylose (20 g) and 98% H2 SO4 (7 mL) were mixed with deionized water (17.7 MΩ·cm)
to create a 100 mL solution with a 1:1 molar ratio of sugar to acid. This was then stirred at
room temperature for 30 min. The resulting solution was then placed inside a 160 mL Teflon
vessel, which was deposited into an aluminum autoclave for 24 h at 120 ◦ C. The resulting
mixture was allowed to cool for 12 h at room temperature and then washed with a 300 mL
equimolar mixture of DI water and ethanol. The solid char phase was separated from the
aqueous phase via vacuum filtration and dried for 24 h at 65 ◦ C.

2.2.2. Sulfonated Activated Carbon (SAC)


Norit activated carbon was sulfonated in a similar fashion to a procedure used by
Foo and Sievers et al. [33] Briefly, 10 g of Norit SX-1 was soaked in 500 mL of deionized
water overnight, after which it was filtered and dried at 65 ◦ C overnight. The activated
carbon was then mixed with 78 mL of deionized water and 54 g of 98% H2 SO4 and put
in an autoclave. The mixture was placed in a preheated oven at 200 ◦ C and allowed to
react for 24 h, after which the autoclave was cooled in an ice bath. Following the treatment,
the solids were filtered and washed twice with 1800 mL of deionized water for 24 h to
remove any residual H2 SO4 , after which the solid material was filtered again and dried
at 65 ◦ C.

2.2.3. Sulfated Zirconia (SZ)


Sulfuric acid-doped zirconia oxide (MEL Chemicals Inc, Flemington, NJ, USA) was
activated by placing the material in a preheated furnace at 550 ◦ C for 16 h.

2.2.4. HZSM-5
HZSM-5 with an Si/Al ratio of 38 (ACS Materials) was calcined at 550 ◦ C for 16 h.

2.2.5. CMP-SO3 H-0.3


The bifunctional polymer solid acid bearing chloromethyl and sulfonic acid groups
was prepared according to the procedure reported by Zuo et al. and reproduced by
Tyufekchiev et al. [35,40].

2.3. Catalyst Wash


The catalysts were washed repeatedly with deionized water (17.8 MΩ·cm) to remove
any soluble acidic spedcies present on the catalyst. Briefly, 1.5 g of catalyst was washed
with 50 mL of water in a centrifuge tube. The pH of the washout and the suspension was
measured until the value stopped changing with further wash. After the wash, the catalysts
were dried at 65 ◦ C overnight.
Sustain. Chem. 2021, 2 650

2.4. Solid-State Titration


Solid-state titration of the catalysts was carried out according to Boehm’s proce-
dure [54]. Briefly, predetermined amounts of 0.1 M NaOH standard solution and solid acid
catalyst were mixed. The mixture was shaken and allowed to ion exchange for 24 h. Next,
an aliquot was taken and acidified with a predetermined volume of standard 0.1 M HCl.
The acidified solution was then titrated with 0.05 M NaOH, while the pH was monitored
with a digital pH meter (Denver Instruments, model 225, Bohemia, NY, USA). The number
of acid sites in the solid catalyst was then calculated.

2.5. Hydrothermal Treatment for Supernatant Generation


Solid acid catalyst (0.2 g) was mixed with deionized water (2 mL) and treated in
hydrothermal conditions at 150 ◦ C for 15 h (10 h for CMP-SO3 H-0.3, Amberlyst-15,
and sulfated zirconia). After the reaction, the pH of the liquid was measured; the liquid and
the solid were centrifuged, and the liquid was extracted with a syringe. The supernatant
liquid was then analyzed with ion chromatography (IC) to characterize the leached species
and stored for cellulose hydrolysis tests.

2.6. Cellulose Hydrolysis


Cellulose was hydrolyzed to determine the activity of the solid acid catalyst and the
leached homogeneous acid. Briefly, solid acid catalyst (0.2 g) was mixed with cellulose
(0.1 g) and deionized water (2 mL) in a 15 mL glass reactor vial and sealed with a Teflon
cap and Viton O-ring. The reactor vial was submerged in a heated oil bath, and the mixture
was left to react for a pre-determined duration (10 h for CMP-SO3 H-0.3, Amberlyst-15,
and sulfated zirconia; 15 h for all other catalysts) to achieve 5–20% soluble product yield.
The reaction temperature in the vial was accurately measured with a modified screw-top
cap equipped with an Omega K-type thermocouple. Temperature measurements made
within reaction vessels placed in multiple locations within the thermal batch indicated that
the reaction temperature was 150 ± 2.5 ◦ C. After the reaction, the vial was centrifuged,
and the liquid was extracted with a syringe. The soluble products of the hydrolysis were
analyzed with HPLC.
In a similar fashion, the activity of the supernatant was determined by mixing 2 mL of
the liquid from hydrothermally treated solid catalyst with cellulose (0.1 g) and reacted at
the same conditions as above. The soluble products were quantified with high-performance
liquid chromatography (HPLC) (Agilent Technologies, Santa Clara, CA, USA).

2.7. Hydrolysate Analysis


Products of cellulose hydrolysis were quantified with high-performance liquid chro-
matography (HPLC) using an Agilent 1200 Series instrument equipped with a refractive
index detector (RID) and a diode array detector (DAD) for quantifying carbohydrates
and furanic compounds, respectively. A Rezex ROA-Organic acid column (Phenomenex)
maintained at 35 ◦ C was used for separation, and the mobile phase was deionized water at
0.6 mL/min. The RID detector operated at 35 ◦ C and the DAD detection wavelength was
set to 284 nm. Calibration curves for cellobiose, glucose, HMF, levulinic acid, and formic
acid were prepared by determining detector response to standardized solutions with con-
centrations of 0.25, 0.5, 0.75, 1, 1.25, 1.5, 1.75, 2, and 2.5 g/L. Yields of products were
m ×M
calculated on a theoretical molar basis as follows: mpc ×Mgu
p
× 100%. where mc is the mass of
cellulose, mp is the mass of the product determined by HPLC, Mgu is the molecular weight
of glucose unit in cellulose, and Mp is the molecular weight of the product.
Sustain. Chem. 2021, 2 651

2.8. Ion Analysis


Anion concentrations in the supernatant were quantified with a Dionex ICS-2100 Ion
Chromatograph equipped with an AERS 500 anion electrolytically regenerated suppressor
and a DS6 heated conductivity cell (ThermoFisher Scientific, Waltham, MA, USA). Anions
were separated on a AS153 250 mm column equipped with an IonPac AG 2 × 50 mm
guard. The mobile phase was 38.00 mM KOH, and its flow rate was held at 0.25 mL/min.
Separation was carried at a column temperature of 30 ◦ C, and detection was performed at
a cell temperature of 35 ◦ C.
The hydronium ion content was measured to within ±0.01 pH units using a pH meter
(Denver Instruments, model 225) equipped with a glass, Ag/AgCl reference, 0–14 pH
probe (Symphony pH Probes), calibrated with buffers of pH of 4 and 7.

2.9. Kinetic Modeling


A kinetic model consisting of ordinary differential equations describing cellulose
hydrolysis to glucose was developed to relate the time evolution of solubilized catalysts
with products of cellulose hydrolysis, in the presence and absence of solid acid. On review
of the literature [55–59], the kinetic models proposed by Saeman were selected for the
framework as the temperature, acid concentration ranges, and cellulose particle size used
in developing Saeman’s kinetic analysis were in the same range as those examined in
the current study. The cellulose hydrolysis reaction was modeled using a first-order
d[C ]
rate law: dt = −k × [C ], where the expression for the reaction rate constant and its
 
n
dependence on acid concentration is defined as k = A × [ H + ] × exp 2.303−×ERa ×T , with
A = 1.73 × 1019 min−1 ·M−1.34 , n = 1.34, and Ea = 179.5 kJ·mol−1 [55].
The time evolution of homogeneous species was captured using a simple first-order
model that accounted for the pH measured in a separate time study. Experimental data
were fit to determine the average value and uncertainty of the degradation rate constant
that led to formation of homogeneous catalyst as H+ . The derived H+ concentration as
a function of time was substituted into the expression for the cellulose hydrolysis rate
constant in Saeman’s model.
MATLAB®® was used for solving the differential equation at variable temperature
and solid acid catalyst degradation constant. The dependence on temperature, to which the
model is especially sensitive, and the deviation for the catalyst degradation rate constant
were simulated using a random number generator function. To model uncertainty, Monte
Carlo simulations of the reaction system were performed, using reaction temperature
(±2.5 ◦ C) as the uncertain variable. Performing Monte Carlo simulations of reaction
temperature captures uncertainty arising from natural variability in the thermal control
system, as well as uncertainties in reaction rate parameters. For comparison, a kinetic study
of cellulose hydrolysis using selected solid acid catalysts was carried out. The quantification
of soluble products and soluble carbon balance was as described previously [40].

2.10. Interactions between Solid Acid Catalyst and Soluble Substrates


Glucose hydrolysis was studied as a model to probe interactions between the soluble
products generated from cellulose hydrolysis. Briefly, the tests were identical to cellulose
hydrolysis; the supernatant activity test was carried out using Amberlyst-15, but for a
period of 2 h. 13 C glucose was reacted with CMP-SO3 H-0.3 for 10 h; the solid material
was then recovered and analyzed with solid-state nuclear magnetic resonance to study
the species deposited on the solid catalyst and potential chemical bonding between those
species and the solid acid.
Sustain. Chem. 2021, 2 652

2.11. Solid-State Nuclear Magnetic Resonance


The structure of the CMP-SO3 H-0.3 catalyst was analyzed using solid-state nuclear
magnetic resonance (NMR) spectroscopy after mixing with 13 C-enriched glucose at the
same conditions used for cellulose reaction. Experiments were performed on a Bruker
Avance 400 spectrometer (Billerica, MA, USA) at a 100 MHz 13 C resonance frequency with
high-power 1 H decoupling, with magic-angle spinning (MAS) of 4 mm zirconia rotors in a
double-resonance probe head at ambient temperature. 13 C chemical shifts were calibrated
to TMS, using 1-13 Cglycine (α-form) at 176.49 ppm as a secondary reference. The composite-
pulse multiCP pulse sequence [60] was used at 14 kHz MAS to obtain quantitative 13 C-NMR
spectra, with a 4 s recycle delay, 10 1.1 ms cross-polarization periods, and a final 0.55 ms
CP time, each separated by a 1 H repolarization time of 1.5 s. Corresponding spectra of
non-protonated C and mobile segments were obtained after 68 µs recoupled 1 H–13 C dipolar
dephasing before detection [61]. To evaluate connectivity between protonated and non-
protonated carbons in coked CMP-SO3 H-0.3, two-dimensional exchange with protonated
and non-protonated spectral editing (EXPANSE) NMR spectra [62] were recorded with a
mixing time of 10 ms. Since small residual diagonal peaks of arenes do not interfere with
detection of most of the cross-peaks, the dipolar dephasing difference was not recorded,
which maximized the signal-to-noise ratio. Standard 4.2 µs 1 H and 13 C 90◦ pulses were
used in the experiments described above.

3. Results and Discussion


The objective of this work was the development and demonstration of a structure-
agnostic method for identifying promising materials with heterogeneous activity for cellu-
lose hydrolysis. In the first section, single-point solid activity and supernatant tests were
performed for six materials, as shown in Scheme 1c. The second section demonstrates
methods for evaluating false negatives, i.e., situations in which the supernatant test falsely
indicates the absence of heterogeneous activity. The third section examines potential for
false positives and introduces methods for evaluating reactant–catalyst interactions, which
in the cases considered here are likely to give rise to false positives, i.e., incorrect iden-
tification of heterogeneous activity when none is present. The fourth and final section
recommends a strategy for identifying solid acids with sufficient heterogeneous activity
for cellulose hydrolysis to justify detailed studies.

3.1. Single-Point Measurements of Solid Acid and Supernatant Cellulose Hydrolysis Activity
For testing the new approach, we selected representative materials from four structural
classes that have been commonly reported for cellulose hydrolysis: (1) Amberlyst-15
(AMB-15) as a polymer catalyst, (2) sulfonated humin (SH) and sulfonated activated
carbon (SAC) as carbonaceous catalysts, (3) ZSM-5 in its protonated form and sulfated
zirconia (SZ) as inorganic catalysts, and (4) partially sulfonated chloromethyl polystyrene
(CMP-SO3 H-0.3) as a bifunctional catalyst [33,35,40,63–69]. Evaluation of these catalyst
categories is sufficient for the current purpose of developing a structure-agnostic method
for distinguishing heterogeneous from homogeneous activity. Testing several catalyst
categories demonstrates the use of the approach for different types of catalysts, establishing
its credibility for future use by others. Simplified structures of the catalyst are provided in
Figure 1. Synthesis and modification procedures are provided in Section 2, as appropriate.
In all cases, catalysts were washed extensively prior to use to remove water-soluble acids
from their surfaces.
Sustain.Sustain.
Chem.Chem.
2021, 2021,
2, FOR2 PEER REVIEW 653 9

Figure 1. Simplified structures of the solid acid catalysts used in the current study [33,35,70–73].
Figure 1. Simplified structures of the solid acid catalysts used in the current study [33,35,70–73].
A structure-agnostic approach for discerning heterogeneous catalytic activity for cellu-
loseA hydrolysis
structure-agnostic
should notapproach for discerning
require catalyst heterogeneous
characterization. catalytic
Nonetheless, activity
to establish for cel-
that
lulose hydrolysis
the selected should
materials hadnot require catalyst
the potential characterization.
to be solid Nonetheless,
acid catalysts before proceeding to toeval-
establish
uating
that their activity,
the selected we first
materials hadquantified their acid
the potential to besite densities
solid using Boehm
acid catalysts titration
before [54].
proceeding to
Data are provided in Table S1 (see Supplementary Materials). In all cases, the
evaluating their activity, we first quantified their acid site densities using Boehm titration materials
selected for this study possess titratable acid sites, with densities greater than 800 µmol·g−1 .
[54]. Data are provided in Table S1 (see Supplementary Materials). In all cases, the mate-
Because of their external surface area and partially microporous pore structures, a fraction
rials selected for this study possess titratable acid sites, with densities greater than 800
of these acid groups will be accessible to cellulose, making the selected materials plausible
μmol· g−1. Because
catalysts of their
for cellulose external surface area and partially microporous pore structures,
hydrolysis.
a fraction
Developing a method forwill
of these acid groups be accessible
discerning to cellulose,
and quantifying making the
heterogeneous selected
catalytic materials
effects
plausible
benefitscatalysts for cellulose
from selection hydrolysis.
of a standard condition that maintains modest conversion to min-
Developing
imize secondarya reactions
method forthatdiscerning andtoquantifying
can contribute formation ofheterogeneous
insoluble species catalytic
that caneffects
complicate data interpretation. Accordingly, the apparent cellulose hydrolysis
benefits from selection of a standard condition that maintains modest conversion to activity of min-
imize secondary reactions that can contribute to formation of insoluble species that can
complicate data interpretation. Accordingly, the apparent cellulose hydrolysis activity of
the selected representative solid acids was determined at a standard reaction temperature
(150 ± 2.5 °C) for reaction times of either 10 or 15 h, with the reaction time selected to
Sustain. Chem. 2021, 2, FOR PEER REVIEW

Sustain. Chem. 2021, 2 654

catalysts shown in Figure 1 were then tested for cellulose hydrolysis at the standa
conditions. Table S2 provides
the selected representative detailed
solid acids product at
was determined distribution data. In
a standard reaction all cases, the pr
temperature
(150 ± 2.5 ◦ C) for reaction times of either 10 or 15 h, with the reaction time selected to
were glucose, levulinic acid, and formic acid, resulting in overall soluble product
maintain product yields obtained for the different catalysts at comparable values. The cat-
ranging from 2% to 25%, consistent with cellulose hydrolysis and subsequent dehyd
alysts shown in Figure 1 were then tested for cellulose hydrolysis at the standardized
ofconditions.
glucose [56].
Table S2 provides detailed product distribution data. In all cases, the products
Measured product
were glucose, levulinic acid,yields shown
and formic inresulting
acid, Table S2inwere used
overall to estimate
soluble averaged p
product yields
ranging from
formation 2% toas
rates 25%,
an consistent
indication withofcellulose
apparenthydrolysis and activity.
catalytic subsequentFigure
dehydration
2 summari
of glucose [56].
averaged product formation rates for each of the six catalysts considered in this
Measured product yields shown in Table S2 were used to estimate averaged product
(gray bars).
formation The
rates as formation
an indication rates were calculated
of apparent by summing
catalytic activity. the mass of
Figure 2 summarizes theall prod
the reaction
averaged media
product as a proxy
formation estimation
rates for of catalysts
each of the six cellulose conversion.
considered in thisThe
studyapparent
(gray for
rates observed when the two polymer catalysts (CMP-SO3H-0.3 and Amberlyst 15
bars). The formation rates were calculated by summing the mass of all products in the
reactionthan
greater mediathose
as a proxy estimation
of the inorganicof cellulose conversion. The
or carbonaceous apparentProduct
catalysts. formationformation
rates ra
observed when the two polymer catalysts (CMP-SO3 H-0.3 and Amberlyst 15) were greater
served for the bifunctional polymer were greater than any of the other materials, a
than those of the inorganic or carbonaceous catalysts. Product formation rates observed for
that could casually
the bifunctional polymerbe were
misattributed
greater thantoanyenhanced catalyst–cellulose
of the other interaction
materials, a result that could due
presence
casually beofmisattributed
chloride groups. Thecatalyst–cellulose
to enhanced purpose of structure-agnostic
interaction due to thedifferentiation
presence of of
chloride groups. The purpose of structure-agnostic differentiation of homogeneous
geneous and heterogeneous promotion is to test this and other similar attributions and
heterogeneous promotion is to test this and other similar attributions.

2.5
Polymeric
Average Rate of Hydrolysis

2.0
-3
x10

1.5
gproducts
h

1.0
Inorganic
Carbonaceous
0.5

0.0
5

5
.3

SZ

SH
C
1

SA
-0

B-

ZS
3H

AM

H
SO
P-
M
C

Figure 2. Comparison of average product formation rate after cellulose hydrolysis with fresh solid
acid catalyst (gray bars) and supernatant (red bars). The dotted line represents the activity of
water-only in the absence of
of catalyst ◦ C (±2.5 ◦ C) reaction
Figure 2. Comparison averageat product
the same conditions.
formationConditions:
rate after 150
cellulose hydrolysis with fre
temperature
acid catalystand either
(gray 10 orand
bars) 15 h supernatant
reaction time, 2(red
mL ofbars).
water,The
0.1 gdotted
of cellulose,
lineand 0.2 g of catalyst.
represents the activity o
only in the absence of catalyst at the same conditions. Conditions: 150 °C (±2.5 °C) reaction
ature and either 10 or 15 h reaction time, 2 mL of water, 0.1 g of cellulose, and 0.2 g of cataly

Product formation rates on their own are easily misinterpreted, and one way
Sustain. Chem. 2021, 2 655

Product formation rates on their own are easily misinterpreted, and one way to reduce
the likelihood of misinterpretation is to characterize the catalyst after use. However, only
use of appropriate, quantitative characterization techniques can prevent misinterpretation.
The tests that are appropriate for a given material depend on its specific chemistry, adding
complexity and ambiguity to the process of post-reaction characterization. Instead, we ad-
vocate performing additional control tests, as shown in Scheme 1. Release of even trace
species can influence observed product formation rates, meaning that characterization
must be sufficiently sensitive to arrive at firm conclusions.
Because cellulose hydrolysis can occur in the absence of added catalyst, soluble prod-
uct formation rates were next measured in the absence of any catalyst, but in the same
reaction conditions under which catalyzed formation rates were measured. The resulting
noncatalytic product formation rate is shown as the horizontal line in Figure 2. In nearly
all cases, the yields (see Table S2) and product formation rates (Figure 2) observed in the
presence of catalyst were greater than under noncatalytic conditions, with the only excep-
tion being H-ZSM-5. Accordingly, H-ZSM-5 can be eliminated from further considerations
as a solid acid catalyst for cellulose hydrolysis, at least under the conditions considered
here, as its activity is no greater than what could be expected without addition of catalyst.
The likely explanation is that H-ZSM-5 has insufficient acid sites on its external surface
to promote cellulose hydrolysis, and that it is sufficiently stable that it does not release
soluble acids.
In addition to promoting greater product formation rates than observed in the absence
of a catalyst, the activity of a solid catalyst toward cellulose hydrolysis must exceed that
observed in the supernatant liquid obtained by treating the catalyst in hot liquid water in
the absence of cellulose, as shown in Scheme 1c. Accordingly, Figure 2 includes product
formation rates observed when the solid acid catalysts were replaced with their supernatant
liquids (red bars) for direct comparison with rates observed in the presence of the solid
catalyst itself; detailed soluble products yields are presented in Table S2. In all cases,
the soluble product formation rates and yields obtained using the supernatant liquid were
equal to (within the uncertainty) or greater than the yield obtained using the corresponding
heterogeneous catalyst. This is a clear indication that all of the materials tested here release
homogeneous catalyst during reaction.
The supernatant obtained from CMP-SO3 H-0.3 is especially active, consistent with
previously observed formation of HCl (aq) when this catalyst is exposed to hydrolysis
conditions [40]. In comparison, the activity of AMB-15 is comparable to its supernatant
(within uncertainty). Unlike CMP-SO3 H-0.3, the activity of the AMB-15 supernatant is
likely a consequence of release of sulfuric acid during reaction [74].
Figure 2 clearly implies that the selected materials either completely lack activity for
cellulose hydrolysis (HZSM-5) or that any activity observed for them included contributions
from soluble species formed during hydrothermal degradation of the solid. It does not,
however, prove that all of the catalysts lack heterogeneous activity. To elucidate further,
we investigated the acid characteristics of the catalysts themselves and the properties of
the supernatant liquids.
To evaluate effects of homogeneous acids, we quantified acid leaching resulting from
hydrothermal treatment. After exposure to hydrothermal conditions (again at 150 ± 2.5 ◦ C
and for either 10 or 15 h) in the absence of cellulose, the pH of the resulting suspension
decreased by several units relative to that observed for a simple water washing at ambient
conditions (gray bars in Figure 3a) for every catalyst, as shown as the red bars in Figure 3a.
Accordingly, Figure 3a clearly establishes that hydrothermal treatment releases acids that
were stable at room temperature. These soluble acid species can then contribute to the
activity observed when the solid materials are used for cellulose hydrolysis, as implied
by comparison of heterogeneous catalyst and supernatant liquid product formation rates
shown in Figure 2.
ustain. Chem. 2021, 2, FOR PEER REVIEW
Sustain. Chem. 2021, 2 656

7 0.20
a) b)
Polymeric
6 Polymeric
*
Inorganic 0.15
5

Ion concentration (M)


Carbonaceous

4
pH

0.10
3

Inorganic
2
* 0.05
* * Carbonaceous
1 *
*
0 0.00
.3 15 -5 .3 15 SZ C
SZ C SH

SH
-5
-0 B- SA -0 B- SA

M
M H
H ZS
ZS
O 3
AM H SO
3
AM
S
P- H
P- M
M C
C
Figure 3. Hydrothermal stability of catalysts and leached acid characterization. In (a), gray bars indicate pH of catalyst
suspension in water after wash and red bars indicate pH after treatment at hydrothermal conditions (0.2 g of catalyst,
mL of water, 150 ◦stability
Figure 3.2 Hydrothermal C, 15 h). Inof(b), red bars represent
catalysts H+ , white
and leached acidbars represent Cl− , and
characterization. Ingray
(a),bars
grayrepresent HSO4 − ions.
bars indicate pH of catalyst
* Duration of the hydrothermal treatment was 10 h.
suspension in water after wash and red bars indicate pH after treatment at hydrothermal conditions (0.2 g of catalyst, 2
mL of water, 150 °C, 15 h.). In (b), red
Thebars represent of
interpretation H+leached
, white acid
barscatalysis
represent Cl–, and gray
is consistent withbars represent
the relative HSOof
activity 4– ions. *

Duration of the hydrothermal treatment was


the polymer 10 h. and the lack of activity observed for SH and HZSM-5 (Figure 2), since
catalysts
the pH measured for the supernatant liquid obtained from the polymers was comparable
3.2. to
Evaluation
pH valuesof False
used Negatives
in dilute acid pretreatment [23], whereas the pH obtained for SH and
especially HZSM-5 was much less acidic. The pH values of the SAC and SZ materials were
Figures 2 and 3 present data obtained at a single time point, which can obscure su
intermediate to these extremes, just as the activities of these catalysts were intermediate to
catalytic effects.and
the polymers In HZSM-5.
fact, the supernatant test risks being overly stringent, as the pH of
actual reaction
As part ofmixture
developingis expected to decrease
a reliable catalyst over
screening time, yet
approach, the the
we took pHextra
of the supernata
step of
fixed at a single value that is the lowest expected during the entire reaction.toAs a res
quantifying anion concentrations in the supernatant liquids. Specific attention was paid
chloride and sulfate species, in accordance with the structures of the original catalysts (as
the supernatant test is prone to identification of false negatives and cannot be relied o
shown in Figure 1). Figure 3b shows anion concentrations measured in the supernatants,
understand productmeasured
plotted alongside distributions.
H+ concentrations for comparison and for confirmation of
To help
charge understand
balance. the potential
All of the catalysts, for identification
except HZSM-5, of false
released anions negatives,
into solution we first sou
that could
be measured using anionic chromatography. HZSM-5 likely releases
to establish whether the supernatant test can identify heterogeneous activity when Al and Si species,
but they
expected to were not measured
be present, usinginglucose
this study [45,75]. Charges
dehydration as a associated with measuredprobe r
simple homogeneous
cation and anion concentrations were nearly balanced, as required for charge neutrality,
tion.indicating
Unlike cellulose
that no other hydrolysis,
anions wereglucose dehydration
released but not identifieddoes not require
and quantified. solid–solid
Of the five in
action, but can
catalysts that still be carried
released out at the
anions, CMP-SO same
3 H-0.3 conditions
released as cellulose
mainly chloride hydrolysis.
ions, consistent with Moreo
control tests on
hydrolysis glucose
of the can help
chloromethyl understand
group that has been thesuggested
potential as aeffect of solid acids on prod
carbohydrate-binding
distributions obtained from cellulose hydrolysis, even when the solidconsistent
group (Figure 1) [40]. The remainder of the catalysts release mainly bisulfate, acid may not
ticipate directly in hydrolysis.
For the purpose of testing solid acid activity for glucose dehydration, we used A
15, its supernatant, and sulfuric acid at the same pH as the supernatant. Analogousl
Figure 2, Figure 4 plots the average glucose conversion rate and shows that, at the co
Sustain. Chem. 2021, 2 657

with degradation of sulfonic acid groups in AMB-15, SAC, and SH and the sulfate ions of
SZ, resulting in sulfuric acid [76].
As a consistency check, the densities of acid sites on the solid catalysts were measured
before and after hydrothermal treatment to determine the source of the soluble ionic
species. Figure S1 in Supplementary Materials plots the amount of leached hydronium
ions, the decrease in acid sites, and the amount of leached bisulfate ions. The data show
that the loss of surface-bound acid sites is nearly in quantitative agreement with the
released homogeneous acid species, consistent with the soluble acids arising directly from
degradation of acid sites in the solid catalyst at the conditions used for hydrolysis.
The data in Figures 2 and 3 were then reanalyzed to identify underlying trends.
Figure S2 (see Supplementary Materials) shows the product formation pseudo-first-order
rate constants of the solid catalysts plotted versus the acid site density (Figure S2a) or the
concentration of the leached acid in the supernatant (Figure S2b). The former shows weak
correlation between acid site density and product formation rates, whereas Figure S2b
reveals a much clearer correlation between acid in the supernatant and product formation
rates. Again, Figure S2 points to a minor role for heterogeneous effects for the catalysts
considered in this study.
The supernatant liquid tests are, therefore, entirely self-consistent. Generation of a
supernatant liquid in the absence of cellulose results in a solution containing H+ and asso-
ciated anions, the identities and concentrations of which are consistent with expectations
and with measurements of acid site densities on the fresh and used solids. The approach is
equally applicable to catalysts of many different categories. Of the six materials studied
here, one lacks discernable activity when compared with noncatalytic conditions, and the
other five result in product formation rates that are less than that observed with the super-
natant liquid. The temptation is to conclude that heterogeneous effects are negligible for
all of these materials. However, questions remain, specifically if the test is too stringent,
resulting in false negatives, or too lax, resulting in false positives.

3.2. Evaluation of False Negatives


Figures 2 and 3 present data obtained at a single time point, which can obscure subtle
catalytic effects. In fact, the supernatant test risks being overly stringent, as the pH of the
actual reaction mixture is expected to decrease over time, yet the pH of the supernatant is
fixed at a single value that is the lowest expected during the entire reaction. As a result,
the supernatant test is prone to identification of false negatives and cannot be relied on to
understand product distributions.
To help understand the potential for identification of false negatives, we first sought
to establish whether the supernatant test can identify heterogeneous activity when it
is expected to be present, using glucose dehydration as a simple homogeneous probe
reaction. Unlike cellulose hydrolysis, glucose dehydration does not require solid–solid
interaction, but can still be carried out at the same conditions as cellulose hydrolysis.
Moreover, control tests on glucose can help understand the potential effect of solid acids
on product distributions obtained from cellulose hydrolysis, even when the solid acid may
not participate directly in hydrolysis.
For the purpose of testing solid acid activity for glucose dehydration, we used AMB-
15, its supernatant, and sulfuric acid at the same pH as the supernatant. Analogously
to Figure 2, Figure 4 plots the average glucose conversion rate and shows that, at the
conditions tested, the presence of solid acid catalyst resulted in nearly double the glucose
conversion when compared to its supernatant. This result is clear evidence of heteroge-
neous interactions between the solid acid and the soluble substrate that can be detected by
the single-point measurement approach used in this study. Interestingly, the supernatant is
still more active than sulfuric acid at the same pH, possibly implicating the role of counteri-
ons in the reaction mixture. Lastly, the greater activity observed for AMB-15 compared with
both the supernatant and sulfuric acid clearly shows that the solid acid can impact product
distributions obtained from cellulose hydrolysis, even under circumstances that apparently
Sustain. Chem. 2021, 2, FOR PEER REVIEW 14

Sustain. Chem. 2021, 2 counterions in the reaction mixture. Lastly, the greater activity observed for AMB-15 com- 658
pared with both the supernatant and sulfuric acid clearly shows that the solid acid can
impact product distributions obtained from cellulose hydrolysis, even under circum-
stances that apparently preclude direct participation in the primary hydrolysis reaction
preclude direct participation
step. Accordingly, incan
solid acids the justifiably
primary hydrolysis reaction
be used to controlstep. Accordingly,
product solida
distribution,
acids can justifiably be used to control product
potentially useful objective in some instances. distribution, a potentially useful objective
in some instances.

Average Conversion Rate 40

30
-3
(g/h) x10

20

10

0
AMB-15 Supernatant Sulfuric Acid
Figure 4. Comparison of the average glucose conversion rates after glucose dehydration with fresh
solid acid
Figure catalyst, its supernatant,
4. Comparison of the averageand sulfuric
glucose acid. The
conversion supernatant
rates wasdehydration
after glucose generated bywith
treating
fresh
AMB-15 at the same conditions as the glucose dehydration reaction. Sulfuric acid at the equivalent
solid acid catalyst, its supernatant, and sulfuric acid. The supernatant was generated by treating
concentration
AMB-15 at the(pH same= conditions
2.94) to that
as of
thethe supernatant
glucose was used
dehydration as a Sulfuric
reaction. control. acid
Conditions: 150 ◦ C
at the equivalent

±2.5 C) reaction
(concentration (pH temperature
= 2.94) to thatand 2 hsupernatant
of the reaction time,
was2used
mL ofas water, 0.1 Conditions:
a control. g of glucose,150
and°C0.2 g
(±2.5
°C)
of reaction temperature and 2 h reaction time, 2 mL of water, 0.1 g of glucose, and 0.2 g of catalyst.
catalyst.

Having
Having established
established that
that the
the supernatant
supernatant testtest yields
yields the
the expected
expected result
result for
for aacase
casein
in
which
whichheterogeneous
heterogeneouscatalysis
catalysisisis
expected,
expected, wewesought
soughtto quantify the the
to quantify effects of time
effects varying
of time var-
pH
yingonpH
product yields.
on product To doTothis,
yields. we adopted
do this, we adopteda reaction pathway
a reaction pathwayanalysis, as shown
analysis, in
as shown
Scheme 2. Here, the solid catalyst can potentially promote cellulose hydrolysis
in Scheme 2. Here, the solid catalyst can potentially promote cellulose hydrolysis or glu- or glucose
dehydration, eithereither
cose dehydration, directly by heterogeneous
directly by heterogeneous activity or indirectly
activity by leaching
or indirectly acidsacids
by leaching into
solution. Each of these reactions, except solid–solid-catalyzed cellulose hydrolysis,
into solution. Each of these reactions, except solid–solid-catalyzed cellulose hydrolysis, can be
described using existing kinetic models [55,56,77]; model comparison with
can be described using existing kinetic models [55,56,77]; model comparison with obser- observations
affords
vationsaaffords
method to identify
a method solid–solid
to identify catalysis.
solid–solid The concentration
catalysis. of H+ available
The concentration for
of H+ availa-
homogeneous
ble for homogeneous catalysis varies over time, an effect that can easily be captured inofa
catalysis varies over time, an effect that can easily be captured in a system
ordinary
system ofdifferential equations. equations.
ordinary differential
Sustain. Chem. 2021, 2 659
Sustain. Chem. 2021, 2, FOR PEER REVIEW 15

Scheme 2. Cellulose hydrolysis, product decomposition, and catalyst leaching reaction pathways. The red and green arrows
Scheme 2. Cellulose hydrolysis, product decomposition, and catalyst leaching reaction pathways. The red and green ar-
indicate homogeneous and heterogeneous catalyzed reactions, respectively.
rows indicate homogeneous and heterogeneous catalyzed reactions, respectively.
For the purposes of identifying solid–solid catalysis, we modeled cellulose hydrolysis
usingFor thetheratepurposes
expressions of identifying
and parameters solid–solid catalysis, we
recommended bymodeled
Saeman cellulose hydroly-
for homogeneous
sis using the rate expressions and parameters recommended by
acids [55]. Figure S3 (see Supplementary Materials) confirms the accuracy of this approach Saeman for homogeneous
byacids [55]. Figure
comparing its S3 (see Supplementary
prediction for hydrolysis Materials)
by each confirms the accuracy of Next,
of the supernatants. this approach
we cou-
by comparing its prediction for hydrolysis by each
pled Saeman’s model with a simple empirical model to capture time-dependent of the supernatants. Next, werelease
coupled of
Saeman’s model
homogeneous acidwith a simple
during empirical
reaction. For the model to capture
purpose of thetime-dependent release model,
empirical acid release of ho-
mogeneous acid
time-resolved during reaction.
measurements of the For
pH theofpurpose of the empirical
the supernatant acid release
liquid obtained from model,
CMP-
SOtime-resolved measurements of the pH of the supernatant liquid obtained from CMP-
3 H-0.3 and AMB-15 were made in the absence of cellulose (shown in Figures S4 and S5
inSO 3H-0.3 and AMB-15 were made in the absence of cellulose (shown in Figures S4 and S5
the Supplementary Materials). The supernatant pH measurements were consistent with
in the Supplementary
a first-order rate law with Materials).
a singleThe supernatant
best-fit pH measurements
rate constant, as shown inwere consistent
Figure with
S6. Potential
a first-order rate law with a single best-fit rate constant, as shown
heterogeneous effects were not explicitly included in the model; instead, model underpre- in Figure S6. Potential
heterogeneous
diction effects were
of experimental notcan
yields explicitly included
be attributed toin the model; instead,
heterogeneous model
catalysis, and underpre-
accurate
diction of experimental yields can be attributed to heterogeneous
predictions are attributable to a lack of heterogeneous effects on cellulose hydrolysis. catalysis, and accurate
predictions
The lastare attributable
element that we to aincluded
lack of heterogeneous
in the hydrolysis effects on cellulose
model is a wayhydrolysis.
to capture the
effects of uncertainties in key parameters. In fact, in our own experiments, capture
The last element that we included in the hydrolysis model is a way to the
we realized
effects of uncertainties in key parameters. In fact, in our own
that differences in thermocouple placement could lead to differences in measured reaction experiments, we realized
that differences
temperature of ±in2.5thermocouple
◦ C. Similarly, placement
the best-fit could
rate lead to differences
constants in measured
used to describe acid reaction
leaching
contained some uncertainty, as shown in Figure S6. Monte Carlo simulation leaching
temperature of ± 2.5 °C. Similarly, the best-fit rate constants used to describe acid methods
containedthe
captured some uncertainty,
effects of parameter as shown in Figure
uncertainty onS6. Monte
model Carlo simulation
predictions, methods
as explained cap-
in more
tured the effects
detail in Section 2. of parameter uncertainty on model predictions, as explained in more de-
tail in Section 2.
The next step was to use the kinetic model to generate a series of predictions of
celluloseTheconversion
next step was for to
useuseof the
AMB-15kinetic andmodel to generate
CMP-SO a series of predictions of cel-
3 H-0.3. These are shown in Figure 5
lulose conversion for use of AMB-15 and CMP-SO 3H-0.3. These are shown in Figure 5 as
as ranges resulting from the Monte Carlo simulations. In both cases, cellulose conversion
ranges resulting
increases smoothly fromwiththeincreasing
Monte Carlo simulations.
reaction In both cases,
times, reaching values cellulose conversion
in the range from
increases smoothly with increasing reaction times, reaching
18–25% for AMB-15 and 48–72% for CMP-SO3 H-0.3, with the difference corresponding values in the range from 18– to
25%
the for AMB-15
more rapid releaseand 48–72%
of acidsfor CMP-SO
from CMP-SO 3H-0.3, with the difference corresponding to the
H-0.3 compared with AMB-15 (Figures S4
3
more
and rapid release of acids from CMP-SO3H-0.3 compared with AMB-15 (Figures S4 and
S5).
S5).
We measured soluble product yields as a proxy for cellulose conversion at specified
timepoints obtained using CMP-SO3H-0.3 and AMB-15 to compare with the predictions
of the reaction model. Figure 6 contains plots of soluble product yields obtained for these
two catalysts. In the case of AMB-15, the model accurately predicts the observed yields
within the limits of uncertainty at all timepoints. The model actually overpredicts cellu-
lose hydrolysis for the case of CMP-SO3H-0.3, despite the fact that the reaction model
completely lacks any heterogeneous pathways. Therefore, Figure 5 clearly shows that the
risk of false negatives associated with the single-point method shown in Figure 3 is gen-
erally manageable. In cases where differentiating heterogeneous activity from homogene-
ous activity is especially important, the time-resolved method shown in Figure 6 can be
used.
Sustain. Chem. 2021, 2, FOR PEER REVIEW
Sustain. Chem. 2021, 2 660

80
CMP-SO3H-0.3
60

40

Soluble Carbon Balance (%)


Cellulose Conversion (%)

20

0
80
AMB-15

60

40

20

0
0 100 200 300 400 500 600
Time (min)
Figure
Figure5. Comparison
5. Comparison of cellulose
of cellulose conversion
conversion (gray)bypredicted
(gray) predicted by using
using Saeman’s Saeman’s
homogeneous acid homoge
cellulose
acid hydrolysis
cellulose model and
hydrolysis time-dependent
model acid concentration
and time-dependent and measured soluble
acid concentration carbon
and measured solub
balance (red) as a function of reaction time. Data presented are for CMP-SO H-0.3 (top)
bon balance (red) as a function of reaction time. Data presented3are for CMP-SO3H-0.3 (to and AMB-
15 (bottom).
AMB-15 (bottom).
We measured soluble product yields as a proxy for cellulose conversion at specified
Performing
timepoints obtainedtheusing
time-resolved
CMP-SO3 H-0.3 study is tedious,
and AMB-15 and less
to compare withlabor-intensive
the predictions appro
are, therefore, desirable. In fact, while the structure-agnostic approach
of the reaction model. Figure 6 contains plots of soluble product yields obtained forprovides
these no
two catalysts. In the case of AMB-15, the model accurately predicts the observed
information on the stability of acid sites on a given catalyst, hypothesizing that som yields
within the limits of uncertainty at all timepoints. The model actually overpredicts cellulose
are more labile than others is not unreasonable. Motivated by this line of thinkin
hydrolysis for the case of CMP-SO3 H-0.3, despite the fact that the reaction model completely
and Sievers
lacks recommended
any heterogeneous pretreating
pathways. candidate
Therefore, Figure 5solids
clearlyinshows
a harshthathydrothermal
the risk of en
ment
falseto removeassociated
negatives labile sites prior
with to using themethod
the single-point treated material
shown as a 3catalyst
in Figure as an alter
is generally
manageable. In cases where differentiating heterogeneous activity
method for eliminating homogeneous contributions to catalytic activity, as shofrom homogeneous
activity is especially important, the time-resolved method shown in Figure 6 can be used.
Scheme 1b [33].
We selected the SH catalyst for endurance testing, as SH activity when used as a
was equal (within uncertainty) to that of the supernatant, whereas the activity of
other catalysts tested here was exceeded by their supernatant. To simulate harsh
ment, the SH catalyst was subjected to a 24 h hydrothermal endurance test at 150 °C
ditions harsher than those used for activity testing, as shown in Scheme 1c. SH was
ered after the pretreatment and then used again for generation of a new supernatant
by treating it in water at 150°C for 15 h. As expected, the supernatant pH recover
the pretreated SH was less acidic than that recovered from the original SH, 2.8 com
with 2.0, while the solid material retained 1.73 mmol/g of surface bound acid sites
observation confirms that the endurance pretreatment successfully removed a fract
were nearly twice that observed for the solid SH. While the results of an endurance pre-
treatment test will depend on the catalyst being studied, Figure 6 clearly shows that an
endurance pretreatment is not sufficient to eliminate potential catalytic effects of leached
acids. Instead, we recommend the endurance test as a supplementary effort to the single-
Sustain. Chem. 2021, 2 point supernatant test and time-resolved study, as the supernatant test is a more discern-661
ing approach for identifying when homogeneous effects are contributing to reactivity and
the time-resolved study a more reliable way to reject false negatives.

Figure 6. Comparison of average product formation rate after hydrolysis of ball-milled cellulose with
hydrothermally pretreated SH solid catalyst and its supernatant. The supernatant was generated
Figure 6. Comparison of average product formation rate after hydrolysis of ball-milled cellulose
as previously
with described.
hydrothermally The dotted
pretreated line catalyst
SH solid represents
andthe
its hydrolysis activity
supernatant. of water. Conditions:
The supernatant was gener-
150 ◦ C (±2.5 ◦ C) reaction temperature 15 h reaction time, 2 mL of water, 0.1 g of cellulose, and 0.2 g
ated as previously described. The dotted line represents the hydrolysis activity of water. Conditions:
of catalyst.
150°C (±2.5 °C) reaction temperature 15 h reaction time, 2 mL of water, 0.1 g of cellulose, and 0.2 g
of catalyst.
Performing the time-resolved study is tedious, and less labor-intensive approaches
are, therefore, desirable.
3.3. False Positives ArisingIn fact,
from while the structure-agnostic
Catalyst–Reactant Interactions approach provides no direct
information on the stability of acid sites on a given catalyst, hypothesizing that some sites
The last question we sought to address was the potential for missing key information
are more labile than others is not unreasonable. Motivated by this line of thinking, Foo and
leading to false positives by generating the supernatant for testing catalytic activity in the
Sievers recommended pretreating candidate solids in a harsh hydrothermal environment to
absence of cellulose. In fact, the actual reaction mixture contains many soluble species that
remove labile sites prior to using the treated material as a catalyst as an alternative method
can potentially react with the catalyst surface, promote acid leaching, or block acid sites,
for eliminating homogeneous contributions to catalytic activity, as shown in Scheme 1b [33].
none of which would be captured by using a supernatant generated in the absence of cel-
We selected the SH catalyst for endurance testing, as SH activity when used as a solid
lulose and other reactants for comparison. For example, sulfonated catalysts have been
was equal (within uncertainty) to that of the supernatant, whereas the activity of most
shown to deactivate
other catalysts by forming
tested here esters with
was exceeded alcohols
by their duringTo
supernatant. esterification
simulate harsh of fatty acids
treatment,
[78] . Model
the SH underprediction
catalyst was subjectedoftoyields a 24 hobtained using CMP-SO
hydrothermal endurance 3H-0.3 as shown
test at 150 ◦ C, previously
conditions
in Figure 6 might plausibly be explained by these types of effects.
harsher than those used for activity testing, as shown in Scheme 1c. SH was recovered after
Accordingly,and
the pretreatment thethenacidused
content
againoffor
thegeneration
supernatants formed
of a new when CMP-SO
supernatant 3H-0.3 and
liquid by treating
AMB-15 were used
◦ for cellulose hydrolysis was measured and compared
it in water at 150 C for 15 h. As expected, the supernatant pH recovered for the pretreated with the super-
natants
SH was formed in the
less acidic thanabsence of cellulose.
that recovered fromFigure 7 provides
the original SH, 2.8 the results ofwith
compared this2.0,
compari-
while
son, showing
the solid that the
material acid content
retained of the supernatant
1.73 mmol/g obtained
of surface bound in the
acid presence
sites. of cellulose
This observation
was greater
confirms than
that the in its absence
endurance for both catalysts,
pretreatment by a factor
successfully removed of two in the case
a fraction of AMB-
of labile acid
15. Figure 7 clearly
sites, as intended. shows that reactant–catalyst interactions can play an important role in
acid leaching
The newand, hence, apparent
supernatant catalytic
and recovered SHreactivity.
were then The resulting
used acid content
for cellulose is greater
hydrolysis. On
when the supernatant is produced in the presence of cellulose, thereby
the basis of previous results, we elected to ball mill the cellulose prior to the activity opening the pos-
test,
sibility
with theofaim
identification
of increasing ofits
false positives
reactivity [4].by the single-point
Figure 6 provides the supernatant test.
results as average product
formation rates observed for the SH and the supernatant recovered after pretreatment. Even
with the pretreatment, product formation rates observed with the supernatant liquid were
nearly twice that observed for the solid SH. While the results of an endurance pretreatment
test will depend on the catalyst being studied, Figure 6 clearly shows that an endurance
pretreatment is not sufficient to eliminate potential catalytic effects of leached acids. Instead,
we recommend the endurance test as a supplementary effort to the single-point supernatant
test and time-resolved study, as the supernatant test is a more discerning approach for
Sustain. Chem. 2021, 2 662

ain. Chem. 2021, 2, FOR PEER REVIEW


identifying when homogeneous effects are contributing to reactivity and the time-resolved
study a more reliable way to reject false negatives.

3.3. False Positives Arising from Catalyst–Reactant Interactions


The last question we sought to address was the potential for missing key information
Practically, a solid catalyst that degrades to release homogen
leading to false positives by generating the supernatant for testing catalytic activity in the
ing with
absence soluble
of cellulose. products
In fact, the actual would not be
reaction mixture applicable
contains forspecies
many soluble cellulose h
that can potentially react with the catalyst surface, promote acid leaching, or block acid
merit
sites, nonefurther consideration.
of which would be captured by usingHowever,
a supernatanttogenerated
safeguard against
in the absence of false
heterogenous
cellulose catalysis,
and other reactants the supernatant
for comparison. test cancatalysts
For example, sulfonated be extended.
have been The
shown to deactivate by forming esters with alcohols during esterification of fatty acids [78].
the solid
Model acid should
underprediction of yieldsbe measured
obtained and3 H-0.3
using CMP-SO compared to the pH
as shown previously in of th
without
Figure 6 might presence
plausibly be of cellulose.
explained by theseIf theof pH
types of the solution after hyd
effects.
Accordingly, the acid content of the supernatants formed when CMP-SO3 H-0.3 and
is strong
AMB-15 evidence
were used of hydrolysis
for cellulose heterogeneous
was measured activity.
and comparedOn with
thethe other
super- hand,
natants
the caseformed ininFigure
the absence7,ofthen
cellulose. Figure 7 provides
additional testthecanresults
be ofcarried
this comparison,
out to qua
showing that the acid content of the supernatant obtained in the presence of cellulose was
products
greater than inon leaching
its absence kinetics.
for both Lastly,
catalysts, by a factor the kinetic
of two model
in the case propose
of AMB-15.
Figure 7 clearly shows that reactant–catalyst interactions
quantify the relative effects of homogeneous and heterogeneous can play an important role in acid
leaching and, hence, apparent catalytic reactivity. The resulting acid content is greater when
analysis
the supernatant presented
is produced inin theFigure
presence of5.cellulose, thereby opening the possibility of
identification of false positives by the single-point supernatant test.

0.14

0.12
Acid Concentration (M)

0.10

0.08

0.06

0.04

0.02

0.00
CMP-SO3H-0.3 AMB-15
Figure 7. Acid concentration in the liquid medium after hydrothermal treatment of CMP-SO3 H-0.3
and AMB-15 in the absence of cellulose (red) and after cellulose hydrolysis at the same conditions.
Figure 7. Acid concentration in the liquid medium after hydrothermal
Reaction was carried out at 150◦ C for 10 h.
t
and AMB-15 in the absence of cellulose (red) and after cellulose hydroly
Reaction was carried out at 150°C for 10 h.

To understand how the reaction mixture can influence ac


Sustain. Chem. 2021, 2 663

Practically, a solid catalyst that degrades to release homogeneous acids due to reacting
with soluble products would not be applicable for cellulose hydrolysis and, thus, not merit
further consideration. However, to safeguard against false-positive assignment of heteroge-
nous catalysis, the supernatant test can be extended. The pH after hydrolysis with the solid
acid should be measured and compared to the pH of the supernatant generated without
presence of cellulose. If the pH of the solution after hydrolysis is greater, then it is strong
evidence of heterogeneous activity. On the other hand, if the pH is lower, as is the case in
Figure 7, then additional test can be carried out to quantify the effect of soluble products
on leaching kinetics. Lastly, the kinetic model proposed here can be applied to quantify
the relative effects of homogeneous and heterogeneous catalysis similarly to the analysis
presented in Figure 5.
To understand how the reaction mixture can influence acid leaching, we treated CMP-
SO3 H-0.3 in a water solution of 13 C-enriched glucose (0.1 g glucose in 2 mL water) at the
reaction conditions applied for cellulose hydrolysis (150◦ C, 10 h), washed the catalyst with
water to remove soluble species, and characterized the remaining solid with solid-state
13 C-NMR. Figure 8a shows spectra arising from all carbon and of non-protonated or mobile

carbon. Figure 8b shows the two-dimensional correlation spectrum between protonated


and non-protonated or mobile carbons obtained for the charred CMP-SO3 H-0.3 catalyst
(see Section 2 for details). The spectra show clear signatures of levulinic acid (~20%)
and HMF (~4%). The majority of the signal is associated with various furanic and other
aromatic structures, ketones, and carboxylic acids typical of hydrothermal char [73,79].
The presence of these species confirms that the acid surface acts as a nucleation agent for
char, likely blocking access to acid sites and simultaneously detracting from the soluble
carbon yield. The appearance of levulinate species—potentially bound directly to sulfonic
acid on the catalyst surface—is especially interesting, as it suggests a reaction to form alkyl
esters on the surface [78]. The reaction between levulinic acid and sulfonic acid liberates
protons, thereby explaining the effect of the reaction mixture on the acid concentration
of the supernatant shown in Figure 7. Accordingly, Figures 7 and 8 provide a caution
that interactions between the reaction mixture and the catalyst can play important roles in
some instances.

3.4. Recommended Protocol


Scheme 3 summarizes a comprehensive structure-agnostic strategy for differenting
heterogeneous effects on cellulose hydrolysis from homogeneous ones. First, the catalyst
itself is tested and its cellulose hydrolysis activity compared to that of water, followed
by catalyst treatment at the same hydrothermal conditions to generate a supernatant;
the cellulose hydrolysis activity of the supernatant is then determined. Second, after the
supernatant comparison test, the stability of the catalyst should be tested in the presence of
soluble cellulose hydrolysis products by pH measurements of the cellulose and cellulose-
free supernatants. Differences should be taken into account when considering the results
of the supernatant test. Third, provided that the solid acid is stable in the presence of the
soluble species, kinetic modeling should be applied for detection and quantification of
heterogeneous activity. After heterogeneous catalysis is quantified, the solid material can
be subjected to structure–activity studies. In contrast, materials that do not exhibit greater
activity than water, are very unstable in water, degrade in the presence of the cellulose
hydrolysis products, or do not possess heterogeneous activity following kinetic modeling
analysis can be excluded from further consideration.
Sustain. Chem. 2021, 2 664
Sustain. Chem. 2021, 2, FOR PEER REVIEW 19

Figure Solid-state1313C-NMR
Figure8.8.Solid-state C-NMR analysis
analysisof ofCMP-SO
CMP-SO33H-0.3
H-0.3after
afterreaction with1313C-enriched
reactionwith C-enrichedglucose
glucose
(150 ◦
(150°C, 10 h).
C, 10 h.).(a)
(a)MultiCP
MultiCP spectra
spectra of of
allall C (black
C (black line)
line) andand
of Cof C bonded
not not bonded
to H to H mobile
or in or in mobile seg-
segments
ments
(red (red
line). line).
The The chemical
chemical shifts ofshifts of levulinic
levulinic acid (simple
acid (simple numbers) numbers) and of
and of HMF HMF (primed
(primed numbers)num-
are
bers) are indicated.
indicated. (b) 2D correlation
(b) 2D correlation between protonated
between protonated and non-protonated
and non-protonated or mobile
or mobile carbons carbons
present in
present
the in the
glucose glucose degradation
degradation products. Characteristic
products. Characteristic cross-peakscross-peaks
are labeled.are labeled.

3.4. Recommended
We recommend Protocol
that future work on solid acid-catalyzed cellulose hydrolysis adopts
at least steps 13and
Scheme 2 in Scheme
summarizes 3. If the results structure-agnostic
a comprehensive of these steps are promising, step
strategy for 3 should
differenting
allow differentiation of heterogeneous activity from what now
heterogeneous effects on cellulose hydrolysis from homogeneous ones. First,appears to be the almost
catalyst
unavoidably homogeneous variety. At that time, provided that heterogeneous
itself is tested and its cellulose hydrolysis activity compared to that of water, followed by activity
accounts for the majority
catalyst treatment of thehydrothermal
at the same observed catalysis, the catalyst
conditions post-reaction
to generate can be charac-
a supernatant; the cel-
terized, regenerated,
lulose hydrolysis and reused.
activity of the Adopting
supernatant these stepsdetermined.
is then will greatly reduce
Second,the likelihood
after of
the super-
future
natantmisattributions
comparison test, andthe
help allocate
stability ofresources to the
the catalyst mostbe
should promising
tested inmaterials.
the presence of
soluble cellulose hydrolysis products by pH measurements of the cellulose and cellulose-
free supernatants. Differences should be taken into account when considering the results
of the supernatant test. Third, provided that the solid acid is stable in the presence of the
soluble species, kinetic modeling should be applied for detection and quantification of
heterogeneous activity. After heterogeneous catalysis is quantified, the solid material can
be subjected to structure–activity studies. In contrast, materials that do not exhibit greater
activity than water, are very unstable in water, degrade in the presence of the cellulose
We recommend that future work on solid acid-catalyzed cellulose hydrolysis adopts
at least steps 1 and 2 in Scheme 3. If the results of these steps are promising, step 3 should
allow differentiation of heterogeneous activity from what now appears to be the almost
unavoidably homogeneous variety. At that time, provided that heterogeneous activity ac-
counts for the majority of the observed catalysis, the catalyst post-reaction can be charac-
Sustain. Chem. 2021, 2 665
terized, regenerated, and reused. Adopting these steps will greatly reduce the likelihood
of future misattributions and help allocate resources to the most promising materials.

Scheme 3. Process
Process flow
flow diagram
diagram of
of the
the structure-agnostic
structure-agnostic method for analyzing solid acid catalyst for cellulose
cellulose hydrolysis.

This study
This studyfocused
focusedon oncellulose
cellulosehydrolysis
hydrolysis asas
anan especially
especially interesting
interesting andand impor-
important
tant case.
test test case. That stated,
That stated, solidcatalysis
solid acid acid catalysis in water-rich
in water-rich solvents
solvents is an active
is an active area ofarea
re-
of research.
search. A major
A major challenge
challenge in liquid-phase
in liquid-phase catalysis
catalysis is differentiating
is differentiating effects effects
due to due
homo-to
homogeneous products of degradation from the desired heterogeneous ones.
geneous products of degradation from the desired heterogeneous ones. Scheme 3 can eas- Scheme 3 can
easily
ily be be adapted
adapted forfor reactions
reactions otherthan
other thancellulose
cellulosehydrolysis,
hydrolysis,providing
providing thethe community
community
with a useful tool for rejecting unpromising catalysts and enabling judicious
with a useful tool for rejecting unpromising catalysts and enabling judicious allocation allocation
of
of resources.
resources.
4. Conclusions
In this work, we developed a structure-agnostic methodology for evaluating solid
acid catalysts for cellulose hydrolysis and detection of heterogeneous activity and applied
it to six solid catalysts representing the main categories currently being considered. Com-
parison of results from a supernatant activity test with those obtained using the solid
material can identify circumstances when solid acid materials leach catalytically active
homogeneous acid species; every catalyst type considered here generated soluble acids
capable of hydrolyzing cellulose, indicating a role of homogeneous acids for these test
cases. By comparison, an alternative method consisting of harsh catalyst pretreatment was
less effective at identifying homogeneous contributions from a heterogeneous catalyst.
A weakness of using a supernatant generated in the absence of cellulose is that it will
contain a fixed amount of acid that may not be the same as that generated in the presence
of cellulose. This can result in identification of false negatives when the acid content of the
supernatant exceeds that of the actual reaction mixture or false positives in the opposite
Sustain. Chem. 2021, 2 666

case. For greater discernment, the time release of acids from the heterogeneous material
into the supernatant liquid can be measured, followed by modeling and measurement of
product yields obtained using the solid material. Similarly, reactions with soluble species,
e.g., glucose, can be used to elucidate reactant–catalyst interactions.
On the basis of those results, we recommend supernatant liquid tests as a simplistic,
yet powerful framework for rapid screening of solid acid materials for cellulose hydrolysis.
The proposed methodology relies only on activity testing and avoids labor-intensive struc-
tural characterization. Only after a material’s heterogeneous activity has been established
and verified should structure–activity studies be initiated. The approach can be utilized
by the community for the design and development of solid acid catalysts for cellulose
hydrolysis. Furthermore, it can be extended and implemented to other systems where
catalyst degradation is of concern for the interpretation and quantification of activity.

Supplementary Materials: The following are available online at https://www.mdpi.com/article/


10.3390/suschem2040036/s1. Figure S1: Comparison of the leached homogeneous acid concentra-
tion per gram of catalyst, the decrease of acid sites of the catalysts post hydrothermal treatment,
and the concentration of the leached bisulfate species per gram of catalyst. Figure S2: Correlation
between first order catalytic hydrolysis rate constant kcat of cellulose hydrolysis using fresh solid
acid catalyst and amount of acid sites and concentration of leached homogeneous acid. Figure S3:
Parity plot of measured soluble carbon balance versus predicted cellulose conversion by Saeman’s
homogeneous acid hydrolysis model. Figure S4: Concentration of the leached homogeneous acid
for Amberlyst-15 as a function of treatment time at 150 ◦ C. Figure S5: Concentration of the leached
homogeneous acid for CMP-SO3 H-0.3 as a function of treatment time at 150 ◦ C. Figure S6: Kinetic
analysis of homogeneous acid leaching from AMB-15 and CMP-SO3 H-0.3 assuming the leaching
obeys first order kinetics. Table S1: Acid site density of the catalysts used in the current study
measured by solid state Boehm titration. Table S2: Comparison of the yields (%) of most abun-
dant soluble products generated by solid acid and leached homogeneous acid catalyzed cellulose
hydrolysis.
Author Contributions: Conceptualization, M.T. and M.T.T.; methodology, M.T., J.F.; software, M.T.
and J.F.; formal analysis, M.T., M.T.T., P.D. and K.S.-R.; investigation, M.T., J.F., C.S., S.S., W.Y. and
Z.Z.; data curation, M.T. and P.D.; writing—M.T. and M.T.T.; writing—M.T., M.T.T., Z.Z., P.D. and
K.S.-R.; supervision, M.T.T.; project administration, M.T.T.; funding acquisition, M.T.T. All authors
have read and agreed to the published version of the manuscript.
Funding: The US National Science Foundation supported this work under ENG/155428.
Data Availability Statement: The data presented in this study are available in the main manuscript
or in the Supplementary Materials.
Conflicts of Interest: The authors declare no conflict of interests.

References
1. Klemm, D.; Heublein, B.; Fink, H.-P.; Bohn, A. Cellulose: Fascinating Biopolymer and Sustainable Raw Material. Angew. Chem.
Int. Ed. 2005, 44, 3358–3393. [CrossRef]
2. Lynd, L. The grand challenge of cellulosic biofuels. Nat. Biotechnol. 2017, 35, 912–915. [CrossRef] [PubMed]
3. Isikgor, F.H.; Becer, C.R. Lignocellulosic biomass: A sustainable platform for the production of bio-based chemicals and polymers.
Polym. Chem. 2015, 6, 4497–4559. [CrossRef]
4. Tyufekchiev, M.; Kolodziejczak, A.; Duan, P.; Foston, M.; Schmidt-Rohr, K.; Timko, M.T. Reaction engineering implications of
cellulose crystallinity and water-promoted recrystallization. Green Chem. 2019, 21, 5541–5555. [CrossRef]
5. Endo, T.; Aung, E.M.; Fujii, S.; Hosomi, S.; Kimizu, M.; Ninomiya, K.; Takahashi, K. Investigation of accessibility and reactivity of
cellulose pretreated by ionic liquid at high loading. Carbohydr. Polym. 2017, 176, 365–373. [CrossRef] [PubMed]
6. Zhao, H.; Jones, C.L.; Baker, G.; Xia, S.; Olubajo, O.; Person, V.N. Regenerating cellulose from ionic liquids for an accelerated
enzymatic hydrolysis. J. Biotechnol. 2009, 139, 47–54. [CrossRef] [PubMed]
7. Chundawat, S.P.S.; Bellesia, G.; Uppugundla, N.; Sousa, L.D.C.; Gao, D.; Cheh, A.M.; Agarwal, U.P.; Bianchetti, C.M.; Phillips,
J.G.N.; Langan, P.; et al. Restructuring the Crystalline Cellulose Hydrogen Bond Network Enhances Its Depolymerization Rate. J.
Am. Chem. Soc. 2011, 133, 11163–11174. [CrossRef]
Sustain. Chem. 2021, 2 667

8. Zhang, Q.; Benoit, M.; Vigier, K.D.O.; Barrault, J.; Jégou, G.; Philippe, M.; Jérôme, F. Pretreatment of microcrystalline cellulose by
ultrasounds: Effect of particle size in the heterogeneously-catalyzed hydrolysis of cellulose to glucose. Green Chem. 2013, 15,
963–969. [CrossRef]
9. Wu, K.; Feng, G.; Liu, Y.; Liu, C.; Zhang, X.; Liu, S.; Liang, B.; Lu, H. Enhanced hydrolysis of mechanically pretreated cellulose in
water/CO2 system. Bioresour. Technol. 2018, 261, 28–35. [CrossRef]
10. Tao, L.; Schell, D.; Davis, R.; Tan, E.; Elander, R.; Bratis, A. NREL 2012 Achievement of Ethanol Cost Targets: Biochemical Ethanol
Fermentation via Dilute-Acid Pretreatment and Enzymatic Hydrolysis of Corn Stover; National Renewable Energy Lab. (NREL): Golden,
CO, USA, 2014.
11. Yin, S.; Pan, Y.; Tan, Z. Hydrothermal Conversion of Cellulose to 5-Hydroxymethyl Furfural. Int. J. Green Energy 2011, 8, 234–247.
[CrossRef]
12. Zhu, Y.; Biddy, M.J.; Jones, S.B.; Elliott, D.; Schmidt, A.J. Techno-economic analysis of liquid fuel production from woody biomass
via hydrothermal liquefaction (HTL) and upgrading. Appl. Energy 2014, 129, 384–394. [CrossRef]
13. Möller, M.; Harnisch, F.; Schröder, U. Hydrothermal liquefaction of cellulose in subcritical water—the role of crystallinity on the
cellulose reactivity. RSC Adv. 2013, 3, 11035–11044. [CrossRef]
14. Questell-Santiago, Y.M.; Zambrano-Varela, R.; Amiri, M.T.; Luterbacher, J.S. Carbohydrate stabilization extends the kinetic limits
of chemical polysaccharide depolymerization. Nat. Chem. 2018, 10, 1222–1228. [CrossRef] [PubMed]
15. Cabiac, A.; Guillon, E.; Chambon, F.; Pinel, C.; Rataboul, F.; Essayem, N. Cellulose reactivity and glycosidic bond cleavage in
aqueous phase by catalytic and non catalytic transformations. Appl. Catal. A Gen. 2011, 402, 1–10. [CrossRef]
16. Negahdar, L.; Delidovich, I.; Palkovits, R. Aqueous-phase hydrolysis of cellulose and hemicelluloses over molecular acidic
catalysts: Insights into the kinetics and reaction mechanism. Appl. Catal. B Environ. 2016, 184, 285–298. [CrossRef]
17. Pala, H.; Mota, M.; Gama, F.M. Enzymatic depolymerisation of cellulose. Carbohydr. Polym. 2007, 68, 101–108. [CrossRef]
18. Nill, J.; Karuna, N.; Jeoh, T. The impact of kinetic parameters on cellulose hydrolysis rates. Process. Biochem. 2018, 74, 108–117.
[CrossRef]
19. Hall, M.; Bansal, P.; Lee, J.H.; Realff, M.J.; Bommarius, A.S. Cellulose crystallinity—A key predictor of the enzymatic hydrolysis
rate. FEBS J. 2010, 277, 1571–1582. [CrossRef]
20. He, J.; Liu, M.; Huang, K.; Walker, T.W.; Maravelias, C.T.; Dumesic, J.A.; Huber, G.W. Production of levoglucosenone and
5-hydroxymethylfurfural from cellulose in polar aprotic solvent–water mixtures. Green Chem. 2017, 19, 3642–3653. [CrossRef]
21. Ghosh, A.; Brown, R.C.; Bai, X. Production of solubilized carbohydrate from cellulose using non-catalytic, supercritical depoly-
merization in polar aprotic solvents. Green Chem. 2015, 18, 1023–1031. [CrossRef]
22. Lynd, L.R.; Liang, X.; Biddy, M.J.; Allee, A.; Cai, H.; Foust, T.; Himmel, M.E.; Laser, M.S.; Wang, M.; Wyman, C.E. Cellulosic
ethanol: Status and innovation. Curr. Opin. Biotechnol. 2017, 45, 202–211. [CrossRef]
23. Pingali, S.V.; Urban, V.S.; Heller, W.T.; McGaughey, J.; O’Neill, H.; Foston, M.B.; Li, H.; Wyman, C.E.; Myles, D.A.; Langan, P.; et al.
Understanding Multiscale Structural Changes During Dilute Acid Pretreatment of Switchgrass and Poplar. ACS Sustain. Chem.
Eng. 2016, 5, 426–435. [CrossRef]
24. Pu, Y.; Hu, F.; Huang, F.; Davison, B.H.; Ragauskas, A.J. Assessing the molecular structure basis for biomass recalcitrance during
dilute acid and hydrothermal pretreatments. Biotechnol. Biofuels 2013, 6, 15. [CrossRef] [PubMed]
25. Foston, M.; Ragauskas, A.J. Changes in lignocellulosic supramolecular and ultrastructure during dilute acid pretreatment of
Populus and switchgrass. Biomass-Bioenergy 2010, 34, 1885–1895. [CrossRef]
26. Yang, B.; Dai, Z.; Ding, S.-Y.; Wyman, C.E. Enzymatic hydrolysis of cellulosic biomass. Biofuels 2011, 2, 421–449. [CrossRef]
27. Huang, Y.-B.; Fu, Y. Hydrolysis of cellulose to glucose by solid acid catalysts. Green Chem. 2013, 15, 1095–1111. [CrossRef]
28. Onda, A.; Ochi, T.; Yanagisawa, K. Selective hydrolysis of cellulose into glucose over solid acid catalysts. Green Chem. 2008, 10,
1033–1037. [CrossRef]
29. Tarabanko, N.; Tarabanko, V.E.; Kukhtetskiy, S.V.; Taran, O. Electrical Double Layer as a Model of Interaction between Cellulose
and Solid Acid Catalysts of Hydrolysis. ChemPhysChem 2019, 20, 706–718. [CrossRef]
30. Dumesic, J.A.; Huber, G.W.; Boudart, M. Principles of Heterogeneous Catalysis. In Handbook of Heterogeneous Catalysis; Wiley:
Hoboken, NJ, USA, 2008.
31. Shuai, L.; Pan, X. Hydrolysis of cellulose by cellulase-mimetic solid catalyst. Energy Environ. Sci. 2012, 5, 6889–6894. [CrossRef]
32. Yuan, S.; Li, T.; Wang, Y.; Cai, B.; Wen, X.; Shen, S.; Peng, X.; Li, Y. Double-adsorption functional carbon based solid acids derived
from copyrolysis of PVC and PE for cellulose hydrolysis. Fuel 2018, 237, 895–902. [CrossRef]
33. Foo, G.S.; Sievers, C. Synergistic Effect between Defect Sites and Functional Groups on the Hydrolysis of Cellulose over Activated
Carbon. ChemSusChem 2014, 8, 534–543. [CrossRef] [PubMed]
34. Hu, L.; Li, Z.; Wu, Z.; Lin, L.; Zhou, S. Catalytic hydrolysis of microcrystalline and rice straw-derived cellulose over a chlorine-
doped magnetic carbonaceous solid acid. Ind. Crop. Prod. 2016, 84, 408–417. [CrossRef]
35. Zuo, Y.; Zhang, Y.; Fu, Y. Catalytic Conversion of Cellulose into Levulinic Acid by a Sulfonated Chloromethyl Polystyrene Solid
Acid Catalyst. ChemCatChem 2014, 6, 753–757. [CrossRef]
36. Qian, X.; Lei, J.; Wickramasinghe, S.R. Novel polymeric solid acid catalysts for cellulose hydrolysis. RSC Adv. 2013, 3, 24280–24287.
[CrossRef]
37. Yang, Q.; Pan, X. Synthesis and Application of Bifunctional Porous Polymers Bearing Chloride and Sulfonic Acid as Cellulase-
Mimetic Solid Acids for Cellulose Hydrolysis. BioEnergy Res. 2015, 9, 578–586. [CrossRef]
Sustain. Chem. 2021, 2 668

38. Bornscheuer, U.; Buchholz, K.; Seibel, J. Enzymatic Degradation of (Ligno)cellulose. Angew. Chem. Int. Ed. 2014, 53, 10876–10893.
[CrossRef] [PubMed]
39. Gazit, O.M.; Katz, A. Understanding the Role of Defect Sites in Glucan Hydrolysis on Surfaces. J. Am. Chem. Soc. 2013, 135,
4398–4402. [CrossRef]
40. Tyufekchiev, M.V.; Duan, P.; Schmidt-Rohr, K.; Granados-Focil, S.; Timko, M.T.; Emmert, M.H. Cellulase-Inspired Solid Acids for
Cellulose Hydrolysis: Structural Explanations for High Catalytic Activity. ACS Catal. 2018, 8, 1464–1468. [CrossRef]
41. Scholz, D.F.; Kröcher, O.; Vogel, F. Deactivation and Regeneration of Sulfonated Carbon Catalysts in Hydrothermal Reaction
Environments. ChemSusChem 2018, 11, 2189–2201. [CrossRef]
42. Liu, Z.; Liu, Z. Comparison of hydrochar- and pyrochar-based solid acid catalysts from cornstalk: Physiochemical properties,
catalytic activity and deactivation behavior. Bioresour. Technol. 2020, 297, 122477. [CrossRef]
43. Sádaba, I.; Granados, M.L.; Riisager, A.; Taarning, E. Deactivation of solid catalysts in liquid media: The case of leaching of active
sites in biomass conversion reactions. Green Chem. 2015, 17, 4133–4145. [CrossRef]
44. Matosin, N.; Frank, E.; Engel, M.; Lum, J.; Newell, K. Negativity towards negative results: A discussion of the disconnect between
scientific worth and scientific culture. Dis. Model. Mech. 2014, 7, 171–173. [CrossRef] [PubMed]
45. Gardner, D.W.; Huo, J.; Hoff, T.C.; Johnson, R.L.; Shanks, B.H.; Tessonnier, J.-P. Insights into the Hydrothermal Stability of ZSM-5
under Relevant Biomass Conversion Reaction Conditions. ACS Catal. 2015, 5, 4418–4422. [CrossRef]
46. Nguyen, V.C.; Dandach, A.; Vu, T.T.H.; Fongarland, P.; Essayem, N. ZrW catalyzed cellulose conversion in hydrothermal
conditions: Influence of the calcination temperature and insights on the nature of the active phase. Mol. Catal. 2019, 476.
[CrossRef]
47. Li, H.-X.; Zhang, X.; Wang, Q.; Yang, D.; Cao, Q.; Jin, L. Study on the hydrolysis of cellulose with the regenerable and recyclable
multifunctional solid acid as a catalyst and its catalytic hydrolytic kinetics. Cellulose 2019, 27, 285–300. [CrossRef]
48. Pham, S.T.; Nguyen, M.B.; Le, G.H.; Nguyen, T.D.; Pham, C.D.; Le, T.S.; Vu, T.A. Influence of Brønsted and Lewis acidity of the
modified Al-MCM-41 solid acid on cellulose conversion and 5-hydroxylmethylfurfuran selectivity. Chemosphere 2020, 265, 129062.
[CrossRef] [PubMed]
49. Hu, J.; Wang, Q.; Wang, W.; Xu, Z.; Fu, J.; Xu, Q.; Wang, Z.; Yuan, Z.; Shen, F.; Qi, W. Synthesis of a Stable Solid Acid Catalyst from
Chloromethyl Polystyrene through a Simple Sulfonation for Pretreatment of Lignocellulose in Aqueous Solution. ChemSusChem
2020, 14, 979–989. [CrossRef]
50. Yang, Q.; Pan, X. Introducing hydroxyl groups as cellulose-binding sites into polymeric solid acids to improve their catalytic
performance in hydrolyzing cellulose. Carbohydr. Polym. 2021, 261, 117895. [CrossRef]
51. Van Pelt, A.H.; Simakova, O.A.; Schimming, S.M.; Ewbank, J.L.; Foo, G.S.; Pidko, E.; Hensen, E.J.; Sievers, C. Stability of
functionalized activated carbon in hot liquid water. Carbon 2014, 77, 143–154. [CrossRef]
52. Gong, R.; Ma, Z.; Wang, X.; Han, Y.; Guo, Y.; Sun, G.; Li, Y.; Zhou, J. Sulfonic-acid-functionalized carbon fiber from waste
newspaper as a recyclable carbon based solid acid catalyst for the hydrolysis of cellulose. RSC Adv. 2019, 9, 28902–28907.
[CrossRef]
53. Parveen, F.; Gupta, K.; Upadhyayula, S. Synergistic effect of chloro and sulphonic acid groups on the hydrolysis of microcrystalline
cellulose under benign conditions. Carbohydr. Polym. 2017, 159, 146–151. [CrossRef]
54. Goertzen, S.L.; Thériault, K.D.; Oickle, A.M.; Tarasuk, A.C.; Andreas, H.A. Standardization of the Boehm titration. Part I. CO2
expulsion and endpoint determination. Carbon 2010, 48, 1252–1261. [CrossRef]
55. Saeman, J.F. Kinetics of Wood Saccharification—Hydrolysis of Cellulose and Decomposition of Sugars in Dilute Acid at High
Temperature. Ind. Eng. Chem. 1945, 37, 43–52. [CrossRef]
56. Girisuta, B.; Janssen, A.L.P.B.M.; Heeres, H.J. Kinetic Study on the Acid-Catalyzed Hydrolysis of Cellulose to Levulinic Acid. Ind.
Eng. Chem. Res. 2007, 46, 1696–1708. [CrossRef]
57. SriBala, G.; Vinu, R. Unified Kinetic Model for Cellulose Deconstruction via Acid Hydrolysis. Ind. Eng. Chem. Res. 2014, 53,
8714–8725. [CrossRef]
58. Camacho, F.; González-Tello, P.; Jurado, E.; Robles, A. Microcrystalline-Cellulose Hydrolysis with Concentrated Sulphuric Acid. J.
Chem. Technol. Biotechnol. 1996, 67, 350–356. [CrossRef]
59. Fagan, R.D.; Grethlein, H.E.; Converse, A.O.; Porteous, A. Kinetics of the acid hydrolysis of cellulose found in paper refuse.
Environ. Sci. Technol. 1971, 5, 545–547. [CrossRef]
60. Duan, P.; Schmidt-Rohr, K. Composite-pulse and partially dipolar dephased multiCP for improved quantitative solid-state 13 C
NMR. J. Magn. Reson. 2017, 285, 68–78. [CrossRef] [PubMed]
61. Mao, J.-D.; Schmidt-Rohr, K. Accurate Quantification of Aromaticity and Nonprotonated Aromatic Carbon Fraction in Natural
Organic Matter by 13C Solid-State Nuclear Magnetic Resonance. Environ. Sci. Technol. 2004, 38, 2680–2684. [CrossRef] [PubMed]
62. Johnson, R.L.; Anderson, J.M.; Shanks, B.H.; Fang, X.; Hong, M.; Schmidt-Rohr, K. Spectrally edited 2D 13C13C NMR spectra
without diagonal ridge for characterizing 13C-enriched low-temperature carbon materials. J. Magn. Reson. 2013, 234, 112–124.
[CrossRef]
63. Zhang, B.; Ren, J.; Liu, X.; Guo, Y.; Guo, Y.; Lu, G.; Wang, Y. Novel sulfonated carbonaceous materials from p-toluenesulfonic
acid/glucose as a high-performance solid-acid catalyst. Catal. Commun. 2010, 11, 629–632. [CrossRef]
64. Kassaye, S.; Pagar, C.; Pant, K.K.; Jain, S.; Gupta, R. Depolymerization of microcrystalline cellulose to value added chemicals
using sulfate ion promoted zirconia catalyst. Bioresour. Technol. 2016, 220, 394–400. [CrossRef]
Sustain. Chem. 2021, 2 669

65. Lanzafame, P.; Temi, D.; Perathoner, S.; Spadaro, A.; Centi, G. Direct conversion of cellulose to glucose and valuable intermediates
in mild reaction conditions over solid acid catalysts. Catal. Today 2012, 179, 178–184. [CrossRef]
66. Suganuma, S.; Nakajima, K.; Kitano, M.; Yamaguchi, D.; Kato, H.; Hayashi, S.; Hara, M. Hydrolysis of Cellulose by Amorphous
Carbon Bearing SO3H, COOH, and OH Groups. J. Am. Chem. Soc. 2008, 130, 12787–12793. [CrossRef] [PubMed]
67. Bhaumik, P.; Dhepe, P.L. Exceptionally high yields of furfural from assorted raw biomass over solid acids. RSC Adv. 2014, 4,
26215–26221. [CrossRef]
68. Chu, S.; Yang, L.; Guo, X.; Dong, L.; Chen, X.; Li, Y.; Mu, X. The influence of pore structure and Si/Al ratio of HZSM-5 zeolites on
the product distributions of α-cellulose hydrolysis. Mol. Catal. 2018, 445, 240–247. [CrossRef]
69. Dhepe, P.L.; Fukuoka, A. Cellulose Conversion under Heterogeneous Catalysis. ChemSusChem 2008, 1, 969–975. [CrossRef]
70. Onda, A.; Ochi, T.; Yanagisawa, K. Hydrolysis of Cellulose Selectively into Glucose Over Sulfonated Activated-Carbon Catalyst
Under Hydrothermal Conditions. Top. Catal. 2009, 52, 801–807. [CrossRef]
71. Kianfar, E. Comparison and assessment of zeolite catalysts performance dimethyl ether and light olefins production through
methanol: A review. Rev. Inorg. Chem. 2019, 39, 157–177. [CrossRef]
72. Rabee, A.I.M.; Mekhemer, G.A.H.; Osatiashtiani, A.; Isaacs, M.A.; Lee, A.F.; Wilson, K.; Zaki, M.I. Acidity-Reactivity Relationships
in Catalytic Esterification over Ammonium Sulfate-Derived Sulfated Zirconia. Catalysts 2017, 7, 204. [CrossRef]
73. Brown, A.; McKeogh, B.; Tompsett, G.; Lewis, R.; Deskins, N.; Timko, M. Structural analysis of hydrothermal char and its models
by density functional theory simulation of vibrational spectroscopy. Carbon 2017, 125, 614–629. [CrossRef]
74. Anderson, J.M.; Johnson, R.L.; Schmidt-Rohr, K.; Shanks, B.H. Hydrothermal degradation of model sulfonic acid compounds:
Probing the relative sulfur–carbon bond strength in water. Catal. Commun. 2014, 51, 33–36. [CrossRef]
75. Maag, A.R.; Tompsett, G.A.; Tam, J.; Ang, C.A.; Azimi, G.; Carl, A.D.; Huang, X.; Smith, L.J.; Grimm, R.L.; Bond, J.Q.; et al. ZSM-5
decrystallization and dealumination in hot liquid water. Phys. Chem. Chem. Phys. 2019, 21, 17880–17892. [CrossRef] [PubMed]
76. Anderson, J.M.; Johnson, R.L.; Schmidt-Rohr, K.; Shanks, B.H. Solid state NMR study of chemical structure and hydrothermal
deactivation of moderate-temperature carbon materials with acidic SO3H sites. Carbon 2014, 74, 333–345. [CrossRef]
77. Weingarten, R.; Cho, J.; Xing, R.; Jr, W.C.C.; Huber, G.W. Kinetics and Reaction Engineering of Levulinic Acid Production from
Aqueous Glucose Solutions. ChemSusChem 2012, 5, 1280–1290. [CrossRef]
78. Fraile, J.M.; García-Bordejé, E.; Pires, E.; Roldán, L. Catalytic performance and deactivation of sulfonated hydrothermal carbon in
the esterification of fatty acids: Comparison with sulfonic solids of different nature. J. Catal. 2015, 324, 107–118. [CrossRef]
79. Brown, A.B.; Tompsett, G.A.; Partopour, B.; Deskins, N.A.; Timko, M.T. Hydrochar structural determination from artifact-free
Raman analysis. Carbon 2020, 167, 378–387. [CrossRef]

You might also like