You are on page 1of 10

pubs.acs.

org/journal/ascecg Research Article

Pretreatment with Sodium Methyl Mercaptide Increases


Carbohydrate Yield during Kraft Pulping
Connor J. Cooper,⊥ Ravikant Patil,⊥ Lintao Bu,⊥ Zhongyu Mou, David B. Turpin,
Adriaan van Heiningen,* Jerry M. Parks,* and Brandon C. Knott*
Cite This: ACS Sustainable Chem. Eng. 2021, 9, 11571−11580 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Downloaded via NATL INST OF TECH WARANGAL on June 14, 2023 at 12:06:24 (UTC).

ABSTRACT: Kraft pulping is the predominant technology in the pulp


and paper industry for removing lignin from wood carbohydrates to
produce paper, board, packaging, tissue, and specialty cellulose. However,
the kraft process is energy intensive and expensive, and its yield is limited
by the degradation of carbohydrates. Pretreatment can increase
carbohydrate yield by limiting degradation via primary peeling of
reducing end groups. However, protection of galactoglucomannan
(GGM), the primary hemicellulose component of softwood, is minimal
when conventional pretreatments are used. Here, we investigated the
effectiveness of sodium methyl mercaptide pretreatment on southern pine
wood chips under a range of experimental conditions. We found that
pretreatment of biomass with 4.38% sodium methyl mercaptide at pH 12
and 105 °C for 60 min provided small but significant increases in xylan
and cellulose yields relative to control conditions, but preservation of GGM was minimal. To provide insight into molecular-scale
details of primary peeling, pretreatment, and alkaline hydrolysis, we performed classical molecular dynamics (MD) simulations under
selected process conditions and quantum mechanical (QM) calculations of selected reactions. MD simulations showed that C1 of
the GGM reducing end is more readily accessible by HO− and CH3S− ions than in cellulose. The free energy barrier for peeling
calculated with QM is lower for GGM than for cellulose, indicating increased susceptibility to peeling. In addition, we found that
GGM may be more susceptible to internal chain cleavage than cellulose. Thus, even though reducing end groups may be protected
initially through pretreatment, new unprotected reducing end groups may be generated through alkaline hydrolysis. Taken together,
these findings show the promise of methyl mercaptide as a pretreatment technology for cellulose retention and also provide
molecular insight for improving its effectiveness toward GGM.
KEYWORDS: Kraft pulping, Pretreatment, Biomass, Degradation, Peeling, Alkaline hydrolysis, MD, DFT

■ INTRODUCTION
The forest products industry creates consumer goods necessary
known for being odorous. The odor comes from a combination
of methyl mercaptan, dimethyl sulfide, dimethyl disulfide, and
for everyday life from renewable resources, accounting for H2S, although the first two are dominant. Methyl mercaptide
(CH3S−) is formed by a nucleophilic attack of HS− on the
approximately 4% of the total U.S. manufacturing gross
methoxy groups of lignin.
domestic product, manufacturing nearly $300 billion in
Wood is composed of 40%−45% cellulose, 25%−30%
products annually and employing approximately 950,000
hemicellulose, and 25−30% lignin, whereas the final kraft
men and women. 1 The industry meets a payroll of
pulp consists of 35%−40% cellulose, 8%−13% hemicellulose,
approximately $55 billion annually and is among the top 10
and 1%−2% lignin.4 Cellulose is a homopolymer consisting of
manufacturing sector employers in 45 states. In the
repeating D-anhydroglucopyranose units linked by β-1,4
manufacture of pulp, the removal of lignin from wood
glycosidic bonds (Figure 1). Galactoglucomannan (GGM),
carbohydrates is predominantly accomplished via the kraft
the primary hemicellulose in softwood, is a heteropolymer
process, resulting in global pulp production of about 200
million metric tonnes annually.2 The kraft process is based on
dissolution of lignin from wood by reaction at high Received: June 25, 2021
temperature (150−170 °C) with an aqueous solution of Revised: August 4, 2021
NaOH and Na2S.3 This process is energy and capital intensive Published: August 20, 2021
and provides a less efficient use of wood resources than desired
due to substantial degradation of wood polysaccharides,
hemicelluloses and cellulose.3 In addition, the kraft process is

© 2021 American Chemical Society https://doi.org/10.1021/acssuschemeng.1c04332


11571 ACS Sustainable Chem. Eng. 2021, 9, 11571−11580
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

Figure 1. Chemical structures of cellulose and galactoglucomannan (GGM).

Figure 2. Selected reactions for (A) primary peeling, (B) stopping, (C) pretreatment, and (D) alkaline hydrolysis.

consisting of β-1,4-linked glucose and mannose (1:3−4 ratio) xylan is retained when the alkalinity is kept low throughout the
as a backbone with some mannose units substituted by α-1,6- pulping process. In contrast, some experiments have shown
linked galactose. Maximizing retention of these polysaccharides primary peeling to degrade more than half of the original GGM
through the pulping process is essential to plant economics. during impregnation with NaOH and Na2S,8 with further
At high pH, the reducing end groups (REGs) of cellulose losses due to secondary peeling. As a result, the final kraft pulp
and GGM polymers can undergo a stepwise depolymerization contains only ∼25% of the native glucomannan.4
reaction called primary peeling (Figure 2A), resulting in In general, pulp yield is less than 50% compared to a
carbohydrate loss to the black liquor. Primary peeling of potential maximum yield of 70%−75% when both cellulose
cellulose becomes significant above 130 °C,5,6 although the and hemicelluloses are fully retained. Estimated pulp yield
onset of primary peeling for GGM occurs at much lower increases from completely eliminating primary peeling is ∼12%
temperature (around 80 °C).7 Chain ends can also undergo for softwoods (4% cellulose, 8% GGM) and ∼6% for
stopping reactions, in which the reducing end groups are hardwoods (4% cellulose, 2% glucomannan).4 Yield increases
converted into alkali-stable carboxylate groups (Figure 2B). In of this magnitude could have a profound impact on plant
addition, at temperatures above 130 °C, the internal glycosidic economics. Various approaches have been developed to
bonds in a polysaccharide chain are susceptible to alkaline minimize primary peeling of carbohydrates by converting the
hydrolysis (Figure 2D). When hydrolysis occurs, newly formed reducing end groups into alkali-stable groups through both
chain ends constitute REGs and thus possible sites for further oxidative and reductive reactions (Figure 2C). Oxidative
stepwise degradation (i.e., secondary peeling). Temperature, pretreatments include the use of anthraquinone or polysulfides,
alkali, and other pulping conditions all influence the relative and examples of reductive pretreatments include hydrogen
rates for peeling, stopping, and hydrolysis. sulfide, ammonium sulfide, and sodium borohydride.9−12
The dissolution of xylan is governed mostly by alkaline Suitable pretreatments should meet the following criteria: (i)
solubility of the residual polymers because the uronic acid side They should have low cost (or efficient recoverability). (ii)
group protects against extensive peeling. As a result, more They should avoid nonprocess elements (i.e., include only H,
11572 https://doi.org/10.1021/acssuschemeng.1c04332
ACS Sustainable Chem. Eng. 2021, 9, 11571−11580
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

C, O, S, and Na).(iii) They should be reducing, avoiding the Assuming a yield loss on odw of 0.15%/kappa,4 these findings
formation of unsaturated bonds that can undergo further imply that methyl mercaptide addition would lead to a pulp
peeling.13 Although sodium borohydride is commonly used, it yield increase of about 2% at the same kappa value. In addition,
is expensive and adds the nonprocess element boron to the methyl mercaptide can be generated from spent kraft pulping
reaction conditions.12 Recently, sodium dithionite was liquor, greatly enhancing the commercializability of the
investigated but was unsuccessful in stabilizing mannose technology compared to prior proposed pretreatment tech-
under kraft pulping conditions.13 nologies.
Sodium methyl mercaptide can be prepared by dissolving Here, we use aqueous SMM as a pretreatment for wood
methyl mercaptan gas in an aqueous solution of sodium chips to achieve a a kraft pulp yield increase of about 2%−3%
hydroxide (Scheme 1). at optimal process conditions. The experiments to identify the
optimum conditions will be reported in a subsequent paper.
Scheme 1. Formation of Sodium Methyl Mercaptide We perform quantitative pulp product composition analysis
and show that SMM is an effective pretreatment for improving
carbohydrate yield, but that it is substantially less effective for
GGM than for xylan and cellulose. We then deploy molecular
modeling studies to investigate the molecular roots for these
Methyl mercaptide is a stronger nucleophile than HS−, and a varying degress of effectiveness. Quantum mechanical
similar attack on the methoxy groups of lignin by CH3S− investigation of the peeling and hydrolysis reactions reveal
produces dimethyl sulfide (CH3SCH3). Dimethyl disulfide slightly lower free energy barriers for GGM compared to
(CH3S2CH3) is then formed by oxidation of CH3SCH3. The cellulose for both types of degradation reactions. Molecular
beneficial effect of methyl mercaptide on carbohydrate dynamics simulations suggest that increased accessibility to
retention during kraft pulping has been reported in two reaction sites by hydroxide ions may render GGM more
patents. Tirado et al. demonstrated a 3% increase in pulp susceptible to degradative reactions under cooking conditions.


production (or about 1.5% based on the original oven dry
wood (odw) weight, assuming a pulp yield of about 50%)
when digester relief gases in a kraft pulp mill containing methyl
MATERIALS AND METHODS
mercaptide, methyl sulfide, and dimethyl sulfide were Pulping Experiments. Pulping experiments were performed
introduced into the fresh kraft cooking liquor, while using southern pine chips obtained from U.S. mills. All experiments
maintaining the same degree of delignification.14 More were performed in quadruplet using four 235 mL cylindrical rocking
digesters. Each digester contained approximately 30 g of wood chips
recently, a Finnish patent reported an increase in the pulp (on oven dry basis, o.d.). To avoid the handling of methyl mercaptan
yield of about 1% when sodium methyl mercaptide (SMM, gas, an aqueous solution of SMM (21% w/w) containing a small
Na+CH3S−) was used as an additive in polysulfide (PS) kraft amount of NaOH (0.4% w/w) at pH 12.9 was obtained from Arkema
pulping of pine, while the kappa number (a measure of the (King of Prussia, PA 19406). The SMM charge was 4.38% on wood
lignin content in pulp15) decreased from 27.9 to 21.7.16 (o.d. basis). The liquor-to-wood ratio for SMM pretreatment was 3

Figure 3. Atomistic models for MD simulations. (A) Model of galactoglucomannan (GGM) composed of three repeating units (one repeating unit
is shaded in red), each consisting of three mannose, one glucose, and one galactose residue. (B) Edge-on view of an example starting configuration
for a 36-chain cellulose microfibril (center, shown with green carbon atoms) surrounded by 18 GGM molecules (various colors). (C) Structures of
wild type glucose dimer (Glu-Glu, top) and mannose dimer (Man-Man, bottom). Note the difference in orientation of the hydroxy group at the O2
position between glucose and mannose. (D) Structures of pretreated Glu-Glu (top) and Man-Man (bottom) dimers, in which the C1 hydroxy has
been replaced by hydrogen to create an alkali-stable 1-deoxypyranose species.

11573 https://doi.org/10.1021/acssuschemeng.1c04332
ACS Sustainable Chem. Eng. 2021, 9, 11571−11580
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

L/kg, and the pretreatment temperature was between 105 to 130 °C. mesh Ewald method24 to calculate electrostatics accurately to long-
The initial pH of the pretreatment liquor was approximately 12. range. All systems were initially minimized and then density
Following pretreatment, the digesters were cooled to room equilibrated via 200 ps constant pressure simulations (NPT) at 300
temperature using a water bath, while makeup white liquor was K. Subsequently, production runs of 300 ns for the GGM/cellulose
being prepared. The final pH of the pretreatment liquor was between microfibril complex and 100 ns for the dimeric systems were
9.5 to 10.5. The total liquor-to-wood ratio after adding makeup white performed in the NVT ensemble.
liquor was 4.5 L/kg. The white liquor charge on wood was one of the Quantum Mechanical (QM) Model Construction. Similar to
process variables, expressed as effective alkali (EA) charge. EA charge the simplified models used in the classical MD simulations, we used a
is calculated as the weight of NaOH (as Na2O) plus half the weight of β-(1,4) glucose dimer (Glu-Glu) to represent cellulose in the QM
Na2S (as Na2O), reported as a percentage of the dry wood chip calculations. Differing from the MD model, β-(1,4)-linked Glu-Man
weight. After adding the makeup white liquor (sulfidity of 30% and was chosen as a model for GGM for the QM calculations of the
causticization efficiency of 80%), the digesters were placed in an oil alkaline hydrolysis reaction pathway because Man-Man is unable to
bath for 1 h at 115 °C to achieve uniform impregnation of white undergo epoxide formation due to the stereochemistry at C2.
liquor into the wood chips. At the end of impregnation, the digesters QM Method Validation. The dispersion-corrected hybrid meta-
were removed from the oil bath and cooled to room temperature, GGA density functional approximation M062X-D325 and the 6-31G
while the oil bath was heated to the cooking temperature of 170 °C. (d,p) basis set were used throughout the study. M062X-D3 was
The digesters were then immersed in the bath and cooked to a target chosen because it is among the most accurate density functionals for
H-factor of ∼1960 h. The H-factor is a measure of the pulping calculating conformational energies of carbohydrates,26 reaction
intensity, which takes into account the cooking time and temperature energies for small systems, and noncovalent interactions in the
compared to pulping at 100 °C (eq 1).4 GMTKN55 benchmarks.27 As an additional evaluation of this model

i 16115 yzz
expjjjj43.2 −
chemistry, we calculated the reaction free energy (ΔGr) and water-
zdt
T (K ) z{
t assisted activation free energy (Ea) for the ring opening of cyclic α-D-
k
H= ∫t 0 (1)
glucose to form acyclic D-glucose. We obtained ΔGr = 9.5 kcal/mol at
the M062X-D3/6-31g(d,p) level of theory, which is in good
Screened pulp was used to determine the kappa number (a agreement with the available reference value of 10.3 kcal/mol
measure of the lignin content of pulp, calculated according to TAPPI obtained at the CCSD(T)/G4 level of theory.28 Thus, we expect that
Standard T-236) and chemical composition. The carbohydrate the M062X-D3/6-31g(d,p) level of theory should provide sufficiently
composition of wood and pulp was determined using high- accurate energetics to investigate the model systems in the present
performance anion exchange chromatography (HPAEC) after acid work.
hydrolysis of the pulp sample. TAPPI standard methods (T-249, T- To account for solution-phase effects, we used the SMD polarizable
222, and T-204) were used to estimate the composition of lignin and continuum model29 with water as the solvent to calculate aqueous free
extractives in the wood/pulp samples. The uronic anhydride content energies. However, using the continuum solvent representation alone
was determined using the chromophoric group analysis method.17 to calculate hydration free energies of charged solutes can be
Residual alkali was determined using a modified SCAN-N 2:88 inaccurate because it neglects specific solute−solvent interactions.
method. Thus, we included one or two explicit water molecules that interact
Molecular Dynamics (MD) Model Construction. GGM is a directly with the solute to account for explicit hydrogen bonding.
heteropolymer consisting of uneven β-(1,4)-linked glucose and Vibrational frequencies were calculated for all optimized geometries
mannose as a backbone with some mannose units substituted by α- to confirm the presence of zero or one imaginary frequencies for
(1,6)-linked galactose. The molar ratio galactose:glucose:mannose in minima and transition states, respectively. Free energies were
GGM is approximately 1:1:3, and the atomic model used here for calculated at 80, 105, 115, and 130 °C. All QM calculations were
performed with Gaussian 16, revision A.03.30


GGM is shown in Figure 3A. To construct an atomistic model of
GGM bound to cellulose, a 36-chain cellulose Iβ microfibril with a
degree of polymerization (DP) of 20 was first generated with a RESULTS
diamond cross-sectional geometry (Figure 3B). Subsequently, 18 Experimental Kinetic Data for Real Wood Chips. We
GGM molecules with DP of 15 were then pseudorandomly arranged
around the cellulose microfibril, resulting in a cellulose/GGM ratio of investigated the effectiveness of 4.38% (w/w% on original
2.7, which is consistent with the experimentally observed cellulose/ wood) SMM pretreatment in preserving carbohydrate yield
GGM ratio of ∼2.6.18 from southern pine wood chips. The operating conditions were
In addition, because cellulose is a homopolymer consisting of adjusted to obtain a kappa number in the target range of 25−
repeating D-anhydroglucopyranose units linked by β-(1,4)-glycosidic 30. The kappa number, a measure of the pulp lignin content, in
bonds, we used a β-(1,4)-glycosidic bonded glucose dimer (Glu-Glu) this range is desirable because it is suitable for subsequent
to represent the cellulose for additional MD simulations (Figure 3C). bleaching.31 It is also desirable to stop the pulping process in
Because of the 1:3 molar ratio between glucose and mannose, we used this kappa range to avoid excessive degradation and dissolution
a β-(1,4)-linked mannose dimer (Man-Man) instead of a glucose- of the wood carbohydrates.
mannose dimer to represent GGM in the classical MD simulations
(Figure 3D). The dimer models were constructed, along with On the basis of literature data,32,33 about 0.18% methyl
pretreated forms of Glu-Glu and Man-Man dimers (i.e., with a 1- mercaptan (on-wood basis) is generated during kraft pulping
deoxypyranose terminus), for computational efficiency as well as to of softwood to produce a standard unbleached pulp. The
remove supramolecular effects given that cellulose forms microfibrillar amount of methyl mercaptan produced in a typical softwood
bundles whereas GGM does not. All generated models were solvated kraft cooking process, regardless of cooking temperature, is
with explicit TIP3P water molecules.19 When necessary in all about 0.235 mol/L of the spent black liquor.33 Using
simulations, charge neutrality was achieved by adding sodium cations. McKean’s data,32 which correlates the concentration of methyl
MD Simulation Details. All MD simulations were conducted mercaptide in the spent black liquor with rate of production of
using the program CHARMM20 with the CHARMM carbohydrate methyl mercaptan in the kraft process, it can be determined
force field, all36_carb.21 The DOMDEC fast parallel CHARMM
method22 was used to accelerate performance. All simulations used a that ∼0.18% methyl mercaptan (on-wood basis) is produced in
2 fs time step and the SHAKE algorithm23 to keep the length of the softwood kraft pulping process. Thus, 1.8 kg of methyl
covalent bonds to hydrogen fixed. Periodic solvated systems used a mercaptan per metric ton of oven-dry wood is equivalent to
nonbonded cutoff of 11 Å, with a 2 Å list buffer and heuristic list 0.18% (on-oven dry wood basis). This amount is significantly
update using a 1 Å switching function for dispersion and the particle- less than the 3% (on-wood basis) of methyl mercaptan
11574 https://doi.org/10.1021/acssuschemeng.1c04332
ACS Sustainable Chem. Eng. 2021, 9, 11571−11580
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

Table 1. Composition of Pulp Samples, Rejects, and Residual Alkali


Δ pulp
composition (%)b yield (%)
experimental conditionsa kappa residual EA
(pretreatment; impregnation; pulping) xylan GGM cellulose lignin calcc measd number % rejects (g/L Na2O)
0 southern pine chips 11.76 15.61 41.73 28.88 n/a 100 − − −
1 1 mM NaOH, 115 °C (ctrl); 115 °C; 15% EA 4.61 3.64 35.09 1.86 n/a n/a 29.4 ± 0.7 0.04 ± 0.02 7.7 ± 0.15
composition (%) rel to ctrl rel pulp
yield (%)
2 none; 115 °C; 16% EA 0.21 −0.04 −0.91 −0.36 −1.10 −1.10 25.1 ± 0.2 0.05 ± 0.00 9.3 ± 0.20
3 none; 115 °C; 17% EA −0.10 −0.11 −0.74 −0.55 −1.50 −1.50 22.7 ± 0.5 0.06 ± 0.05 10.6 ± 0.20
4 4.38% SMM 105 °C; 115 °C; 15% EA 0.20 0.29 −0.50 −0.39 −0.40 −0.40 19.1 ± 0.5 0.03 ± 0.02 9.5 ± 0.68
5 4.38% SMM 105 °C; 115 °C; 12% EA 0.96 0.23 1.07 0.24 2.49 2.50 29.0 ± 0.8 0.01 ± 0.01 6.2 ± 0.04
6 4.38% SMM 115 °C; 115 °C; 12% EA 0.87 0.12 1.85 0.06 2.90 2.90 30.0 ± 0.5 0.04 ± 0.00 5.6 ± 0.06
7 4.38% SMM 130 °C; 115 °C; 12% EA 1.31 0.11 0.31 0.38 2.10 2.10 29.7 ± 0.9 0.01 ± 0.01 5.8 ± 0.08
a
Pretreatment at condition 1 was carried out at 1 mM NaOH at pH 12 for 60 min. No pretreatment in conditions 2 and 3. Pretreatments at other
conditions were at pH 12 for 60 min unless otherwise noted. bBased on dry wood mass. cCalculated from pulp analysis data. dMeasured from pulp
yield data.

Figure 4. Effect of kappa number on total yield for the experiments included in Table 1. Seven experimental conditions were examined, each in
triplicate.

(equivalent to 4.38% as SMM) that was added in most of the and kappa number for the control cooks (Figure 4). Increasing
present experiments. Thus, additional mercaptan was needed EA to 16% (condition 2) and 17% (condition 3) resulted in a
in the present experiments to achieve 3% methyl mercaptan. reduction of the kappa number to 25.1 ± 0.2 and 22.7 ± 0.5,
Prior to pulping, we determined the composition of respectively. Thus, we selected condition 1 as the control.
southern pine wood chips to be 11.76% xylan, 15.61% Next, we considered the effects of SMM pretreatment and
GGM, 41.73% cellulose, and 28.88% lignin (Table 1, condition found that at 4.38% SMM charge it was possible to reduce the
0). Under pulping conditions of 1 mM NaOH, 115 °C, and pH EA charge during pulping from 15% to 12% (conditions 4 and
12 for 60 min, we obtained a kappa number of 29.4 ± 0.7 5, respectively) to obtain a kappa number similar to that of the
when the effective alkalinity (EA) was 15% (Table 1, condition control (condition 1). This approach led to a measured pulp
1). Under these conditions, the relative compositions of xylan, yield increase of 2.5% (on wood) for experiment 5 compared
GGM, and cellulose were reduced to 4.61, 3.64, and 35.09, to the control (Table 1 and Figure 4). For most of the
respectively, reflecting a loss of 60.79% of xylan, 76.68% of experiments, there is good agreement between the calculated
GGM, and 15.91% cellulose at 93.56% lignin removal. The and experimentally measured pulp yield increase (Table 1).
largest carbohydrate loss occurred for GGM. A significant The calculated pulp yield increase is the sum of the changes in
increase in GGM retention would have a major positive effect the percentages (on-wood basis) of xylan, GGM, cellulose, and
on the pulp yield. lignin, all relative to the control. The percentage of rejects was
We then analyzed the effect of changing the EA on kappa less than 0.1%, and the residual alkali was generally within the
number to identify pulping conditions that led to kappa acceptable range of 4−10 g/L.
numbers in the desired range (Table 1 and Figure 4). As The initial SMM pretreatment experiments were performed
expected, there is a linear relationship between total pulp yield at a charge of 4.38% SMM on o.d. wood for 60 min at 105 °C
11575 https://doi.org/10.1021/acssuschemeng.1c04332
ACS Sustainable Chem. Eng. 2021, 9, 11571−11580
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

Figure 5. Snapshot of the GGM/cellulose complex at the end of a 300 ns MD simulation. (A) End-on view and (B) side view rotated 90° with
cellulose microfibril shown in green surface representation, and bound GGM residues (determined as those within 3 Å of the cellulose surface)
shown in solid surface model of various colors. Unbound GGM residues are shown in stick representation. Hydroxide ions are shown as spheres
(red, oxygen; white, hydrogen). (C) Radial distribution function of OH− ions with respect to O1 and O2 at the reducing end of cellulose or GGM.
Results shown are with NaOH concentration of 0.9 M; 1.5 M NaOH loading showed qualitatively similar behavior.

(condition 4). The temperature of 105 °C was chosen Molecular Dynamics (MD) Simulations. Molecular
assuming that, at this temperature and pH ∼ 12 of the dynamics (MD) simulations were performed to determine
pretreatment liquor, the kinetics of the primary peeling the spatial distributions of ionic additives of the pretreatment
reactions of GGM and cellulose would be slower than the chemistry around each biomass polysaccharide. The inter-
kinetics of the reducing end group stabilization reactions by actions of these ions were probed with the unmodified forms
methyl mercaptide ions. Under these pretreatment conditions of cellulose and GGM as well as with the corresponding
with 15% EA, the kappa number dropped to 19.1. Therefore, polymers with terminal 1-deoxypyranoses expected to be
to obtain a kappa number that is closer to the target value we formed during pretreatment.
performed a 12% EA cook after pretreatment with 4.38% SMM Cellulose/GGM Simulations. To investigate the distribu-
(condition 5), which resulted in a kappa number of 29. The tion of OH− ions under cooking conditions, i.e., in the
results show that the addition of SMM increased the retention presence of NaOH, we performed a 300 ns simulation of a
of wood carbohydrates with an increase of ∼2.5% in the pulp GGM/cellulose microfibril complex (Figure 5A, B). Simu-
yield and the same kappa as the control. It is also important to lations were performed at both 0.9 and 1.5 M NaOH.
note that the 4.38% SMM pretreatment allowed a decrease of Although much more concentrated than the experiments
the EA charge in kraft cooking from 15% to 12%, while still described above, these conditions are consistent with past
achieving the same kappa number as the control kraft cook at investigations of GGM stabilization during kraft cooking.5 In
15% EA. Under these conditions, we observed increases in the addition, whereas the experimental conditions were chosen to
yields of xylan (0.96%) and cellulose (1.07%) but a much less maintain a balance between lignin removal and polysaccharide
significant increase in the GGM yield (0.23%). Details on the retention, the higher ion concentrations used in the
simulations resulted in a much larger data set from which to
calculations of the pulp composition will be reported
gather statistics. Simulations at both NaOH concentrations
separately.
indicate that OH− ions have a much higher propensity to bind
Next, we increased the temperature to 115 °C (condition 6).
to the reducing end groups of GGM than to cellulose, as
We observed a 1.85% increase in cellulose yield relative to the
revealed by comparing the radial distribution functions of each
control and a 0.12% increase in GGM yield. Further increasing
polymer (Figure 5C).
the temperature to 130 °C (condition 7) resulted in a slight Dimer Simulations under Cooking Conditions. A
reduction in the cellulose yield, likely due to increased peeling higher affinity of hydroxide ions for GGM than for cellulose
reactions. We hypothesized that the mercaptide ion stabiliza- (as deduced from the radial distribution functions) was also
tion reaction of cellulose occurs during the 60 min observed in 100 ns simulations of Glu-Glu and Man-Man
impregnation phase at 115 °C, but apparently the stabilization dimers (Figure S1). Glucose and mannose differ only in the
reaction of GGM by mercaptide is too slow compared to the stereochemistry at the C2 carbon, and our results demonstrate
GGM primary peeling reaction to affect the retention of GGM. that OH− ions exhibit a stronger propensity to localize at the
Because primary peeling of GGM starts at about 80 °C7 but bridging position between the O1−O2 atoms in mannose.
primary peeling of cellulose is only significant at or above 130 Analogous simulations with pretreated dimers indicate that
°C,5,6 it is possible that mercaptide ion stabilizes some OH− ions retain a stronger affinity to the hydroxy group in
cellulose before peeling but that GGM stabilization by mannose but in this case around the O2−O3 atoms.
mercaptide is slow compared to primary peeling. The degree Dimer Simulations under Pretreatment Conditions.
of xylan retention was mostly related to alkali charge because CH3S− ions form upon dissociation of sodium methyl
more xylan is dissolved at higher alkaline concentrations. The mercaptan and are thought to facilitate the conversion of
experiments with 4.38% SMM pretreatment at 130 °C reducing end groups to alkali-stable 1-deoxypyranoses and
(condition 7) resulted in a higher retention of xylan, possibly other products during the pretreatment reaction.
presumably because more alkali was consumed by cellulose Dimer simulations indicate that, similar to the behavior of
peeling, which in turn lowered the solubility of xylan in the HO−, CH3S− ions exhibit affinity for C1 in both Glu-Glu and
cooking liquor. Man-Man dimers (Figure 6). There is a slightly higher
11576 https://doi.org/10.1021/acssuschemeng.1c04332
ACS Sustainable Chem. Eng. 2021, 9, 11571−11580
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

Figure 6. Simulations of pretreated Glu-Glu and Man-Man dimers in SMM solution. Snapshots of CH3S− ions around C1 at the reducing end of
(A) a Glu-Glu dimer and (B) Man-Man dimer. The sulfur atom of each CH3S− ion within the first well in the RDF (approximately 4.25 Å of C1) is
shown every 100 ps of the 100 ns trajectory. (C) Radial distribution function of CH3S− ions with respect to C1 at the reducing end of a Glu-Glu
(green) or Man-Man (blue) dimer (CH3S− concentration = 1.5 M).

probability of CH3S− ions being located in close proximity to


C1 in Glu-Glu than in Man-Man.
MD simulations may also provide insight into the
susceptibility of cellulose and GGM polymers to alkaline
hydrolysis, which can lead to secondary peeling. Thus, we
measured the radial distribution of OH− ions with respect to
the glycosidic oxygen (O4) of cellulose and GGM. The
simulations indicate a much higher local concentration of OH−
ions around O4 in GGM than in cellulose (Figure 7).

Figure 8. Selected free energies for the primary peeling reaction of


Glu-Glu (representing cellulose) and Man-Man (representing GGM).
Free energies were calculated at 105 °C, and structures for the Glu-
Glu reactions are shown.

(Glu-Glu) and mannose dimer (Man-Man) to represent


cellulose and GGM, respectively.
Comparing the GGM and cellulose pathways, the corre-
sponding free energies for intermediates 1−4 are generally
within 1−2 kcal/mol of each other, indicating similar
thermodynamic favorability in the two systems. The overall
Figure 7. Radial distribution function of OH− ions at the glycosidic free energy barriers for the reactions were calculated as the
bond (O4 atom) in cellulose and GGM polymer MD simulations. difference in free energy between the lowest-energy state, 1,
and that of the transition state (TS-1). The calculated barrier is
slightly lower for Man-Man (22.9 kcal/mol) than for Glu-Glu
Quantum Mechanical (QM) Calculations of Peeling (26.5 kcal/mol).
and Hydrolysis Reactions. We performed quantum We then calculated reaction free energy diagrams for the
mechanical calculations to estimate the thermodynamic and generally accepted mechanism of alkaline hydrolysis of
kinetic feasibility of peeling and alkaline hydrolysis reactions. cellulose and GGM, with a particular emphasis on the
These two reactions were chosen to determine whether there presumed rate-limiting step of converting the O2-deprotonated
are differences in free energy barriers between the dimeric species to the epoxide intermediate, which then results in the
models representing cellulose and GGM. These calculations formation of a terminal α-D-glucose unit (Figure 9). We used
provide insight into detailed mechanisms that have not been β-(1,4)-linked Glu-Man as a model of GGM for the QM
tested or are unable to be tested experimentally. calculations of the alkaline hydrolysis reaction pathway because
Alkali-catalyzed endwise degradation, or primary peeling, is a Man-Man is unable to undergo epoxide formation due to the
major cause of the loss of carbohydrates during kraft pulping. stereochemistry at C2. The only difference between the two
The generally accepted mechanism of primary peeling involves models is in the stereochemistry of C2 in one sugar, so similar
ring opening, enolization, β-elimination, and benzylic acid results for the two systems were anticipated. We obtained free
rearrangement.4 Using this mechanism, we calculated free energy barriers of 40.8 kcal/mol for Glu-Glu and 38.7 kcal/mol
energies for the primary peeling reaction at the experimentally for Glu-Man (Figure 9). These values should be considered
relevant temperatures of 80, 105, 115, and 130 °C (Figure 8 essentially the same because they are within the expected error
and Figure S2). As was done for the classical MD simulations for the QM method used. However, it is possible that GGM
of dimeric models, we used a β-(1,4)-linked glucose dimer may be slightly more readily hydrolyzed than cellulose.
11577 https://doi.org/10.1021/acssuschemeng.1c04332
ACS Sustainable Chem. Eng. 2021, 9, 11571−11580
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

larger scale structural features such as crystallinity may play an


important role in governing accessibility of reactive sites and
therefore peeling susceptibility. In fact, our classical MD
simulations of cellulose and GGM indicate that OH− ions have
a greater propensity to localize near the reducing end groups of
GGM than in cellulose. This difference in affinity could be due
to the chemical differences between cellulose and GGM or to
the crystalline nature of cellulose possibly rendering physical
protection against peeling. We posit that both effects are likely
at work, on the basis of these polymer-scale results as well as
our findings from dimer models. MD simulations of a mannose
dimer reveal an enhanced propensity for OH− ions to locate at
the bridging position between the O1 and O2 atoms at the
reducing end. This finding indicates that, even without a
polymer chain and in the absence of supramolecular structure
Figure 9. Selected free energies for the alkaline hydrolysis reaction of (i.e., a cellulose microfibril), peeling of mannose from GGM
Glu-Glu (representing cellulose) and Glu-Man (representing GGM).
Free energies were calculated at 105 °C, and structures for the Glu-
may be promoted more so than glucose from cellulose. This
Glu reactions are shown. finding holds even in analogous simulations wherein the
reducing end group is converted to an alkali-stable 1-

■ DISCUSSION AND CONCLUSIONS


Wood remains the largest cost in the manufacture of wood
deoxypyranose functionality. The enhanced affinity is possibly
due to cis-OH groups at these two carbon ring atoms and may
promote both the peeling reaction in the WT Man-Man dimer
pulp for paper, making up over half of the noncapital operating and the dehydration reaction in the pretreated Man-Man
cost of producing pulp.34 Therefore, increasing pulp yield dimer.
would have a major financial impact on pulp mill economics. In addition to primary peeling, internal chain scission by
Unfortunately, and despite intensive R&D efforts, attempts to alkaline hydrolysis leads to the creation of new chain ends that
increase kraft pulp yield by an industry-relevant process have can subsequently undergo secondary peeling. Thus, even
not yet been successful. Here, we have demonstrated a 3% though SMM pretreatment can protect REGs against primary
increase in pulp yield from softwood via pretreatment with peeling, new unprotected REGs may be generated during
SMM. The primary means by which pulp yield is increased is pulping. Our DFT calculations on dimeric models indicate that
in the stabilization of cellulose against degradation by alkaline GGM may be slightly more susceptible to internal alkaline
peeling (Table 1). Unfortunately, stabilization of the primary hydrolysis of a carbohydrate linkage than cellulose, although
softwood hemicellulose galactoglucomannan (GGM) by this the calculated free energy barriers are within ∼2 kcal/mol of
pretreatment is minimal, although the cause for the each other. The DFT-derived energy barriers can be compared
discrepancy between the two polysaccharides is not immedi- qualitatively with previously reported experimental values. Our
ately clear. GGM model for alkaline hydrolysis yielded a substantially
To obtain a better fundamental understanding of the higher value (38.7 kcal/mol) than the results reported by
molecular basis for the experimentally observed differences in Montagna et al. (23.7 kcal/mol)7 and by Nieminen et al. (14.9
stabilization of cellulose and GGM by SMM pretreatment, we kcal/mol).6 The origin of the large discrepancies among these
carried out a computational analysis of relevant model systems. values is unclear. In contrast, our calculated value of 40.8 kcal/
Numerous kinetic models have been developed to predict mol for alkaline hydrolysis with the cellulose model is quite
carbohydrate degradation during pulping.5,6,35,36 While useful close to the corresponding value from kinetic modeling of 41.8
for predicting the effect of various kinetic parameters on kcal/mol.6 Overall, our calculations follow the same trend in
product distributions, such models do not provide the reactivity; i.e., the barrier for cellulose hydrolysis is higher than
molecular level insight critical to developing actionable that for GGM.
hypotheses for pulp yield enhancement. Molecular dynamics Finally, we reiterate that the 2.5% pulp yield increase
studies have investigated various aspects of carbohydrate obtained with the addition of 3% methyl mercaptan (on o.d.
conformations and dynamics in aqueous solution at the wood chips, equivalent to 4.38% as SMM) is rather modest.
molecular scale,23,37−39 but detailed molecular-scale insight However, it has been proposed that a yield increase of 12% for
under kraft pulping conditions has not been demonstrated. softwood (4% cellulose and 8% galactoglucomannan) could be
Quantum mechanical results presented here indicate a achieved if primary peeling could be eliminated by the addition
slightly lower peeling barrier for Man-Man compared to Glu- of methyl mercaptan. In addition, during regular softwood kraft
Glu. The DFT-derived energy barriers can be compared with pulping, about 0.2% methyl mercaptan (on wood) is
experimentally derived values for cellulose and GGM peeling, formed.32,33 Thus, methyl mercaptan is already part of normal
calculated from experiments as 23.4 and 6.2 kcal/mol, industrial operations, albeit at lower levels than those used in
respectively.6 The latter value was noted in the original the present experiments. As a result, implementation of the
publication to be surprisingly lower than values from previous concept would require emission control techniques similar to
work.5 The value computed for GGM peeling (27.7 kcal/mol) what is presently done in modern “odorless” kraft mills. Of
by Montagna et al.7 is much more in line with the results course, the methyl mercaptan would also have to be
presented here. The lower peeling barrier for the mannose regenerated from the spent pulping so that addition of makeup
dimer suggests that more successive peeling events will occur methyl mercaptan would be kept to a minimum in a closed-
for GGM than for cellulose. However, we note that the peeling loop process; recycling of methyl mercaptan is a topic that will
rates will also be affected by the whole polymer context, and be described in a subsequent publication. The primary
11578 https://doi.org/10.1021/acssuschemeng.1c04332
ACS Sustainable Chem. Eng. 2021, 9, 11571−11580
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

motivation for the present work, and especially the MD and Author Contributions

QM calculations, is to provide a better understanding of why C. J. Cooper, R. Patil, and L. Bu contributed equally.
methyl mercaptan addition produced only a pulp yield increase Notes
of 2.5% and that this fundamental insight might allow The authors declare no competing financial interest.


identification of conditions that would minimize or even
suppress primary peeling of galactoglucomannan.
In conclusion, we have demonstrated that pretreatment of ACKNOWLEDGMENTS
softwood with sodium methyl mercaptan can increase kraft This research was supported by the High-Performance
pulp yield by 3%, while maintaining a kappa value in the Computing for Manufacturing Project Program (HPC4Mfg),
desirable range. This result is thought to be achieved by the which is managed by the U.S. Department of Energy (DOE),
capping cellulose REGs with alkali-stable functionality that Advanced Manufacturing Office, within the Energy Efficiency
protects against peeling. Hemicellulose GGM is left largely and Renewable Energy Office. This work was partially
unprotected, representing a further target for increasing pulp authored by the Alliance for Sustainable Energy, LLC, the
yield. The results from atomistic MD simulations and DFT manager and operator of NREL for the U.S. Department of
calculations indicate that the molecular roots for these Energy (DOE) under Contract No. DE-AC36-08GO28308.
observations include both the chemical differences between The views expressed in the article do not necessarily represent
mannose and glucose as well as differences in the supra- the views of the DOE or the U.S. Government. The U.S.
molecular structures, namely, that cellulose forms microfibrils Government retains and the publisher, by accepting the article
whereas GGM does not. This molecular insight may be useful for publication, acknowledges that the U.S. Government
for guiding future strategies for increased preservation of wood retains a nonexclusive, paid-up, irrevocable, worldwide license
carbohydrates during kraft pulping. to publish or reproduce the published form of this work, or


*
ASSOCIATED CONTENT
sı Supporting Information
allow others to do so, for U.S. Government purposes.
Computer time was provided by the National Renewable
Energy Laboratory Computational Sciences Center supported
The Supporting Information is available free of charge at by the DOE Office of EERE under Contract No. DE-AC36-
https://pubs.acs.org/doi/10.1021/acssuschemeng.1c04332. 08GO28308. This research also used resources of the
Compute and Data Environment for Science (CADES) at
Additional figures (PDF) ORNL, which is managed by UT-Battelle, LLC, for the U.S.
Calculated energies (XLSX) DOE under Contract No. DE-AC05-00OR22725.


Cartesian coordinates for all optimized geometries (ZIP)

■ AUTHOR INFORMATION
Corresponding Authors
REFERENCES
(1) State Industry Economic Impact Report; American Forest & Paper
Association: January 2021.
Adriaan van Heiningen − Department of Chemical and (2) Pulp and Paper Capacities − Survey; FAO: Rome, 2020.
Biomedical Engineering, University of Maine, Orono, Maine (3) Smook, G. A. Handbook for Pulp & Paper Technologists, 3rd ed.;
04469-5737, United States; Email: adriaan.vanheiningen@ Angus Wilde Publications, 2002; Vol. 75.
maine.edu (4) Sixta, H. Handbook of Pulp; Wiley-VCH Verlag GmbH & Co.
Jerry M. Parks − Biosciences Division, Oak Ridge National KGaA: Weinheim, Germany, 2006; pp 3−20.
(5) Paananen, M.; Tamminen, T.; Nieminen, K.; Sixta, H.
Laboratory, Oak Ridge, Tennessee 37831-6309, United Galactoglucomannan stabilization during the initial kraft cooking of
States; orcid.org/0000-0002-3103-9333; Scots pine. Holzforschung 2010, 64 (6), 683−692.
Email: parksjm@ornl.gov (6) Nieminen, K.; Paananen, M.; Sixta, H. Kinetic Model for
Brandon C. Knott − Renewable Resources and Enabling Carbohydrate Degradation and Dissolution during Kraft Pulping. Ind.
Sciences Center, National Renewable Energy Laboratory, Eng. Chem. Res. 2014, 53 (28), 11292−11302.
Golden, Colorado 80401, United States; orcid.org/0000- (7) Montagna, P. N.; Inalbon, M. C.; Paananen, M.; Sixta, H.;
0003-3414-3897; Email: brandon.knott@nrel.gov Zanuttini, M. A. Diffusion Dynamics in Pinus sylvestris Kraft
Impregnation: Effect of Deacetylation and Galactoglucomannan
Authors Degradation. Ind. Eng. Chem. Res. 2013, 52 (10), 3658−3662.
Connor J. Cooper − Biosciences Division, Oak Ridge National (8) Matthews, C. H. Carbohydrate losses at high temperature in
Laboratory, Oak Ridge, Tennessee 37831-6309, United kraft pulping. Svensk Papperstidn 1974, 77 (17), 629−635.
States; orcid.org/0000-0002-5527-9948 (9) Hartler, N.; Olson, L. A. Hydrogen Sulfide Cooking Part I. First
Ravikant Patil − Department of Chemical and Biomedical Stage Variables. Svensk Papperstidn 1972, 75 (13), 559−565.
(10) Pekkala, O.; Palenius, I. Hydrogen sulfide pretreatment in
Engineering, University of Maine, Orono, Maine 04469-5737, alkaline pulping. Paperi ja Puu 1973, 55 (9), 659−668.
United States (11) Chiang, V. L.; Sarkanen, K. V. Ammonium sulfide organosolv
Lintao Bu − Renewable Resources and Enabling Sciences pulping. Wood Sci. Technol. 1983, 17, 217−226.
Center, National Renewable Energy Laboratory, Golden, (12) Wang, Y.; Azhar, S.; Lindström, M. E.; Henriksson, G.
Colorado 80401, United States Stabilization of polysaccharides during alkaline pre-treatment of wood
Zhongyu Mou − Biosciences Division, Oak Ridge National combined with enzyme-supported extractions in a biorefinery. J. Wood
Laboratory, Oak Ridge, Tennessee 37831-6309, United Chem. Technol. 2015, 35 (2), 91−101.
States; orcid.org/0000-0003-2240-3129 (13) Lindén, P. A.; Lindström, M. E.; Lawoko, M.; Henriksson, G.
David B. Turpin − Alliance for Pulp and Paper Technology Stabilising mannose using sodium dithionite at alkaline conditions.
Innovation (APPTI), Washington, DC 20005, United States Holzforschung 2020, 74 (2), 131−140.
(14) Tirado, A.; Guevara, M.; Gonzalez, V. Method of digesting
Complete contact information is available at: wood by the kraft process utilizing organic sulphides. U.S. Patent
https://pubs.acs.org/10.1021/acssuschemeng.1c04332 US3,451,889, 1969.

11579 https://doi.org/10.1021/acssuschemeng.1c04332
ACS Sustainable Chem. Eng. 2021, 9, 11571−11580
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

(15) TAPPI Standard T236 cm-85: Kappa Number of Pulp; TAPPI: Rendell, A. P.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.;
Atlanta, GA, 1993. Millam, J. M.; Klene, M; Adamo, C.; Cammi, R.; Ochterski, J. W.;
(16) Vaelttilae, O.; Pekkala, O. Method for making chemical Martin, R. L.; Morokuma, K.; Farkas, O.; Foresman, J. B.; Fox, D. J.
cellulose pulp by a polysulfide cooking process with the addition of Gaussian 16, Revision A.03; Gaussian, Inc.: Wallingford, CT, 2016.
organic mercaptans. Finland Patent FI118347B20071015, 2007. (31) Annergren, G.; Bowman, G. M.; Sandstrom, P. Principles of
(17) Scott, R. W. Colorimetric determination of hexuronic acids in Multi-stage Bleaching of Softwood Kraft Pulp. In Proceedings of the
plant materials. Anal. Chem. 1979, 51 (7), 936−941. International Pulp Bleaching Conference, Helsinki, Finland, 1998.
(18) Negahdar, L.; Delidovich, I.; Palkovits, R. Aqueous-phase (32) McKean, W. T.; H, B. F.; Sarkanen, K. V. Kinetic analysis of
hydrolysis of cellulose and hemicelluloses over molecular acidic odor formation in the kraft pulping process. Tappi J. 1965, 48 (12),
catalysts: Insights into the kinetics and reaction mechanism. Appl. 699−704.
Catal., B 2016, 184, 285−298. (33) McKean, W. T.; H, B. F.; Sarkanen, K. V.; Price, L.; Douglass, I.
(19) Jorgensen, W. L.; Chandrasekhar, J.; Madura, J. D.; Impey, R. B. Effect of kraft pulping conditions on the formatino of methyl
W.; Klein, M. L. Comparison of simple potential functions for mercaptan and dimethyl sulfide. Tappi J. 1967, 50 (8), 400−405.
simulating liquid water. J. Chem. Phys. 1983, 79 (2), 926−935. (34) Kangas, P.; Kaijaluoto, S.; Määttänen, M. Evaluation of future
(20) Brooks, B. R.; Brooks, C. L., 3rd; Mackerell, A. D., Jr.; Nilsson, pulp mill concepts − Reference model of a modern Nordic kraft pulp
L.; Petrella, R. J.; Roux, B.; Won, Y.; Archontis, G.; Bartels, C.; mill. Nord. Pulp Pap. Res. J. 2014, 29 (4), 620−634.
Boresch, S.; Caflisch, A.; Caves, L.; Cui, Q.; Dinner, A. R.; Feig, M.; (35) Fearon, O.; Nykänen, V.; Kuitunen, S.; Ruuttunen, K.; Alén, R.;
Fischer, S.; Gao, J.; Hodoscek, M.; Im, W.; Kuczera, K.; Lazaridis, T.; Alopaeus, V.; Vuorinen, T. Detailed modeling of the kraft pulping
Ma, J.; Ovchinnikov, V.; Paci, E.; Pastor, R. W.; Post, C. B.; Pu, J. Z.; chemistry: carbohydrate reactions. AIChE J. 2020, 66 (8), e16252.
Schaefer, M.; Tidor, B.; Venable, R. M.; Woodcock, H. L.; Wu, X.; (36) da Silva Perez, D.; van Heiningen, A. Prediction of alkaline
Yang, W.; York, D. M.; Karplus, M. CHARMM: the biomolecular pulping yield: equation derivation and validation. Cellulose 2015, 22
simulation program. J. Comput. Chem. 2009, 30 (10), 1545−614. (6), 3967−3979.
(21) Guvench, O.; Greene, S. N.; Kamath, G.; Brady, J. W.; Venable, (37) Peric-Hassler, L.; Hansen, H. S.; Baron, R.; Hunenberger, P. H.
R. M.; Pastor, R. W.; Mackerell, A. D., Jr. Additive empirical force Conformational properties of glucose-based disaccharides investigated
field for hexopyranose monosaccharides. J. Comput. Chem. 2008, 29 using molecular dynamics simulations with local elevation umbrella
(15), 2543−2564. sampling. Carbohydr. Res. 2010, 345 (12), 1781−801.
(22) Hynninen, A. P.; Crowley, M. F. New faster CHARMM (38) Berglund, J.; Angles d’Ortoli, T.; Vilaplana, F.; Widmalm, G.;
molecular dynamics engine. J. Comput. Chem. 2014, 35 (5), 406−413. Bergenstrahle-Wohlert, M.; Lawoko, M.; Henriksson, G.; Lindstrom,
(23) Krautler, V.; Muller, M.; Hunenberger, P. H. Conformation, M.; Wohlert, J. A molecular dynamics study of the effect of glycosidic
dynamics, solvation and relative stabilities of selected beta- linkage type in the hemicellulose backbone on the molecular chain
hexopyranoses in water: a molecular dynamics study with the flexibility. Plant J. 2016, 88 (1), 56−70.
GROMOS 45A4 force field. Carbohydr. Res. 2007, 342 (14), 2097− (39) Berglund, J.; Azhar, S.; Lawoko, M.; Lindström, M.; Vilaplana,
2124. F.; Wohlert, J.; Henriksson, G. The structure of galactoglucomannan
(24) Darden, T.; York, D.; Pedersen, L. Particle mesh Ewald: An N· impacts the degradation under alkaline conditions. Cellulose 2019, 26
log(N) method for Ewald sums in large systems. J. Chem. Phys. 1993, (3), 2155−2175.
98 (12), 10089−10092.
(25) Zhao, Y.; Truhlar, D. G. The M06 suite of density functionals
for main group thermochemistry, thermochemical kinetics, non-
covalent interactions, excited states, and transition elements: two new
functionals and systematic testing of four M06-class functionals and
12 other functionals. Theor. Chem. Acc. 2008, 120 (1−3), 215−241.
(26) Marianski, M.; Supady, A.; Ingram, T.; Schneider, M.; Baldauf,
C. Assessing the accuracy of across-the-scale methods for pedicting
carbohydrate conformational energies for the examples of glucose and
alpha-maltose. J. Chem. Theory Comput. 2016, 12 (12), 6157−6168.
(27) Goerigk, L.; Hansen, A.; Bauer, C.; Ehrlich, S.; Najibi, A.;
Grimme, S. A look at the density functional theory zoo with the
advanced GMTKN55 database for general main group thermochem-
istry, kinetics and noncovalent interactions. Phys. Chem. Chem. Phys.
2017, 19 (48), 32184−32215.
(28) Assary, R. S.; Curtiss, L. A. Comparison of sugar molecule
decomposition through glucose and fructose: a high-level quantum
chemical study. Energy Fuels 2012, 26 (2), 1344−1352.
(29) Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. Universal
solvation model based on solute electron density and on a continuum
model of the solvent defined by the bulk dielectric constant and
atomic surface tensions. J. Phys. Chem. B 2009, 113 (18), 6378−6396.
(30) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.;
Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Petersson,
G. A.; Nakatsuji, H.; Li, X.; Caricato, M.; Marenich, A. V.; Bloino, J.;
Janesko, B. G.; Gomperts, R.; Mennucci, B.; Hratchian, H. P.; Ortiz, J.
V.; Izmaylov, A. F.; Sonnenber, J. L.; Williams-Young, D.; Ding, F.;
Lipparini, F.; Egidi, F.; Goings, J.; Peng, B.; Petrone, A.; Henderson,
T.; Ranasinghe, D.; Zakrzewsk, V. G.; Gao, J.; Rega, N.; Zheng, G.;
Liang, W.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.;
Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H; Vreven, T.;
Throssell, K.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.;
Bearpark, M. J.; Heyd, J. J.; Brothers, E. N..; Kudin, K. N.; Staroverov,
V. N.; Keith, T. A.; Kobayashi, R.; Normand, J.; Raghavachari, K.;

11580 https://doi.org/10.1021/acssuschemeng.1c04332
ACS Sustainable Chem. Eng. 2021, 9, 11571−11580

You might also like