You are on page 1of 11

International Journal of Biological Macromolecules 256 (2024) 128506

Contents lists available at ScienceDirect

International Journal of Biological Macromolecules


journal homepage: www.elsevier.com/locate/ijbiomac

Review

A review on the calculation and application of lignin Hansen


solubility parameters
Qingzhi Ma a, *, Changqing Yu b, Yuran Zhou c, Dinggen Hu c, Jianbin Chen a, c, Xuejin Zhang a, *
a
Key Laboratory of Recycling and Eco-treatment of Waste Biomass of Zhejiang Province, Zhejiang University of Science and Technology, Hangzhou 310023, China
b
Provincial Key Laboratory of New Polyolefin Materials, School of Chemistry & Chemical Engineering, Northeast Petroleum University, Daqing 163318, China
c
Winbon Schoeller New Materials Co., Ltd., Quzhou 324400, China

A R T I C L E I N F O A B S T R A C T

Keywords: Hansen solubility parameters (HSPs) play a critical role in the majority of processes involving lignin depoly­
Hansen solubility parameters merization, separation, fractionation, and polymer blending, which are directly related to dissolution properties.
Atomic and functional group contribution However, the calculation of lignin HSPs is highly complicated due to the diversity of sources and the complexity
(AFGC) method
of lignin structures. Despite their important role, lignin HSPs have been undervalued, attracting insufficient
Lignin dissolution behavior
attention. This review summarizes the calculation methods for lignin HSPs and proposes a straightforward
method based on lignin subunits. Furthermore, it highlights the crucial applications of lignin HSPs, such as
identifying ideal solvents for lignin dissolution, selecting suitable solvents for lignin depolymerization and
extraction, designing green solvents for lignin fractionation, and guiding the preparation of lignin-based com­
posites. For instance, leveraging HSPs to design a series of solvents could potentially achieve sequential
controllable lignin fractionation, addressing issues of low value-added applications of lignin resulting from poor
homogeneity. Notably, HSPs serve as valuable tools for understanding the dissolution behavior of lignin.
Consequently, we expect this review to be of great interest to researchers specializing in lignin and other
macromolecules.

1. Introduction product upgraded oxygen-rich bio-oils can be obtained via catalysis


[13,14], (4) preparation of lignin-based composites, including poly­
Lignin, characterized as a heterogeneous polymer with a phenyl­ urethane, phenolic resin, and aldehyde free adhesive [15–18], and (5)
propane structure, is the sole polymer capable of naturally supplying a development of lignin nanoparticles for several applications, such as
significant quantity of phenolic compounds [1,2]. Depending on the adhesives, composites, coatings, and antibacterial agents. [19–21].
number of methoxyl in the benzene ring, the phenylpropane structure is However, despite its broad utility, only approximately 5 % of industrial
subdivided into guaiacyl (G), syringyl (S), and p-hydroxyphenyl (H) lignin is utilized in high value applications, with the remaining 95 %
units [3]. The side chain of the phenylpropane structure contains burned to recover heat [22–24].
different types and numbers of functional groups, such as hydroxyl, The complex structure, heterogeneity, and poor solubility of lignin
carboxyl, carbonyl, acyl, and olefin structures [4]. The extraction, (except for cases involving alkaline solutions and lignosulfonate) pose
characterization, and utilization of lignin have attracted considerable significant challenges, hindering its high value-added applications
attention owing to its renewability and versatility [5,6]. Lignin has [25–27]. Of these, the correlation and compatibility (insolubility,
demonstrated rich application value [7]: (1) unmodified or modified swelling, dissolution, compatibility, and incompatibility) of lignin with
(phosphating, hydroxymethylation, phenolation, sulfonation, carbox­ solvents (or other polymers) profoundly impact its extraction, separa­
ylation, epoxidation, esterification, and etherification) lignin can be tion, purification, modification, fractionation, and other various utili­
directly used as a dispersant, emulsifier, sustained-release agent, sur­ zation processes. However, the importance of the correlation and
factant, and energy storage material [8,9], (2) lignin can be depoly­ compatibility behaviors, especially the dissolution behavior of lignin,
merized, encompassing physical, biological, chemical, and pyrolysis has not received adequate attention [28,29]. Dissolution is the process
methods, into smaller molecular platform compounds [10–12], (3) of mixing multiple substances into a homogeneous phase in a molecular

* Corresponding authors.
E-mail addresses: maqingzhi@zust.edu.cn (Q. Ma), xuejinzhang@126.com (X. Zhang).

https://doi.org/10.1016/j.ijbiomac.2023.128506
Received 30 August 2023; Received in revised form 24 November 2023; Accepted 28 November 2023
Available online 29 November 2023
0141-8130/© 2023 Elsevier B.V. All rights reserved.
Q. Ma et al. International Journal of Biological Macromolecules 256 (2024) 128506

state. The formation of a solution is a gradual process in which energy is when ΔHM is positive. Furthermore, there is no database available for
consumed to overcome the attraction between solute-solute and solvent- lignin mixing heat. Directly measuring mixing heat by combining lignin
solvent, which is an endothermic process. Conversely, in the solvation and solvent would provide explicit information on whether the lignin is
process, where solute-solvent attraction force is established, heat is soluble. Therefore, utilizing ΔG to determine the dissolution behavior of
released, which is an exothermic process. Following dissolution, an lignin is more complicated than conducting a direct experiment.
equilibrium is attained by the intermolecular forces of the solute-solute, Consequently, alternative tools are required to determine the solubility
solute-solvent, and solvent-solvent systems. However, the dissolution of solutes in solvents.
behavior of polymers like lignin is notably more complex compared to
small molecules, as it necessitates adequate swelling and dispersion
before the occurrence of dissolution [30]. 2.2. Flory-Huggins theory and interaction parameter χ
The overall (total) solubility parameter (δ), defined as the square
root of cohesive energy (E) density, was proposed to aid in under­ The Flory–Huggins theory plays a vital role in assessing the mutual
standing the compatibility and dissolution behavior of substances miscibility between a polymer and a solvent or another polymer. This
[31–33]. The cohesion energy of the solute and solvent is derived from theory was derived from the following three assumptions [37–39]: (1)
the energy of evaporation, which is the total energy required to hold each polymer chain is represented as a chain comprising “x” segments,
liquid molecules together. Hansen divided the total energy of vapor­ with each segment having exactly the same volume as a solvent mole­
ization of a substance into three parts: dispersion forces, permanent cule; (2) the polymer chain is completely free, and all conformations
dipole–permanent dipole forces, and hydrogen bonding [32,33]. possess the same energy; and (3) the polymer segments in solution
Furthermore, he proposed the concept of Hansen solubility parameters exhibit uniformly distribution, occupying any lattice equally. Therefore,
(HSPs), suggesting that the overall solubility parameter of a substance a polymer solution can be depicted by a lattice divided into cells, with
comprises three partial solubility parameters: dispersive cohesion (δD), each cell occupying a segment of the polymer or solvent molecule.
polar cohesion (δP), and hydrogen bonding cohesion (δH) [34]. Ac­ Therefore, the Flory-Huggins interaction parameter (χ12) between the
cording to Hansen’s theory, two substances are more compatible when polymer and the solvent is defined by Eq. (2) [38]:
their three forces (or parameters) are similar. HSPs have found extensive Z△w12
application in dissolution-related fields, including but not limited to the χ 12 = , (2)
RT
coating industry (for solvent selection), prediction of polymer compat­
ibility and solubility, assessing chemical resistance and permeation where Z represents the coordination number of the crystalline network
rates, and surface characterization of pigments, fibers, and fillers [35]. model, Δw12 signifies the mutual interaction energy between polymer
However, to the best of our knowledge, no study has systematically and solvent molecules, R denotes the universal gas constant, and T is the
introduced calculation methods for lignin HSPs and their comprehensive absolute temperature.
applications. This concise review aims to: 1) summarize polymer The χ12 is a dimensionless parameter, with χ12kT expressing the
dissolution-related theories, such as the Gibbs free energy of dissolution, difference in energy of a solvent molecule in pure solvent compared to
Flory-Huggins interaction parameter, and HSPs, 2) offer an overview of its immersion in pure polymer. The osmotic pressure (π) generated when
the essential methods for calculating lignin HSPs and propose a quick mixing a pure solvent with a polymer solution can be expressed by Eq.
method to compute the partial and overall solubility parameters of (3). Accordingly, the following conclusions can be drawn from further
lignin based on its structure; and 3) outline the critical applications of analysis of Eq. (3) (the detailed derivation is summarized in the refer­
lignin HSPs, aiming to more effectively guide the selection of solvents ence and omitted here) [33,40,41]: When χ12 = 1/2, the polymer ex­
during the lignin extraction and fractionation processes, as well as hibits ideal behavior during mixing with the solvent, referred to as the
predict the compatibility of lignin with other polymers. θ-point. Under these conditions, the polymer chains in the solution have
no repulsion or attraction, maintaining their original dimensions un­
2. Critical theories and factors influencing polymer dissolution disturbed. In instances where χ12 < 1/2, the polymer chains unfold in
behavior the solvent, indicating favorable or “good” solvent conditions. In
contrast, when χ12 > 1/2, the polymer chains fold or precipitate in the
2.1. Gibbs free energy of dissolution solvent, signifying unfavorable or “poor” solvent conditions. The Flory-
Huggins interaction parameter χ12 substantially contributes to under­
The changes in Gibbs free energy (ΔG) play a fundamental role in standing the dissolution behavior of polymers in various solvents.
thermodynamics, determining whether the reactions, including disso­ [ ( ) ]
(π/C2 ) = RT (1/M2 ) + ϕ2 2 (1/2 − χ 12 )C2 , (3)
lution, can occur spontaneously. According to Eq. (1) [36], the disso­
lution process can only occur spontaneously when the free energy of
where C2 denotes the polymer concentration, M2 represents the molar
mixing (ΔGM) is either zero or negative.
mass of the polymer, and ϕ2 signifies the volume fraction.
△GM = △HM − T△SM , (1) However, the Flory–Huggins theory does not consider the entropy
caused by polymer chain folding [42]. When a crystalline polymer dis­
where ΔHM represents the heat of mixing, T is the absolute temperature, solves, its crystal structure undergoes changes. Even for amorphous
and ΔSM denotes the change in entropy during the mixing process. polymers, the conformation of the molecular chains transforms, leading
Entropy is essentially a description of the degree of chaos, increasing to additional entropy and energy variations. Moreover, the actual dis­
as the solute dissolves in the solvent, which means that ΔSM is always tribution of polymers in the solution is not uniform, further complicating
positive. Considering the absolute temperature is positive, TΔSM and the scenario.
–TΔSM are always positive and negative, respectively. When the disso­
lution is exothermic, where both ΔHM and ΔGM are negative, the
2.3. HSPs
dissolution can occur spontaneously. Likewise, in a dissolution process
without heat exchange, wherein ΔHM = 0 and ΔGM is negative, the
The concept of solubility parameter was initially proposed by Hil­
dissolution can also occur spontaneously. However, in endothermic
debrand and Scott [31,43], defined as the square root of cohesive energy
dissolution (ΔHM is positive), determining whether the ΔGM value is
density, as shown in Eq. (4). The one-component overall solubility
positive or negative becomes uncertain. Therefore, employing ΔG alone
parameter formulated by Hildebrand and Scott is generally applicable to
to ascertain lignin solubility in solvents has limitations, particularly
regular solutions lacking molecular or specific interactions.

2
Q. Ma et al. International Journal of Biological Macromolecules 256 (2024) 128506

δ = (E/V)1/2 , (4) δ = (ED /V + EP /V + EH /V)1/2 , (6)

where E is the total cohesion energy (or the energy of vaporization) and ( )1/2
δ = δD 2 + δP 2 + δH 2 . (7)
can be measured. V denotes the molar volume of both the pure solvent
and polymer. Notably, the numerical value of the solubility parameter in The dispersion force, or London force, is a weak intermolecular force
MPa1/2 is 2.0455 times greater than that in cal1/2⋅cm–3/2. existing between two atoms or molecules when they are in close prox­
A shortcoming of the solubility parameter introduced by Hildebrand imity. It arises due to the transient dipoles present in molecules or atoms.
and Scott is that the approach is limited to regular solutions and must Every molecule (or atom) has a transient dipole, which induces a similar
consider the influences of polarity and hydrogen-bonding interactions. transient dipole in a neighboring molecule (or atom), causing an
Blanks and Prausnitz [44] addressed this by categorizing solubility pa­ attractive force between them. Dispersion forces are universal and exist
rameters into two parts: “nonpolar” and “polar.” Hansen expanded upon in all molecules and atoms. Conversely, polar cohesive energy is pri­
this concept by dividing the total vaporization energy of a liquid into marily induced by permanent dipole-permanent dipole interactions.
three constituent parts: dispersion forces, permanent dipole–permanent Similar to permanent dipole-permanent dipole interactions, polar
dipole forces, and hydrogen bonding. Alternatively, the total cohesive cohesive energy is present in most molecules. Hydrogen bonding is an
energy can be considered to comprise three components: dispersion intermolecular interaction that resembles the polar cohesive energy and
cohesive (ED), polar cohesive (EP), and hydrogen bonding (EH), as shown is a permanent dipole force. Hydrogen bonding occurs between a
in Eq. (5) [33,43]. Therefore, the overall solubility parameter is the hydrogen atom covalently bonded to other atoms and another atom (X-
square root of the sum of these energy densities (that is, dispersion, H… Y); the atoms (X and Y) on either side of a hydrogen atom are
polar, and hydrogen bonding), as presented in Eq. (6). Notably, the total usually highly electronegative.
solubility parameter of Hansen and the solubility parameters of Hilde­ In particular, W. Zeng et al. abbreviated Hildebrand solubility
brand and Scott are largely the same [45]. Following the definition of parameter and Hansen total solubility parameter to δ and δT, respec­
the total solubility parameter, Hansen regarded the energy densities: tively (Eq. (8)) [45]. They suggested that “the total solubility parameter
dispersion cohesive, polar cohesive, and hydrogen bonding as the pa­ of Hansen δT should be equal to the Hildebrand parameter δ, although
rameters of dispersion cohesion, polar cohesion, and hydrogen bonding the two quantities may differ for materials with specific interactions
cohesion, respectively. Finally, the HSPs are formulated, wherein the when they are determined by different methods [45].” The close or even
total solubility parameter δ is composed of three partial solubility pa­ identical actual measured data of Hansen’s total solubility parameters
rameters (δD, δP, and δH), satisfying the relationship expressed in Eq. (7). and the Hildebrand parameters (Fig. 1) support the conclusion of Zeng
E = ED + EP + EH , (5) et al. However, in contrast, M. Belmares et al. [46] and S. Venkatram
et al. [47] abbreviated Hildebrand parameter as δ without introducing

Fig. 1. Representation of the consistency between Hansen total parameter and Hildebrand parameter of the actual measured substance. The substances are
numbered 1–68 based on the Hildebrand parameter and Hansen total parameter in an ascending order. Solvents 1–68 were perfluoroheptane (1), diethyl ether,
trichlorofluoromethane, butyl acetate (iso), benzonitrile, tricresyl phosphate, amyl acetate, butyl acetate (normal), carbon tetrachloride, ethylbenzene (10), mesi­
tylene, xylene, diethyl carbonate, diethyl ketone, toluene, mesityl oxide, ethyl acetate, tetrahydrofuran, isophorone, benzene (20), trichloroethylene, diacetone
alcohol, chloroform, styrene, tetrachloroethylene, dibenzyl ether, benzaldehyde, acetyl chloride, chlorobenzene, tetrahydronaphthalene (30), diethylene glycol
monobutyl ether (normal), methyl acetate, ethyl bromide, tetrachloroethane-1,1,2,2, ethylene dichloride, cyclohexanone, naphthalene, acetone, bromobenzene,
carbon disulfide (40), dichlorobenzene (ortho), diethyl phthalate, nitrobenzene, acetic acid, acetaldehyde, acetic anhydride, aniline, butyronitrile, hexamethyl­
phosphoramide, acetophenone (50), pyridine, quinoline, epichlorhydrin, nitroethane, cyclohexanol, allyl alcohol, acetonitrile, benzyl alcohol, formic acid, ethylene
diamine (60), furfuryl alcohol, nitromethane, methanol, ethylene carbonate, ethylene cyanohydrin, succinic anhydride, formamide, and water (68) [45],
respectively.

3
Q. Ma et al. International Journal of Biological Macromolecules 256 (2024) 128506

Hansen’s total solubility parameter. Importantly, Hildebrand parameter interaction parameter χ12 and then used to judge the dissolution
δ, as presented in the references of M. Belmares et al. [46] and S. Ven­ behavior of polymer in the solvent [50–52]:
katram et al. [47], was calculated by Eq. (7), which has the same right-
△GM = RT(n1 lnϕ1 + n2 lnϕ2 + χ 12 n1 ϕ2 ) (13)
hand side as in Eq. (8). In summary, there was no essential difference in
the introduction of Hansen’s total solubility parameter, which is
equivalent to the Hildebrand parameter. For brevity, the Hildebrand 2.5. The relationship between χ 12 and HSPs
parameter and Hansen’s total solubility parameter are treated as
equivalent in this study. Both the Flory-Huggins theory and the HSPs concern the law of the
( ) solute dissolution in the solvent, signifying that there should be some
δT = δD 2 + δP 2 + δH 2 1/2
(8) connection between χ12 and HSPs. The Flory-Huggins interaction
parameter χ12 for a system can be estimated where HSPs are known via
Hansen’s approach divided the cohesive energy into three parts,
Eq. (14) [32,53]. However, the reverse is impossible, that is, estimating
while Hildebrand and Scott viewed it as a unified whole. This funda­
the HSPs when χ12 are known. This is because HSPs comprise three
mental difference lies at the core of two approaches. Notably, other
partial solubility parameters: dispersion forces, permanent dipole-
sources of cohesion energy (such as induced dipoles, metallic bonds, and
permanent dipole forces, and hydrogen bonding. However, a single
electrostatic interactions) are generated in various types of molecules.
χ12 cannot be divided into the above three parts.
Given that organic molecules contain none or limited cohesion energy,
Hansen ignored considering their influence on the cohesive energy. χ 12 =
v1 [ ]
(δD1 − δD2 )2 + (δP1 − δP2 )2 + (δH1 − δH2 )2 + β, (14)
The HSPs of a polymer can be considered as the length of a vector in a RT
3-dimensional system with three partial solubility parameters as co­
where v1 represents the molar volume of the solvent. β denotes an
ordinates [48,49]. The radius of the Hansen sphere, Ro, defines the
empirical constant that serves as a correction factor for the Flory
extent to which a difference in solubility parameters can be tolerated.
combinatorial entropy in polar polymer systems, typically with an
The boundary of the spherical characterization depends on the distance
average value of around 0.34. For weak polar polymers, like lignin, the
of good solvents from the center of the sphere (the coordinate of the
constant β is negligible and can be set as zero [53].
polymer) and is calculated using a computerized optimization method
Notably, accurate χ12 values from numerous solvents (with known
[33]. A solvent, soluble in a polymer, encompasses in the system of
HSPs) and the same polymer can be used to determine the HSPs of the
coordinates a volume in which the polymer vector terminates. The
polymer. Other data such as viscosity, solvency, or swelling from
distance between the polymer and the solvent in the sphere (Ra) was
numerous solvents (with known HSPs) and the same polymer can also be
computed in accordance with their respective partial solubility param­
used to ascertain the HSPs of the polymer [54]. Segarceanu et al. [55]
eters in the coordinates, as illustrated in Eq. (9). This equation was
proposed that the higher the interactions between the solvent and so­
developed from the plots of experimental data where the constant “4”
lute, the greater the intrinsic viscosity of the solution. Therefore, the
was determined to be suitable, and accurately represented the solubility
overall and partial solubility parameters of the polymer can be predicted
data within a sphere encompassing the good solvents.
from the intrinsic viscosity of the solution and the solubility parameters
[
Ra = 4(δD1 − δD2 )2 + (δP1 − δP2 )2 + (δH1 − δH2 )2
]1/2
, (9) of the solvents. Segarceanu et al. [55] then calculated the solubility
parameters of polyesterimide using their method, confirming their ac­
where δD1, δP1 and δH1 are the partial solubility parameters of solvent, curacy by comparing it with the solubility parameters computed using
and δD2, δP2 and δH2 are the partial solubility parameters of solvent. Hansen’s dissolution method. However, the main drawback of Segar­
The relative energy difference (RED) was used to determine the ceanu’s method is that the relationship between the intrinsic viscosity
miscibility of lignin with the solvent, and is defined as the ratio of Ra to and solubility of solutes in the solvents is complex and is only set as a
Ro, as shown in Eq. (10): positive correlation. However, this method has not yet been employed to
calculate lignin HSPs.
RED = Ra /R0 (10)
RED numbers <1.0 indicate a high affinity, suggesting that the 2.6. Temperature and molecular weight on polymer dissolution
polymer is more likely to be soluble, mostly soluble, or partially soluble
in the solvent. In contrast, progressively higher RED values >1 indicate Temperature and molecular weight are also critical factors that affect
decreasing affinities, implying that the polymer is more likely to exhibit polymer solubility. From the perspective of Gibbs free energy, high
swelling or minority swelling in the solvent. temperature is favorable for endothermic dissolution, while low tem­
perature is favorable for exothermic dissolution [33,50].
An increase in the temperature can reduce the energy required for
2.4. The relationship between ΔGM and χ 12
vaporization, thereby decreasing the solubility parameter of the sub­
stance. However, the change in the solubility parameter of the solute
In essence, there is a close relationship between the free energy of
with increasing or decreasing temperature is usually slightly less than
mixing (ΔGM in Eq. (1)) and the Flory-Huggins interaction parameter
that of the solute. Therefore, the impact of temperature on solubility
(χ12 in Eq. (2)). According to the definition of mixed entropy, ΔSM can be
depends on the solubility parameters of the solute and solvent [52]. For
expressed by Eq. (11):
example, when the solubility parameter of the solvent is greater than
△SM = − R(n1 lnϕ1 + n2 lnϕ2 ), (11) that of the solute, an increase in temperature reduces the difference
between the solubility parameters of the solvent and solute, thereby
where n1 and n2 are the mole numbers of solvent and polymer, and ϕ1 increasing solubility. Conversely, when the solubility parameter of the
and ϕ2 are the volume fractions of solvent and polymer. solvent is smaller than that of the solute, an increase in temperature
The heat of mixing ΔHM can also be expressed by Eq. (12) based on increases the difference between the solubility parameters of the solvent
the Flory–Huggins theory [37–39]: and solute, thereby decreasing the solubility.
A higher molecular weight enhances polymer chain entanglement
△HM = − χ 12 RTn1 ϕ2 (12)
and thus reduces the tendency for environmental stress cracking when
Hence, Eq. (12) can be obtained by substituting Eqs. (11) and (12) the solvent is absorbed [56]. A higher molecular weight leads to stronger
into Eq. (1). Importantly, this equation revealed that the free energy of entanglements between the polymers, thus reducing the possibility of
mixing ΔGM can be calculated according to the Flory-Huggins environmental stress cracking when the solvent is absorbed and swelled.

4
Q. Ma et al. International Journal of Biological Macromolecules 256 (2024) 128506

The fact that cross-linked and crystalline polymers are almost insoluble (∑ )1/2
is suitable evidence [33,52,57]. Notably, when the solubility parameters δH =

∑ ei-H , (18)
are the same, small-molecule solvents are more likely to dissolve the △Vi
polymers than larger solvent molecules. This is because small-molecule
solvents are more likely to approach polymers, especially large-molecule where Δei-D, Δei-P, and Δei-H are the evaporation energies required to
polymers, with significant steric hindrance. overcome the dispersion cohesion, polar cohesion, and hydrogen
bonding cohesion of atoms or functional groups, respectively, and Δvi
represents the molar volume of an atom or functional group.
3. The calculation of polymer HSPs

4. The calculation of lignin HSPs


The cohesive energy, which holds the molecules together, can be
directly measured by vaporizing the substance, particularly the solvent,
4.1. Hansen dissolution method
from liquid to gas. The vaporization method has proven to be reliable for
measuring the total cohesive energy of solvent molecules. However, the
Hansen first measured the HSPs in lignin using the dissolution
cohesive energy of high-molecular-weight polymers, such as lignin,
method [33,49]. Eighty-two solvents with known HSPs were used to
cannot be measured using the evaporation method. As the heating
dissolve milled wood lignin (MWL), with the results divided into five
temperature increased, the high-molecular-weight polymers underwent
categories based on visual observations: soluble, majority-soluble,
pyrolysis and were depolymerized into more minor molecular com­
partially soluble, swelling, and minority-swelling. Then, the cate­
pounds before reaching the boiling point. In addition, the three partial
gories: soluble, majority soluble, and partially soluble were coded as
cohesive energies (ED, EP, and EH) cannot be determined depending on
soluble, while swelling and minor swelling were labeled as insoluble
the evaporation method because the contributions of these three types of
[33,49,66]. The soluble or insoluble fractions and the solubility pa­
partial cohesive energies to vaporization cannot be clearly
rameters of the solvents were processed by the Hansen computer pro­
distinguished.
gram [67]. Finally, the solubility parameters δD, δP, δH, and δ (10.7,
The polar and dispersion cohesive energies are usually measured
6.89, 8.26, and 15.2 cal1/2⋅cm–3/2 respectively), as well as Ro (6.70 cal1/
through complex experiments and calculations rooted in the theory of 2
⋅cm–3/2), for MWL were obtained (Table 1). Depending on the lignin
statistical thermodynamics [33]. These experiments and calculations are
solubility parameters, we can quantitatively (rather than qualitatively)
so refined that the ED and EP values of most monomolecular solvents and
analyze the dissolution behavior of lignin [33,49].
polymers can be obtained by referring to a handbook. The δD parameter
Numerous researchers have improved Hansen’s dissolution method
was calculated according to the procedure outlined by Blanks and
[64,66,68–71]. Vebber et al. [66] optimized the HSPs calculation pro­
Prausnitz [44]. Standard figures can be used to determine this param­
cess for polymers using a genetic algorithm. The partial and overall
eter, depending on the aliphatic, cycloaliphatic, or aromatic type. The
solubility parameters of MWL calculated by Vebber et al. [66] based on
ED of straight-chain hydrocarbons was shown to be a function of the
the improved genetic algorithms method were 10.6, 6.93, 8.28, and
molar volume and reduced temperature (Tr). If the molar volume and Tr
15.1 cal1/2⋅cm–3/2, which is highly consistent with the results of the
of a molecule are known, the ED can be found in the standard figure, and
Hansen dissolution method 10.7, 6.89, 8.26, and 15.2 cal1/2⋅cm–3/2
the δD parameter can be confirmed. The dispersion cohesive energy
[33,49]. Díaz de los Ríos et al. [71] determined the HSPs and Hansen
density (ED/V) for cycloalkanes and aromatic hydrocarbons was shown
sphere radii using the Solver add-in in Microsoft Excel. The result for
to be a function of molar volume and Tr. If the molar volume and Tr of a
lignin HSPs is an optimal region rather than a point [71,72]. The Hansen
molecule are known, ED/V can be found in the standard figure, and then
dissolution method provides an important basis for the calculation of
the δD parameter can be obtained by taking the square root of ED/V.
lignin HSPs and the exploration of lignin dissolution behavior. However,
Hansen and Beerbower [58] proposed Eq. (15) to calculate the δP. An
the Hansen dissolution method has two shortcomings: (1) its reliance on
advantage of this equation is that δP can be calculated using the dipole
qualitative rather than quantitative measurement can lead to theoretical
moment, which can be obtained from the McClellan standard table [59].
inaccuracies, although the use of a large number of solvents compen­
Notably, the results calculated using this equation are reliable. The EH
sates for this inaccuracy; and (2) the method is complex, owing to the
can be computed by subtracting the polar and dispersion energies of
necessity of employing large amounts of solvents [64].
vaporization from the total vaporization energy.
/
δP = 18.3(DM) V1/2 , (15) 4.2. Inverse gas chromatography (IGC) method
1/2
where the constant 18.3 yields this parameter in units of cal ⋅cm –3/2
, Ren et al. [69] calculated lignin solubility parameters using the IGC
and DM represents the dipole moment of the molecule. method rooted in the Flory-Huggins interaction parameter [57]. First,
The atomic and functional group contribution (AFGC) method is a the retention volumes of the six solutes were determined by gas chro­
valuable and widely-used tool for calculating the HSPs of polymers matography employing alkali lignin as the stationary phase. The ther­
[58,60–63]. When the structure of the repeating unit of the polymer is modynamics parameters of the probe solvents (primarily including
known, and the reliable latent heat and dipole moment data are absent, weight fraction activity coefficient, molar absorption enthalpy, and
the HSPs of the polymer can be computed using AFGC. The partial sol­ molar evaporation enthalpy) and the interaction parameters between
ubility parameters were calculated employing Eqs. (16)–(18). The alkali lignin and solvents were calculated in turn. Finally, the average
overall solubility parameter can then be obtained utilizing Eq. (11). The overall solubility parameter of 11.9 cal1/2⋅cm–3/2 for alkali lignin be­
related molar volume, dispersion cohesive energy, polar cohesive en­ tween the temperatures of 25–100 ◦ C was calculated (Table 1). More
ergy, and hydrogen bonding of atoms and functional groups comprising details regarding the use of the IGC method to measure overall lignin
the polymer can be obtained from the references [33, 43, 64, 65]. solubility parameters can be found in a previous study [69]. The overall
(∑
△ei-D
)1/2 solubility parameter of alkali lignin measured by IGC displayed the
δD = ∑ , (16) result of 11.9 cal1/2⋅cm–3/2. This is nearly identical to that measured by
△Vi
the Hansen dissolution method (12.0 cal1/2⋅cm–3/2), proving the accu­
(∑ )1/2 racy of the IGC method. Additionally, when the temperature increased

δP = ∑ ei-P , (17) from 25 to 100 ◦ C, the overall solubility parameter of the alkali lignin
△Vi
decreased slowly from 12.0 to 11.8 cal1/2⋅cm–3/2 (Table 1), indicating
that the investigated condition temperature showed limited influence on

5
Q. Ma et al. International Journal of Biological Macromolecules 256 (2024) 128506

Table 1 the overall solubility parameter. Ren et al. [69] presented a new method
Lignin solubility parameters calculated via different methods. for determining the overall lignin solubility parameters depending on
Calculation methods and lignin samples Solubility parameters (cal1/2⋅cm–3/ the IGC. However, the three partial solubility parameters, and the radius
2
) of the Hansen sphere of alkali lignin, were not (or cannot be) determined
δD δP δH δ by the IGC method [69,79].

Hansen dissolution method


MWL [33] 10.7 6.89 8.26 15.2 4.3. AFGC method
MWL on genetic algorithm optimized Hansen 10.6 6.93 8.28 15.1
dissolution method [66]
Sugar cane bagasse lignin [68] 10.5 4.17 10.7 15.5
Ni et al. [65] computed the overall solubility parameter of lignin
Alkali lignin1 [69] 8.42 6.04 6.03 12.0 employing the AFGC method based on the empirical C9 formula devel­
oped by Hoy et al. [60]. First, the C9 molecular formula of the hardwood
IGC method ALCELL lignin was determined by elemental analysis. Secondly, the G/
Alkali lignin1 at 25, 40, 60, 80, and 100 ◦ C S/H ratio of the hardwood ALCELL lignin (76/20/4) was reasonably
12.0, 11.9, 11.9, 11.9, and 11.8
(Overall solubility parameters) [69] inferred according to the C9 molecular formula and hardwood charac­
Average of Alkali lignin1 at 25 ◦ C–100 ◦ C [69] N/A N/A N/A 11.9 teristics. Third, the atomic and functional groups or the empirical C9
Alkali lignin2 [73] N/A N/A N/A 9.82
formulas of the G, S, and H units were further clarified (Fig. 2). For
example, the numbers of methylene (>CH2), methine (>CH–), and
AFGC method based on the lignin empirical C9 formula
olefin (>C=) structures, methoxyl and hydroxyl groups, and atomic
AlCELL lignin (G: S: H = 74: 20:6) [65] N/A N/A N/A 13.7
Almond shell lignin (G: S = 1: 1.25) [74] N/A N/A N/A 14.6
oxygen were 1, 1, 1, 1 and 1, and 1 for the G unit; 1, 2, 0, 2 and 2, and 1
Cornstalk lignin (G: S: H = 40: 50: 10) [75] N/A N/A N/A 13.7 for the S units; and 1, 1, 1, 0 and 1, and 2 for the H units (Fig. 2). At last,
Bagasse enzymatic hydrolysis/mild acidolysis
N/A N/A N/A 14.0
the overall solubility parameter of G, S, and H units was calculated as
lignin (EMAL) (G: S: H = 2.65: 1: 5.07) [76] 13.5, 14.2, and 14.1 cal1/2⋅cm–3/2, respectively, according to Eqs. (16)–
(18) and (7), respectively. Then, the δ of hardwood ALCELL lignin was
AFGC method based on the lignin structure (the lignin structure was not clearly calculated as 13.7 cal1/2⋅cm–3/2 depending on its G/S/H ratio (Table 1).
revealed in the reference) [77] Ni et al. [65] provided a more straightforward method without
Pine kraft lignin 8.17 6.70 5.72 12.0
Acetylated pine kraft lignin 7.87 5.87 4.99 11.0
complicated experimental measurements and tedious calculations to
Propionated pine kraft lignin 7.97 5.62 4.74 10.8 determine the overall solubility parameter of lignin compared to the
Butyrated pine kraft lignin 8.02 5.43 4.55 10.7 Hansen dissolution [33,49] and the IGC methods [69]. However, Ni
Maleated pine kraft lignin 7.63 6.45 5.23 11.2 et al. [65] did not consider the differences between lignin atoms and
Methacrylated pine kraft lignin 7.78 5.72 4.50 10.7
functional groups caused by different raw materials and separation
Hardwood kraft lignin 8.17 6.60 5.53 11.9
Acetylated hardwood kraft lignin 7.87 5.92 5.04 11.1 methods. In addition, the lignin partial solubility parameters and the
Propionated hardwood kraft lignin 7.97 5.62 4.74 10.9 radius of the Hansen sphere were not determined in their study.
Butyrated hardwood kraft lignin 8.02 5.43 4.55 10.7 Wool et al. [77] calculated the partial and overall solubility param­
Maleated hardwood kraft lignin 7.63 6.41 5.23 11.3 eters of several lignin samples, including that of pine kraft, hardwood
Methacrylated hardwood kraft lignin 7.78 5.72 4.50 10.7
kraft, and esterified, using the AFGC model. The result revealed that the
partial and overall solubility parameters were 8.17, 6.70, 5.53, and 12.0
Molecular dynamics simulations based on the lignin model polymeric structure [56]
cal1/2⋅cm–3/2 for pine kraft lignin and 8.17, 6.60, 5.53, and 11.9 cal1/
9.33 (δe = 2
Lignin (DP = 8) 9.14
δP + δH)
13.1 ⋅cm–3/2 for hardwood kraft lignin (Table 1). Regrettably, a detailed
9.04 (δe = lignin structure and calculation process have not yet been reported [77].
Lignin (DP = 11) 8.82 12.6
δP + δH) This may explain why the calculation of partial solubility parameters
7.83 (δe = using the AFGC method has not been widely adopted.
Lignin (DP = 26) 8.09 11.3
δP + δH)
Mohan et al. [56] calculated the HSPs of empirical model lignin,
composed of all possible linkages, such as β-1, β-5, 5–5, β-O-4, α-O-4, and
AFGC method based on the lignin expanded C9 formula [78]
4-O-5, depending on the molecular dynamics simulations (MDS).
12.87;
Bioethanol residual lignin (L1); acetylated L1 N/A N/A N/A
10.97
However, the explicit separation of the contributions of polar and
13.09; hydrogen bonds from electrostatic terms (δe) cannot be achieved
Hardwood kraft lignin (L2); acetylated L2 N/A N/A N/A
11.32 through the MDS method [80]. The empirical model of lignin having a
13.42; degree of polymerization (DP) of 8, 11, and 26 exhibited δD of 9.14,
Softwood kraft lignin (L3); acetylated L3 N/A N/A N/A
11.03
9.33, and 13.1 cal1/2⋅cm–3/2; δe of 8.82, 9.04, and 12.6 cal1/2⋅cm–3/2; and
12.92;
Non-wood soda lignin (L4); acetylated L4 N/A N/A N/A
11.04

AFGC method based on the lignin expanded C9 formula [64]


wheat straw organic acid lignin (WSOAL) 10.8 4.15 6.30 13.2
eucalyptus kraft lignin (EKL) 10.9 4.27 6.47 13.4
bamboo kraft lignin (BKL) 10.9 4.28 6.51 13.4
softwood kraft lignin (SKL) 11.1 4.31 7.38 14.0

AFGC method based on the lignin empirical C9 formula (quick method to calculate
HSPs proposed in this review)
WSOAL (G: S: H = 51: 42: 7) [64] 10.5 4.58 7.72 13.8
EKL (G: S = 19.3: 80.7) [64] 10.4 4.81 8.23 14.1
BKL (G: S: H = 22.2: 63.1: 14.7) [64] 10.5 4.74 8.05 14.0
SKL (G: H = 95.8: 4.2) [64] 10.7 4.30 7.11 13.5

Fig. 2. The empirical structure of lignin units.

6
Q. Ma et al. International Journal of Biological Macromolecules 256 (2024) 128506

δ of 8.09, 7.83, and 11.3 cal1/2⋅cm–3/2, as shown in Table 1. The Table 2


advantage of the MDS approach is evident: it helps to understand the The calculation process partial and overall solubility parameters of G, S, and H
solubility parameters from the perspective of the lignin structure, units based on the AFGC method.
especially the DP and linkages. However, the inability to distinguish The calculation process partial and overall solubility parameters (G unit as an
between polar and hydrogen-bonding cohesion parameters introduces example)
certain limitations. Atom or group Number Δei-D Δei-P Δei-H Δei cal/ Δvi
Lignin structures determine the lignin solubility parameters; how­ in C9 cal/mol cal/ cal/ mol cm3/
ever, none of the above methods establish a relationship between the mol mol mol
exact lignin structure and solubility parameters. Samend et al. [78,81] Tri-
analyzed the element content through elemental analysis. Furthermore, substituted 1 7530 50 50 7630 33.4
they examined the functional group content (aliphatic hydroxyl (OHal), phenyl (G)
-OHal 1 1370 1100 4650 7120 10
phenolic hydroxyl (OHph), carboxyl) and G/S/H ratio via 31P nuclear >CH2 1 1180 0 0 1180 16.1
magnetic resonance (NMR), investigated the methoxyl groups, the >C= 1 800 60 180 1040 − 5.5
number of aliphatic hydrogens (Hal), and aromatic hydrogens (Hph) by -O- 1 0 425 375 800 3.8
1 -CH< 1 820 0 0 820 − 1
H NMR. Subsequently, they measured the expanded C9 formula
-OCH3 1 1125 425 375 1925 37.3
(C9Hal ph al ph
y1Hy2 Oz(OH)m1(OH)m2 (OCH3)n), calculating the overall solubility
2×9
parameters of lignin employing the AFGC method. The overall solubility Δv[a]
o
= 18.0
parameters of bioethanol residual lignin, hardwood kraft lignin, soft­ Sum – 12,825 2060 5630 20,515 112.1
wood kraft lignin, and non-wood soda lignin were 12.9, 13.1, 13.4, and Solubility parameter of G δD δP δH δ

12.9 cal1/2⋅cm–3/2, respectively (Table 1). The overall solubility pa­ (cal1/2⋅cm–3/2) 10.7 4.29 7.09 13.5

rameters calculated using the expanded C9 formula were considered to


be more accurate than those calculated using the empirical C9 formula.
However, partial solubility parameters were ignored. Partial and overall solubility parameters (cal1/2⋅cm–3/2) of G, S, and H units

In our previous study [64], we utilized reliable analytical methods to δD δP δH δ


determine the exact structure of lignin. We subsequently conducted (1) G 10.7 4.29 7.09 13.5
elemental analysis to determine the C, H, N, S, and O contents, (2) 1H S 10.3 4.93 8.50 14.2
NMR spectroscopy analysis to determine the methoxyl content, (3) 31P H 10.9 4.59 7.60 14.1
NMR spectroscopy analysis to determine the phenolic hydroxyl (OHal),
aliphatic hydroxyl (OHph), carboxyl, and acetoxyl group contents, (4)
two-dimensional (2D) heteronuclear single-quantum coherence (HSQC) δD-L = 10.7 × G% + 10.3 × S% + 10.9 × H%, (19)
NMR spectroscopy analysis to determine the ratios of G, S, and H units
[64]. We derived a more detailed expanded C9 formula δP-L = 4.29 × G% + 4.93 × S% + 4.59 × H%, (20)
(C9HyOz(OH)al ph
m1(OH)m2(OCH3)n1(COOH)n2(CH3COO)n3) for lignin and
calculated the HSPs of the four lignin samples using the AFGC method δH-L = 7.09 × G% + 8.50 × S% + 7.60 × H%, (21)
[64]. The result revealed the δD, δP, δH, and δ of wheat straw organic acid ( )
lignin (WSOAL), eucalyptus kraft lignin (EKL), bamboo kraft lignin δL = δD-L 2 + δP-L 2 + δH-L 2 1/2
, (22)
(BKL), and softwood kraft lignin (SKL) as (10.8, 4.15, 6.30, and 13.2
cal1/2⋅cm–3/2), (10.9, 4.27, 6.47, and 13.4 cal1/2⋅cm–3/2), (10.9, 4.28, where δD-L, δP-L, δH-L, and δL were the HSPs of lignin, and G%, S%, and H
6.51, and 13.4 cal1/2⋅cm–3/2), and (11.1, 4.31, 7.38, and 14.0 cal1/ % were the ratio of G, S and H subunits, respectively.
2
⋅cm–3/2), respectively (Table 1). Thus, we established a universal and The HSPs of lignin were based on the calculation methods. Different
accurate method for determining the solubility parameters of lignin testing methods may yield different levels of response (or error) to lignin
according to its structure. However, the analysis and calculation of the impurities (such as polysaccharides, sugars, and inorganic components).
solubility parameters rooted in the expanded C9 formula are compli­ At present, there is a lack of research on the influence of impurities on
cated and tedious. the HSP results using various test methods. However, if the G/S/H ratio
of lignin can be accurately determined using the quick AFGC method,
4.4. A quick method to calculate lignin HSPs the HSPs of lignin can also be precisely computed to eliminate the in­
fluence of impurities.
Existing methods for determining the solubility parameters of lignin We further recalculated the HSPs of WOSAL, EKL, BKL, and SKL
face limitations. Some are tedious, such as the Hansen dissolution employing an established rapid method. As the ratio of the G, S, and H
method [33,49] and the AFGC method derived from the accurate subunits of these lignin samples was calculated utilizing the 2D NMR
expanded C9 formula outlined in our previous study [64]. Others are semi-quantitative method (Table 1) [64], their HSPs were calculated
unable to determine all the partial solubility parameters, including the through Eqs. (19)–(22). The result revealed that the (δD, δP, δH, and δ) of
IGC method [69] and the AFGC method based on the empirical by Ni WOSAL, EKL, BKL, and SKL were (10.5, 4.85, 7.72, and 13.8 cal1/2⋅cm–3/
2
et al. [65]. ), (10.4, 4.81, 8.23, and 14.1 cal1/2⋅cm–3/2), (10.5, 4.74, 8.05, and 14.0
In this review, we propose a more accessible and quick approach for cal1/2⋅cm–3/2), and (10.7, 4.30, 7.11, and 13.5 cal1/2⋅cm–3/2) respec­
measuring the HSPs of lignin using the AFGC method derived from the tively (Table 1). When compared to the HSPs of WOSAL, EKL, BKL, and
empirical C9 formula of lignin. First, the HSPs of the G, S, and H subunits SKL, calculated based on expanded C9 structure (Table 1), the values
of the empirical C9 formula (Fig. 2) were calculated using the AFGC calculated via the quick method displayed a small absolute error of
method. The calculation process was shown in Table 2, and the partial 0.30–0.50 cal1/2⋅cm–3/2 for δD, 0–0.54 cal1/2⋅cm–3/2 for δP, 0.27–1.76
and overall solubility parameters (δD, δP, δH, and δ) were (10.7, 4.29, cal1/2⋅cm–3/2 for δH, and 0.50–0.70 cal1/2⋅cm–3/2 for δ. In summary, this
7.09, and 13.5 cal1/2⋅cm–3/2) for G subunit, (10.3, 4.93, 8.50, and 14.2 is a straightforward and effective method for measuring lignin HSPs
cal1/2⋅cm–3/2) for S subunit, and (10.9, 4.59, 7.60, and 14.1 cal1/2⋅cm–3/ using the AFGC method according to the empirical C9 formula proposed
2
) for H subunit (Table 2). The HSPs of lignin can then be easily calcu­ in this review. This particularly when the accuracy requirements are not
lated using Eqs. (19)–(22) since the HSPs are proportional to the con­ exceptionally high. For further accuracy, the lignin HSPs were calcu­
tents of the G, S, and H subunits. This method is largely based on the lated using the AFGC method and the expanded C9 formula of lignin,
findings of Ni et al. [65,77]. which was measured using a set of detailed analyses of the lignin

7
Q. Ma et al. International Journal of Biological Macromolecules 256 (2024) 128506

structure proposed in our previous study [64]. meleated, and methacrylated) to decrease their HSPs and, therefore,
reduce their difference from that of solvents [77]. As expected, the
5. Application of lignin HSPs solubility of esterified lignin samples (such as styrene-containing ther­
moset resins) in nonpolar solvents was improved [77]. Phloretic acid-
5.1. Searching for ideal solvents to dissolve lignin esterified wheat straw soda lignin also displayed lower HSPs and
higher solubility than unesterified lignin [86]. Esterified lignin also
According to HSPs theory, substances with similar solubility pa­ demonstrates a higher onset temperature of thermal degradation and
rameters show high compatibility and are more likely to dissolve in each glass transition temperature, rendering it suitable for unsaturated
other [33]. Therefore, HSPs are crucial tools in the search for solvents to thermosetting composites.
dissolve lignin [82]. S. Venkatram et al. proposed that when the dif­
ference between the solvent and polymer δ value is <0.978 cal1/2⋅cm–3/2 5.2. Selecting solvents for lignin depolymerization and extraction
(2 MPa1/2), the solvent is considered as a good solvent for the polymer.
Conversely, when the difference between the solvent and polymer δ HSPs have been proven helpful in selecting solvents for lignin
value is >0.978 cal1/2⋅cm–3/2 (2 MPa1/2), the solvent is classified as a depolymerization and extraction from lignocellulosic raw materials
non-solvent for the polymer [47]. Furthermore, a radius Ro of 3.91 (8 [87,88]. The lignin in hydrolyzed almond shells, with a total solubility
MPa1/2) is the boundary of the Hansen sphere. Solvents that fall inside parameter of 14.6 cal1/2⋅cm–3/2, was extracted under acetone/water,
the sphere of the polymer are deemed good solvents, and those that fall ethanol/water, and dioxane/water systems [74]. The lignin extraction
outside the sphere are considered non-solvents [47]. yield reached its maximum value when the total solubility parameters of
The dissolution behavior of ALCELL lignin with a total solubility acetone/water, ethanol/water, and dioxane/water were 14.4, 15.6, and
parameter of 13.7 cal1/2⋅cm–3/2 in an ethanol/water system was inves­ 14.0 cal1/2⋅cm–3/2 respectively, which was close to that of lignin. The
tigated when the total solubility parameter of solvents decreased degree of lignin removal from enzymatically hydrolyzed cornstalks
continuously from 22.3 to 12.1 cal1/2⋅cm–3/2, as the ethanol ratio tends to increase when the total solubility parameters of lignin and
increased from 0 to 100 % [65]. ALCELL lignin exhibited higher solu­ ethanol/water, 1, 4-dioxane/water, and tetrahydrofuran/water solvents
bility of 17.8 to 18.0 g/L in ethanol/water system with total solubility become more closely matched [75]. The relationship between RED
parameters ranging from 13.2 to 14.1 cal1/2⋅cm–3/2, which is quite (calculated based on the HSPs of lignin and various solvents) and the
close to the solubility parameter of lignin (13.7 cal1/2⋅cm–3/2). The degree of delignification of acid-hydrolyzed sugarcane bagasse was
relationship between the solubility of bagasse EMAL solubility in 1, 4- established [68]. A linear relationship with a correlation of 0.93856 was
butanediol/water mixture and the total solubility parameter has been observed between RED and the extent of delignification. The smaller the
established [76]. EMAL, with a total solubility parameter of 14.0 cal1/ difference in the partial solubility parameters between lignin and the
2
⋅cm–3/2, showed the highest solubility at 1, 4-butanediol/water (80/20, solvents, the higher the degree of delignification [68]. The degree of
v/v) solvent, with the total solubility parameter of 14.6 cal1/2⋅cm–3/2. lignin removal from biomass and RED demonstrated a correlation of
This further demonstrates that the search for good solvents to dissolve 0.8432 [89]. Similar results were achieved for the depolymerization of
lignin is possible based on the overall solubility parameters. coir and poplar using an alcohol/water/acidic ionic liquid solvent sys­
In our previous study, we established a dissolution model of WSOAL, tem [90]. These results prove that solubility parameters can be used for
EKL, BKL, and SKL in formic acid/water and acetic acid/water systems. solvent screening in biomass depolymerization.
Furthermore, the relationship between lignin solubility and HSPs had Ionic liquids (IL) have proven efficient for the pretreatment of
conformed to a Gaussian curve with a correlation of 0.954 [64]. biomass, such as carbohydrate dissolution and lignin extraction, under
Significantly, the established dissolution model accurately predicted mild conditions [91]. However, the difficulty in ionic liquid recovery
lignin solubility in a formic acid/acetic acid/water system. The model hinders the broad application of this approach. The solubility and
adhered to the solubility parameter theory, overcoming the limitations Kamlet-Taft parameters play an essential role as reference criteria when
of the “like dissolves like” principle in organic acid/water systems, and selecting suitable organic solvents for extracting IL from the IL-lignin
provided a concise method for selecting solvent systems, as well as their system [92]. [BMIM]OAc/water solvents (60/40, wt%), with an over­
ratios, to dissolve or precipitate lignin. all solubility parameter of 12.8 cal1/2⋅cm–3/2, reached the maximum
Organosolv lignin, with a solubility parameter of 12.5 cal1/2⋅cm–3/2, lignin (with an overall solubility parameter of 12.9 cal1/2⋅cm–3/2)
exhibited its maximum solubility in γ-valerolactone (GVL)/water sol­ extraction degree for water hyacinth [54]. The theoretical interpreta­
vents (92–96/8–4, wt%), with similar solubility parameters ranging tion, which is based on the solubility parameter concept, agrees with the
from 11.8 to 12.4 cal1/2⋅cm–3/2 [83]. The selected GVL/water systems experimental results when solvents containing IL are selected for lignin
with solubility parameters close to those of lignin can be used for lignin delignification [93].
extraction [83]. HSPs reveal a significant correlation with the solubility In summary, HSPs can be used to select solvents and determine their
of lignin and guaiacol in different tetrahydrofuran(THF)/water systems ratio for lignin depolymerization and extraction from lignocellulosic and
[84]. In conclusion, the solubility parameters can determine the THF/ acid- or enzyme-pretreated raw materials. Notably, during lignin
water ratio during the depolymerization of lignin for guaiacol produc­ delignification and extraction, the following processes also occurred:
tion [84]. impregnation of solvents into the raw material cell wall, cleavage of
Lignin exhibited higher solubility in five deep eutectic solvents covalent bonds between lignin and carbohydrates, cleavage of ether
(DESs: choline chloride/glycerol, choline chloride/oxalic acid, choline bonds between lignin molecules, and formation of a lignin condensed
chloride/malic acid, choline chloride/lactic acid, and choline chloride/ structure. Therefore, these complex reactions should be considered
glacial acetic acid) than in common solvents, such as water, ether, when HSPs are used to select solvents for lignin extraction.
ethanol, chloroform, dimethyl sulfoxide, dioxane, tetrahydrofuran,
acetone, carbon tetrachloride, n-hexane, and ethyl acetate [85]. This is 5.3. Designing green solvents for lignin fractionation
because the hydrogen-bonding cohesion parameter of DESs is closer to
that of lignin than that of common solvents, and the hydrogen-bonding The instability in performance caused by polydispersity is a bottle­
force is the most dominant force in determining lignin solubility. neck that restricts high-value applications of lignin. Lignin fractionation,
In addition to searching for good solvents to dissolve lignin especially via sequential solvent extraction rooted in lignin solubility,
depending on the solubility parameters, it is also feasible to enhance its can potentially obtain lignin fractions with excellent homogeneity and
solubility by modifying the lignin solubility parameters. Hardwood and stable performance [25,94–96]. Two to five solvent systems with
softwood lignins are esterified (acetylated, propionated, butyrated, different solubility parameters comprising water [65,97,98] and organic

8
Q. Ma et al. International Journal of Biological Macromolecules 256 (2024) 128506

solvents, such as hexane, acetone [97–102], ether, dichloromethane, solubility parameter of 9.14 cal1/2⋅cm–3/2, exhibited the highest tensile
[103], ethyl acetate [97,99,100,104], methanol [99,100,102–104], strength of 18.86 MPa and a hydrophilic contact angle of 89o [73]. The
ethanol [65,99,101,104], dioxane [102,104], 2− /n-butanol, n-butanol, blends of alkaline lignin/ acrylonitrile− butadiene− styrene (ABS) pre­
1, 4-butanediol [76], formic acid [105], acetic acid and methyl ethyl pared by mixing showed the total solubility parameters of 8.51, 9.39,
ketone [99], are often used for lignin fractionation. Ajao et al. [99] 9.28, and 8.49 cal1/2⋅cm–3/2 when the proportions of incorporated lignin
fractionated lignin using methanol, ethanol, acetone, methyl ethyl ke­ were 0, 10, 20, and 30 %, respectively [113]. The total solubility pa­
tone, and ethyl acetate. The result suggested that the molecular weight rameters of the alkaline lignin/ABS blends initially increased and then
of the dissolved fraction is below that of the insoluble fraction. Jiang decreased as the tensile strength and thermal weight loss temperature
et al. [100] dissolved lignin using a mixture of methanol/acetone and increased. These results provide vital support for the preparation of
successively added ethyl caproate, ethyl acetate/petroleum ether, and high-performance alkaline lignin blends containing PVA, ABS, or other
petroleum ether to reduce the solvent solubility parameters for contin­ polymers. Haridevan et al. observed that solubility parameters, molec­
uous fractionation. The molecular weight and glass transition temper­ ular weight, dielectric constants, and structure determine the dispersion
ature of the insoluble lignin fractions decreased successively, whereas and solubility of lignin in polyols [114]. Therefore, selecting polyols
those of the phenolic hydroxyl and carboxyl groups increased. Sun et al. depending on their solubility enables the preparation of sustainable,
conducted sequential extraction of kraft-AQ lignin using hexane, diethyl compatible, and high-performance polyurethanes [114,115]. HSPs have
ether, methylene chloride, methanol, and dioxane (with total solubility helped predict and rationalize the behavior of lignin-based rubber
parameters of 7.3, 7.4, 9.7, 14.5, and 10.0 cal1/2⋅cm–3/2, respectively) composites [116]. The results highlighted that the affinity between
[90]. The five resulting fractions exhibited lower polydispersity and lignin and rubber could be quantified according to the HSP theory.
better thermal stability.
Sequential fractionation is essential for obtaining homogeneous and 6. Conclusions
stable lignin fractions [106]. However, the relationship between the
difference in solubility parameters (between lignin and solvents) and the This review summarizes dissolution-related theories, such as the
characteristics of the fractions has yet to be established. The following Gibbs free energy of dissolution, Flory-Huggins interaction parameters,
two issues need to be studied to reveal the mechanisms underlying the and HSPs. HSPs reflect intermolecular or intramolecular interaction
solubility parameters of lignin fractionation. The HSPS of the lignin forces and can be useful for a deeper understanding of the lignin
benzene ring and the side-chain propane structure were different. For dissolution behavior.
example, the HSPs (δD, δP, δH, and δ) of the benzene ring structure and The HSPs of lignin can be calculated using the Hansen dissolution,
side chain propane structure of bamboo kraft lignin are (15.1, 1.23, IGC, and AFGC methods depending on the lignin solubility in various
1.23, and 15.2 cal1/2⋅cm–3/2) and (8.75, 5.78, 9.85 and 14.4 cal1/2⋅cm–3/ solutions, interaction parameters between lignin and solvents, and the
2
), respectively. Although their total solubility parameter is close to each empirical model or the accurate expanded structure of lignin, respec­
other, their partial solubility parameters are significantly different. tively. However, owing to the complexity of the lignin structure, these
Thus, the difference between the lignin benzene ring and the side-chain methods are not sufficiently concise or cannot calculate the three partial
propane structure during fractionation should be further studied. In solubility parameters. Therefore, we propose a quick method for
contrast, the three partial solubility parameters are fundamentally measuring the partial and overall solubility parameters of lignin using
derived from different intramolecular or intermolecular interactions of the AFGC method based on empirical C9 subunits. The partial and
dispersion, polar, and hydrogen bond forces. However, differences in overall solubility parameters (δD, δP, δH, and δ) of G, S, and H subunits
fractionation have not yet been revealed. were calculated as (10.7, 4.29, 7.09, and 13.5 cal1/2⋅cm–3/2), (10.3, 4.93,
8.50, and 14.2 cal1/2⋅cm–3/2), and (10.9, 4.59, 7.60, and 14.1 cal1/
2
5.4. Guiding the preparation of lignin-based composites ⋅cm–3/2), respectively. The HSPs of the lignin were calculated using Eqs.
(19)–(22), once the ratios of the G, S, and H subunits are determined.
Lignin is a renewable and biodegradable aromatic polymer that can HSPs are crucial for guiding most processes related to lignin swelling
be used to produce high-value thermosetting plastics, thermoplastic and dissolution. At present, HSPs primarily aid in the search for ideal
plastics, and other composites or blends [107]. Compared with the solvents to dissolve lignin, selection of solvents for lignin depolymer­
original lignin and polymers, the synergistic effects in lignin polymer ization and extraction, design of green solvents for lignin fractionation,
blends can result in superior curing time, durability, and mechanical and preparation of lignin-based composites. Therefore, it is necessary to
performance [108,109]. Importantly, lignin-based polymers cause focus more attention on lignin HSPs to better guide scientific and
minimal or no environmental pollution associated with existing poly­ applied research on lignin.
mers produced by petroleum refining platform products [110]. The
HSPs are practical tools for predicting the compatibility of lignin with CRediT authorship contribution statement
other polymers. The more significant the difference in solubility pa­
rameters between lignin and other polymers, the more difficult it is to Qingzhi Ma: Investigation, Methodology, Writing – original draft,
become miscible because more electron-donor/electron-acceptor in­ Writing – review & editing. Changqing Yu: Conceptualization, Formal
teractions are needed to bridge the energy gap than in a less miscible analysis. Yuran Zhou: Funding acquisition, Supervision, Writing – re­
system [107,110]. view & editing. Dinggen Hu: Funding acquisition, Investigation. Jian­
The total solubility parameter of acetylated pristine lignin (10.2 cal1/ bin Chen: Conceptualization, Formal analysis, Funding acquisition.
2
⋅cm–3/2) was closer to that of poly lactic acid (PLA) 9.88 cal1/2⋅cm–3/2 Xuejin Zhang: Funding acquisition, Supervision.
when compared to unacetylated pristine lignin (12.9 cal1/2⋅cm–3/2)
[111]. As expected, the acetylated lignin demonstrated enhanced Declaration of competing interest
compatibility with PLA, and the resulting acetylated lignin/PLA film
exhibited well-balanced optical and mechanical properties [111,112]. The authors declare that they have no known competing financial
Alkaline lignin and polyvinyl alcohol (PVA) with total solubility pa­ interests or personal relationships that could have appeared to influence
rameters of 9.82 and 13.5 cal1/2⋅cm–3/2 were mixed to form new blends the work reported in this paper.
[73]. The total solubility parameter of lignin/PVA membranes were
8.56, 9.14, 8.14, and 8.00 cal1/2⋅cm–3/2 when the proportions of incor­ Acknowledgments
porated lignin were 10, 15, 20, and 25 %, respectively. Interestingly, the
composite membrane with 15 % lignin, which had the highest total The authors are grateful for the supporting of Research Start-up

9
Q. Ma et al. International Journal of Biological Macromolecules 256 (2024) 128506

Foundation (Grant No. F701119L02) and Basic Research Foundation [28] J. Guo, X. Chen, J. Wang, Y. He, H. Xie, Q. Zheng, The influence of compatibility
on the structure and properties of PLA/lignin biocomposites by chemical
(Grant No. 2023QN056) of Zhejiang University of Science and
modification, Polymers 12, (1) (2019) 56.
Technology. [29] C. Pouteau, S. Baumberger, B. Cathala, P. Dole, Lignin–polymer blends:
evaluation of compatibility by image analysis, C. R. Biol. 327 (9–10) (2004)
935–943.
Appendix A. Supplementary data [30] M. Elodie, A. Filipe, D. Sousa, A.J.M. Valente, A. Romano, F.E. Antunes,
B. Medronho, Dissolution of Kraft lignin in alkaline solutions, Int. J. Biol.
Supplementary data to this article can be found online at https://doi. Macromol. 148 (2020) 688–695.
[31] J.H. Hildebrand, R.L. Scott, The Solubility of Nonelectrolytes, 3rd ed., Dover
org/10.1016/j.ijbiomac.2023.128506.
Publications, New York, 1950.
[32] C.M. Hansen, Cohesion parameters for surfaces, pigments, and fillers, Surf. Coat.
References Int. 80 (8) (1997) 386–391.
[33] C.M. Hansen, Hansen solubility parameters: a user’s handbook, in: C.M. Hansen
(Ed.), Hansen Solubility Parameters: A User’s Handbook, CRC Press, Florida, US,
[1] J. Ralph, C. Lapierre, W. Boerjan, Lignin structure and its engineering, Curr.
2002.
Opin. Biotechnol. 56 (2019) 240–249.
[34] J.S. Howell, B.O. Stephens, D.S. Boucher, Convex solubility parameters for
[2] C. Chio, M. Sain, W. Qin, Lignin utilization: a review of lignin depolymerization
polymers, J. Polym. Sci. B 53 (2015) 1089–1097.
from various aspects, Renew. Sustain. Energy Rev. 107 (2019) 232–249.
[35] O. Ajao, M. Benali, N. El Mehdi, Experimental and computer aided solubility
[3] J. Wen, S. Sun, B. Xue, R. Sun, Recent advances in characterization of lignin
quantification of diverse lignins and performance prediction, Chem. Commun. 57
polymer by solution-state nuclear magnetic resonance (NMR) methodology,
(14) (2021) 1782–1785.
Materials 6 (1) (2013) 359–391.
[36] P. Atkins, J.D. Paula, Atkins’ Physical Chemistry, Oxford University Press,
[4] V.K. Ponnusamy, D.D. Nguyen, J. Dharmaraja, S. Shobana, J.R. Banu, R.
Wellington Square, Oxford, UK, 2006.
G. Saratale, S.W. Chang, G. Kumar, A review on lignin structure, pretreatments,
[37] J.W. Gooch, Flory-Huggins Theory, Springer New York, New York, NY, 2011.
fermentation reactions and biorefinery potential, Bioresour. Technol. 271 (2019)
[38] E. Langer, K. Bortel, S. Waskiewicz, M. Lenartowicz-Klik, Essential quality
462–472.
parameters of plasticizers, Plasticizers Derived from Post-Consumer PET (2020)
[5] E.A. Agustiany, M. Rasyidur Ridho, D.N Rahmi .M., E.W. Madyaratri, F. Falah, M.
45–100.
A.R. Lubis, N.N. Solihat, F.A. Syamani, P. Karungamye, A. Sohail, Recent
[39] P.J. Flory, In Principles of Polymer Chemistry, Cornell university press, 1953.
developments in lignin modification and its application in lignin-based green
[40] M. Muthukumar, S.F. Edwards, Chain statistics and scaling concepts,
composites: a review, Polym. Compos. 43 (8) (2022) 4848–4865.
Comprehensive Polymer Science Supplements 2 (9) (1989) 1–47.
[6] R.C. Sun, Lignin source and structural characterization, ChemSusChem 13 (17)
[41] T. Tadros, Flory-Huggins interaction parameter, in: T. Tadros (Ed.), Encyclopedia
(2020) 4385–4393.
of Colloid and Interface Science, Springer Berlin Heidelberg, Berlin, Heidelberg,
[7] A.J. Ragauskas, G.T. Beckham, M.J. Biddy, R. Chandra, F. Chen, M.F. Davis, B.
2013, pp. 523–524.
H. Davison, R.A. Dixon, P. Gilna, M. Keller, Lignin valorization: improving lignin
[42] J.F. Kadla, S. Kubo, Miscibility and hydrogen bonding in blends of poly (ethylene
processing in the biorefinery, Science 344 (6185) (2014), 1246843.
oxide) and Kraft lignin, Macromolecules 36 (20) (2003) 7803–7811.
[8] A.E. Kazzaz, P. Fatehi, Technical lignin and its potential modification routes: a
[43] A.F.M. Barton, Solubility parameters, Chem. Rev. 75 (6) (1974) 731–753.
mini-review, Industrial Crops Products 154 (2020), 112732.
[44] R.F. Blanks, J. Prausnitz, Thermodynamics of polymer solubility in polar and
[9] M.U. Khan, B.K. Ahring, Lignin degradation under anaerobic digestion: influence
nonpolar systems, Ind. Eng. Chem. Fundam. 3 (1) (1964) 1–8.
of lignin modifications-a review, Biomass Bioenergy 128 (2019), 105325.
[45] W. Zeng, Y. Du, Y. Xue, H.L. Frisch, Solubility parameters, in: J.E. Mark (Ed.),
[10] D. Bourbiaux, J. Pu, F. Rataboul, L. Djakovitch, C. Geantet, D. Laurenti, Reductive
Physical Properties of Polymers Handbook, Springer New York, New York, NY,
or oxidative catalytic lignin depolymerization: an overview of recent advances,
2007, pp. 289–303.
Catal. Today 373 (2021) 24–37.
[46] M. Belmares, M. Blanco, W.A.G. Iii, R.B. Ross, G. Caldwell, S.H. Chou, J. Pham, P.
[11] Z. Sun, B. Fridrich, A. De Santi, S. Elangovan, K. Barta, Bright side of lignin
M. Olofson, C. Thomas, Hildebrand and Hansen solubility parameters from
depolymerization: toward new platform chemicals, Chem. Rev. 118 (2) (2018)
molecular dynamics with applications to electronic nose polymer sensors,
614–678.
J. Comput. Chem. 25 (15) (2004) 1814–1826.
[12] Y. Han, B.A. Simmons, S. Singh, Perspective on oligomeric products from lignin
[47] S. Venkatram, C. Kim, A. Chandrasekaran, R. Ramprasad, Critical assessment of
depolymerization: their generation, identification, and further valorization, Ind.
the Hildebrand and Hansen solubility parameters for polymers, J. Chem. Inf.
Chem. Mater. 1 (2) (2023) 207–223.
Model. 59 (10) (2019) 4188–4194.
[13] A. Kumar, J. Kumar, T. Bhaskar, Utilization of lignin: a sustainable and eco-
[48] C.M. Hansen, The universality of the solubility parameter, Ind. Eng. Chem. Prod.
friendly approach, J. Energy Inst. 93 (1) (2020) 235–271.
Res. Dev. 8 (1) (1969) 2–11.
[14] J. Gracia-Vitoria, S.C. Gándara, E. Feghali, P. Ortiz, W. Eevers, K. Triantafyllidis,
[49] C.M. Hansen, A. Björkman, The ultrastructure of wood from a solubility
K. Vanbroekhoven, The chemical and physical properties of lignin bio-oils, facts
parameter point of view, Holzforschung 52 (4) (1998) 335–344.
and needs, Curr. Opin. Green Sustain. Chemistry 40 (2023), 100781.
[50] D. Patterson, Free volume and polymer solubility. A qualitative view,
[15] M.N. Collins, M. Nechifor, F. Tanasă, M. Zănoagă, A. McLoughlin, M.A. Stróżyk,
Macromolecules 2 (6) (1969) 672–677.
M. Culebras, C.-A. Teacă, Valorization of lignin in polymer and composite systems
[51] L.P. Mcmaster, Aspects of polymer solution thermodynamics, Macromolecules 6
for advanced engineering applications–a review, Int. J. Biol. Macromol. 131
(5) (1973) 760–773.
(2019) 828–849.
[52] D. Patterson, Role of free volume changes in polymer solution thermodynamics,
[16] P. Jędrzejczak, M.N. Collins, T. Jesionowski, Ł. Klapiszewski, The role of lignin
in: Journal of Polymer Science Part C: Polymer Symposia, 1967, Wiley Online
and lignin-based materials in sustainable construction–a comprehensive review,
Library, 1967, pp. 3379–3389.
Int. J. Biol. Macromol. 187 (2021) 624–650.
[53] D.W. Van Krevelen, Properties of Polymers: Their Correlation with Chemical
[17] V.K. Thakur, M.K. Thakur, P. Raghavan, M.R. Kessler, Progress in green polymer
Structure; their Numerical Estimation and Prediction from AdditiVe Group
composites from lignin for multifunctional applications: a review, ACS Sustain.
Contributions, Elsevier Science, Amsterdam, 1997.
Chem. Eng. 2 (5) (2014) 1072–1092.
[54] G. Gogoi, S. Hazarika, Dissolution of lignocellulosic biomass in ionic liquid-water
[18] G. Yang, Z. Gong, X. Luo, L. Chen, L. Shuai, Bonding wood with uncondensed
media: interpretation from solubility parameter concept, Korean J. Chem. Eng. 36
lignins as adhesives, Nature (2023) 1–2.
(10) (2019) 1626–1636.
[19] Q. Tang, Y. Qian, D. Yang, X. Qiu, Y. Qin, M. Zhou, Lignin-based nanoparticles: a
[55] O. Segarceanu, M. Leca, Improved method to calculate Hansen solubility
review on their preparations and applications, Polymers 12 (11) (2020) 2471.
parameters of a polymer, Prog. Org. Coat. 37 (1997) 307–310.
[20] P.S. Chauhan, Lignin nanoparticles: eco-friendly and versatile tool for new era,
[56] M. Mohan, K. Huang, V.R. Pidatala, B.A. Simmons, S. Singh, K.L. Sale, J.
Bioresour. Technol. Rep. 9 (2020), 100374.
M. Gladden, Prediction of solubility parameters of lignin and ionic liquids using
[21] W.D.H. Schneider, A.J.P. Dillon, M. Camassola, Lignin nanoparticles enter the
multi-resolution simulation approaches, Green Chem. 24 (3) (2022) 1165–1176.
scene: a promising versatile green tool for multiple applications, Biotechnol. Adv.
[57] D. Patterson, Introduction to thermodynamics of polymer solubility, J. Paint
47 (2021), 107685.
Technol. 41 (536) (1969) 489–493.
[22] C.G. Yoo, A.J. Ragauskas, Lignin Utilization Strategies: From Processing to
[58] C.M. Hansen, A. Beerbower, “Solubility Parameters” in Kirk-Othmer
Applications vol. 1377, American Chemical Society, Washington, DC, USA, 2021.
Encyclopaedia of Chemical Technology. New York, 1971.
[23] K. Khitrin, S. Fuks, S. Khitrin, S. Kazienkov, D. Meteleva, Lignin utilization
[59] A. Goetz, Tables of Experimental Dipole Moments, W.H. Freeman, San Francisco,
options and methods, Russ. J. Gen. Chem. 82 (2012) 977–984.
1963.
[24] Y. Xu, C. Ma, S. Sun, C. Zhang, J. Wen, T. Yuan, Fractionation and evaluation of
[60] K.L. Hoy, Solubility parameter as a design parameter for water borne polymers
light-colored lignin extracted from bamboo shoot shells using hydrated deep
and coatings, J. Coated Fabrics 19 (1) (1989) 53–67.
eutectic solvents, Bioresour. Technol. 129679 (2023).
[61] D.W. Van Krevelen, P.J. Hoftyzer, Properties of Polymer–Their Estimation and
[25] T. Pang, G. Wang, H. Sun, W. Sui, C. Si, Lignin fractionation: effective strategy to
Correlation with Chemical Structure, Elsevier, Amsterdam, 1976.
reduce molecule weight dependent heterogeneity for upgraded lignin
[62] A. Beerbower, Environmental capability of liquid lubricants, NASA Special
valorization, Ind. Crops Prod. 165 (2021), 113442.
Publication 318 (1973) 365–431.
[26] S. Stiefel, C. Marks, T. Schmidt, S. Hanisch, G. Spalding, M. Wessling, Overcoming
[63] R.F. Fedors, A method for estimating both the solubility parameters and molar
lignin heterogeneity: reliably characterizing the cleavage of technical lignin,
volumes of liquids, Polym. Eng. Sci. 14 (2) (1974) 147–154.
Green Chem. 18 (2) (2016) 531–540.
[27] W.L. Archer, Determination of Hansen solubility parameters for selected cellulose
ether derivatives, Ind. Eng. Chem. Res. 30 (10) (1991) 2292–2298.

10
Q. Ma et al. International Journal of Biological Macromolecules 256 (2024) 128506

[64] Q. Ma, L. Wang, H. Zhai, H. Ren, Lignin dissolution model in formic acid–acetic [91] P. Weerachanchai, S.S.J. Leong, M.W. Chang, C.B. Ching, J.-M. Lee, Improvement
acid–water systems based on lignin chemical structure, Int. J. Biol. Macromol. of biomass properties by pretreatment with ionic liquids for bioconversion
182 (2021) 51–58. process, Bioresour. Technol. 111 (2012) 453–459.
[65] Y. Ni, Q. Hu, Alcell® lignin solubility in ethanol–water mixtures, J. Appl. Polym. [92] P. Weerachanchai, K.H. Lim, J.-M. Lee, Influence of organic solvent on the
Sci. 57 (12) (1995) 1441–1446. separation of an ionic liquid from a lignin–ionic liquid mixture, Bioresour.
[66] G.C. Vebber, P. Pranke, C.N. Pereira, Calculating Hansen solubility parameters of Technol. 156 (2014) 404–407.
polymers with genetic algorithms, J. Appl. Polym. Sci. 131 (1) (2014) [93] Y. Wang, L. Wei, K. Li, Y. Ma, N. Ma, S. Ding, Linlin Wang, D. Zhao, B. Yan,
39696–39707. W. Wan, Q. Zhang, X. Wang, J. Wang, H. Li, Lignin dissolution in
[67] C. Etxabarren, M. Iriarte, C. Uriarte, A. Etxeberría, J.J. Iruin, Polymer–solvent dialkylimidazolium-based ionic liquid-water mixtures, Bioresour. Technol. 170
interaction parameters in polymer solutions at high polymer concentrations, (2014) 499–505.
J. Chromatogr. A 969 (1) (2002) 245–254. [94] T. Yuan, J. He, F. Xu, R. Sun, Fractionation and physico-chemical analysis of
[68] L.P. Novo, A.A. Curvelo, Hansen solubility parameters: a tool for solvent selection degraded lignins from the black liquor of Eucalyptus pellita KP-AQ pulping,
for organosolv delignification, Ind. Eng. Chem. Res. 58 (31) (2019) Polym. Degrad. Stab. 94 (7) (2009) 1142–1150.
14520–14527. [95] M. Zhou, O.A. Fakayode, A.E. Ahmed Yagoub, Q. Ji, C. Zhou, Lignin fractionation
[69] H. Ni, S. Ren, G. Fang, Y. Ma, Determination of alkali lignin solubility parameters from lignocellulosic biomass using deep eutectic solvents and its valorization,
by inverse gas chromatography and Hansen solubility parameters, BioResources Renew. Sustain. Energy Rev. 156 (2022), 111986.
11 (2) (2016) 4353–4368. [96] C. Lai, C. Yang, Y. Jia, X. Xu, K. Wang, Q. Yong, Lignin fractionation to realize the
[70] M. Weng, Determination of the Hansen solubility parameters with a novel comprehensive elucidation of structure-inhibition relationship of lignins in
optimization method, J. Appl. Polym. Sci. 133, (16) (2015) 43328. enzymatic hydrolysis, Bioresour. Technol. 355 (2022), 127255.
[71] M. Díaz de los Ríos, E. Hernández Ramos, Determination of the Hansen solubility [97] C.G. Boeriu, R.J. Gosselink, A.E. Frissen, J. Stoutjesdijk, F. Peter, Fractionation of
parameters and the Hansen sphere radius with the aid of the solver add-in of five technical lignins by selective extraction in green solvents and
Microsoft Excel, SN Appl. Sci. 2 (4) (2020) 676. characterisation of isolated fractions, Ind. Crops Prod. 62 (2014) 481–490.
[72] J. Ruwoldt, M. Tanase-Opedal, K. Syverud, Ultraviolet spectrophotometry of [98] J. Domínguez-Robles, T. Tamminen, T. Liitia, M.A.S. Peresin, A. Rodríguez, A.-
lignin revisited: exploring solvents with low harmfulness, lignin purity, Hansen S. Jaaskelainen, Aqueous acetone fractionation of Kraft, organosolv and soda
solubility parameter, and determination of phenolic hydroxyl groups, ACS Omega lignins, Int. J. Biol. Macromol. 106 (2018) 979–987.
7 (50) (2022) 46371–46383. [99] O. Ajao, J. Jeaidi, M. Benali, O.Y. Abdelaziz, C.P. Hulteberg, Green solvents-based
[73] G. Zhao, H. Ni, S. Ren, G. Fang, Correlation between solubility parameters and fractionation process for Kraft lignin with controlled dispersity and molecular
properties of alkali lignin/PVA composites, Polymers 10 (2018) 3. weight, Bioresour. Technol. 291 (2019), 121799.
[74] J. Quesada-Medina, F.J. López-Cremades, P. Olivares-Carrillo, Organosolv [100] X. Jiang, D. Savithri, X. Du, S. Pawar, H. Jameel, H.-M. Chang, X. Zhou,
extraction of lignin from hydrolyzed almond shells and application of the δ-value Fractionation and characterization of Kraft lignin by sequential precipitation with
theory, Bioresour. Technol. 101 (21) (2010) 8252–8260. various organic solvents, ACS Sustainable Chemistry and Engineering 5 (1)
[75] Y. Ye, Y. Liu, J. Chang, Application of solubility parameter theory to organosolv (2017) 835–842.
extraction of lignin from enzymatically hydrolyzed cornstalks, BioResources 9 (2) [101] A.S. Jaaskelainen, T. Liitia, A. Mikkelson, T. Tamminen, Aqueous organic solvent
(2014) 3417–3427. fractionation as means to improve lignin homogeneity and purity, Ind. Crop.
[76] Q. Wang, K. Chen, J. Li, G. Yang, S. Liu, J. Xu, The solubility of lignin from Prod. 103 (2017) 51–58.
bagasse in a 1, 4-butanediol/water system, BioResources 6 (3) (2011) [102] J.-Y. Kim, S.Y. Park, J.H. Lee, I.-G. Choi, J.W. Choi, Sequential solvent
3034–3043. fractionation of lignin for selective production of monoaromatics by Ru catalyzed
[77] W. Thielemans, R.P. Wool, Lignin esters for use in unsaturated thermosets: lignin ethanolysis, RSC Adv. 7 (84) (2017) 53117–53125.
modification and solubility modeling, Biomacromolecules 6 (4) (2005) [103] A. Arshanitsa, J. Ponomarenko, T. Dizhbite, A. Andersone, R.J. Gosselink, J. van
1895–1905. der Putten, M. Lauberts, G. Telysheva, Fractionation of technical lignins as a tool
[78] J. Sameni, S. Krigstin, M. Sain, Solubility of lignin and acetylated lignin in organic for improvement of their antioxidant properties, J. Anal. Appl. Pyrolysis 103
solvents, BioResources 12 (1) (2017) 1548–1565. (2013) 78–85.
[79] G. Zhao, X. Liu, S. Ren, W. Tan, G. Fang, Quantitative comparison of surface [104] M. Li, S. Sun, F. Xu, R. Sun, Sequential solvent fractionation of heterogeneous
properties of enzymatic hydrolysis lignin before and after degradation, Ind. Crops bamboo organosolv lignin for value-added application, Sep. Purif. Technol. 101
Prod. 125 (2018) 468–472. (2012) 18–25.
[80] K. Vanommeslaeghe, E. Hatcher, C. Acharya, S. Kundu, S. Zhong, J. Shim, [105] X. Li, M. Li, Discrepancy of lignin dissolution from eucalyptus during formic acid
E. Darian, O. Guvench, P. Lopes, I. Vorobyov, A.D.M. Jr, CHARMM general force fractionation, Int. J. Biol. Macromol. 164 (2020) 4662–4670.
field: a force field for drug-like molecules compatible with the CHARMM all-atom [106] A. Duval, F. Vilaplana, C. Crestini, M. Lawoko, Solvent screening for the
additive biological force fields, J. Comput. Chem. 31 (4) (2010) 671–690. fractionation of industrial Kraft lignin, Hozforschung 70 (1) (2016) 11–20.
[81] J. Sameni, S. Krigstin, M. Sain, Characterization of lignins isolated from industrial [107] G.F. Bass, T.H. Epps, Recent developments towards performance-enhancing
residues and their beneficial uses, BioResources 11 (4) (2016) 8435–8456. lignin-based polymers, Polym. Chem. 12 (29) (2021) 4130–4158.
[82] W.C.d. Ribeiro, P.F.M. Martinez, V. Lobosco, Solubility parameters analysis of [108] D. Feldman, Lignin and its polyblends-a review, Pure Appl. Chem. 8 (9) (1995)
Eucalyptus urograndis Kraft lignin, Bioresources 15 (2020) 8577–8600. 81–99.
[83] Huy Quang Lê, A. Zaitseva, J.-P. Pokki, M. Ståhl, V. Alopaeus, H. Sixta, Solubility [109] J. Kim, T.V.T. Nguyen, Y.H. Kim, F. Hollmann, C.B. Park, Lignin as a
of Organosolv lignin in g-Valerolactone/water binary mixtures, ChemSusChem 9 multifunctional photocatalyst for solar-powered biocatalytic oxyfunctionalization
(20) (2016) 2939–2947. of C–H bonds, Nature Synthesis 1 (3) (2022) 217–226.
[84] D. Raikwar, S. Majumdar, D. Shee, Thermocatalytic depolymerization of Kraft [110] J.J. Meister, D.R. Patil, Solvent effects and initiation mechanisms for graft
lignin to guaiacols using HZSM-5 in alkaline water–THF co-solvent: a realistic polymerization on pine lignin, Macromolecules 18 (8) (1985) 1559–1564.
approach, Green Chem. 21 (14) (2019) 3864–3881. [111] Y. Kim, J. Suhr, H.-W. Seo, H. Sun, S. Kim, I.-K. Park, S.-H. Kim, Y. Lee, K.-J. Kim,
[85] N. Wang, B. Xu, X. Wang, J. Lang, H. Zhang, Insights into the mechanism of lignin J.-D. Nam, All biomass and UV protective composite composed of compatibilized
dissolution via deep eutectic solvents by using Hansen solubility theory, J. Mol. lignin and poly (lactic-acid), Sci. Rep. 7, (1) (2017) 43596.
Liq. 366 (2022), 120294. [112] H. Sadeghifar, A. Ragauskas, Lignin as a UV light blocker-a review, Polymers 12,
[86] A. Adjaoud, R. Dieden, P. Verge, Sustainable esterification of a soda lignin with (5) (2020) 1134.
phloretic acid, Polymers 13, (4) (2021) 637. [113] G. Zhao, H. Ni, L. Jia, S. Ren, G. Fang, Quantitative analysis of relationship
[87] F. Cheng, T. Ouyang, J. Sun, T. Jiang, J. Luo, Using solubility parameter analysis between Hansen solubility parameters and properties of alkali lignin/
to understand delignification of poplar and rice straw with catalyzed organosolv acrylonitrile-butadiene-styrene blends, ACS Omega 3 (8) (2018) 9722–9728.
fractionation processes, Bioresources 14 (1) (2019) 486–499. [114] H. Haridevan, D.A.C. Evans, D.J. Martin, P.K. Annamalai, Rational analysis of
[88] P. Weerachanchai, S.K. Kwak, J.M. Lee, Effects of solubility properties of solvents dispersion and solubility of Kraft lignin in polyols for polyurethanes, Industrial
and biomass on biomass pretreatment, Bioresour. Technol. 170 (2014) 160–166. Crops and Products 185 (2022), 115129.
[89] Q. Zhang, X. Tan, W. Wang, Q. Yu, Q. Wang, C. Miao, Y. Guo, X. Zhuang, Z. Yuan, [115] H.J.M. Grünbauer, D. Areskogh, A Composition in the Form of a Lignin Polyol, A
Screening solvents based on Hansen solubility parameter theory to depolymerize Method for the Production Thereof and Use Thereof, 2013.
lignocellulosic biomass efficiently under low temperature, ACS Sustain. Chem. [116] D. Barana, M. Orlandi, L. Zoia, L. Castellani, T. Hanel, C. Bolck, R. Gosselink,
Eng. 7 (9) (2019) 8678–8686. Lignin based functional additives for natural rubber, ACS Sustain. Chem. Eng. 6
[90] F. Cheng, X. Zhao, Y. Hu, Lignocellulosic biomass delignification using aqueous (9) (2018) 11843–11852.
alcohol solutions with the catalysis of acidic ionic liquids: a comparison study of
solvents, Bioresour. Technol. 249 (2018) 969–975.

11

You might also like