Anda di halaman 1dari 16

The Journal of Supercritical Fluids 133 (2018) 550–565

Contents lists available at ScienceDirect

The Journal of Supercritical Fluids


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

Review

Understanding biomass fractionation in subcritical & supercritical water T



María José Cocero , Álvaro Cabeza, Nerea Abad, Tijana Adamovic, Luis Vaquerizo,
Celia M. Martínez, María Victoria Pazo-Cepeda
High Pressure Processes Group, Department of Chemical Engineering and Environmental Technology, University of Valladolid (Spain), Doctor Mergelina s/n, 47011,
Valladolid, Spain

G RA P H I C A L AB S T R A C T

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

Keywords: Biomass fractionation into its individual building blocks poses a major challenge to the biorefinery concept. The
Hemicellulose recalcitrance of the lignocellulose matrix and the high crystallinity of cellulose make typical feed stocks difficult
Bound phenolics to separate into their components. Hydrothermal processing fractionates biomass by its hydrolysis. However, a
Cellulose dissolution deep knowledge of hydrolysis principles is required since an inappropriate selection of the operating parameters
Hydrolysis
such as an excessive temperature and a long residence times causes dramatic selectivity losses. This review is
Lignin
divided in four main sections which present the fundamentals of lignocellulosic biomass fractionation in
hemicelluloses, cellulose and lignin. As the biomass structure plays an important role, a section to study the
extraction of the linked phenols that joint lignin and hemicelluloses is included.

1. Introduction bio-sourced commodity chemicals, but due do their differing chemical


functionalities (lignin made up of linked aromatic units, hemicellulose
Shifting the chemical industry away from petrochemical feedstocks of C5 sugars and cellulose of C6 sugars) separation steps are necessary
towards renewable, bio-based chemicals and materials is a long-term to isolate the appropriate fraction and break it into its individual
strategy of the European Union. This “biorefinery” concept, despite building blocks (e.g. sugars for cellulose/hemicellulose and aromatic
being proposed as early as the late 1980s, has still not come to fruition units for lignin) (Fig. 1).
because the cost and complexity of processing biomass to generate This fractionation of biomass into its individual building blocks
practical, usable, saleable feedstocks makes it unfeasible. poses a major challenge to the biorefinery concept, because the re-
Lignocellulose is the most abundant, cheapest and easiest grown calcitrance of the lignocellulose matrix and high crystallinity of cellu-
form of biomass, and it is composed of three main fractions: cellulose lose makes typical feedstocks difficult to separate into their compo-
(40–50%), hemicellulose (25–35%) and lignin (10–30%), in addition to nents. For this reason, it typically requires long reaction times (from
minor compounds. These fractions represent potential feedstocks for 30 min for the hydrothermal hydrolysis to 24–70 h for enzymatic


Corresponding author.
E-mail address: mjcocero@iq.uva.es (M.J. Cocero).

http://dx.doi.org/10.1016/j.supflu.2017.08.012
Received 31 May 2017; Received in revised form 21 August 2017; Accepted 22 August 2017
Available online 26 August 2017
0896-8446/ © 2017 Elsevier B.V. All rights reserved.
M.J. Cocero et al. The Journal of Supercritical Fluids 133 (2018) 550–565

Fig. 1. Lignocellulosic biomass structure.

Fig. 2. Subcritical and supercritical water properties around the cri-


tical point.

hydrolysis) and the presence of strong reagents (sodium hydroxide and rapid and presents a mean to achieve significantly more process in-
sodium sulfide during Kraft pulping, for example). This leads to de- tensive fractionation of biomass.
gradation of the non-cellulosic fractions as well as large volumes of Reaction speed – whilst being an advantage to process intensifica-
effluent which requires expensive treatment to reduce environmental tion – is also a significant disadvantage to selectivity at longer reaction
load. times, leading to degradation of hydrolysis products and resulting in
To truly harness the potential of the biorefinery concept, this frac- complex reaction mixtures. This degradation and mixture complexity
tionation step needs to be revolutionized. It needs to be considerably leads to inefficient recovery of biomass derived products and inter-
more process intensive (ideally seconds per unit volume of biomass -as mediates. There is therefore a need for understanding the hydrothermal
opposed to minutes or hours) to enable modular units to deal with large fractionation processes to improve processes selectivity, which can
volumes of biomass at decentralized locations. It must not involve the harness the potential of subcritical and supercritical water fractiona-
use of harsh reagents in order to minimize environmental impact and tion.
cost, whilst maintaining quality of the fractions. Even under water’s critical point, certain fractions of biomass face
Water above its critical point (Tc 374 °C, 22 MPa), is an alternative reactions that proceed too rapidly to be controlled by conventional
solvent for dissolution/hydrolysis of biomass. Its low viscosity and high methods. For instance, lignin undergoes rapid hydrolysis and sub-
diffusivity facilitate the penetration of water into the complex structure sequent hydrolysis product conversion in less than 1 s at 350 °C. Whilst
of the lignocellulosic matrix, whilst its low dielectric constant, similar poor selectivity is common to both sub- and supercritical water (SCW),
to non-polar organic solvents, enhances solubility of organic com- there are some significant differences between the reaction media –
pounds. Physical properties of water (such as density, ionic product, most notably the difference in ionic product of water (for example the H
dielectric constant) can be finely tuned by varying temperature and +/OH− concentration at 300 °C and 22 MPa is around 3·10−6 mol L−1
pressure. At these conditions, the hydrolysis of biomass fractions is vs 3·10−10 at 400 and 22 MPa) which means that subcritical water has a

551
M.J. Cocero et al. The Journal of Supercritical Fluids 133 (2018) 550–565

Fig. 3. Liquid profiles at the output of a packed bed reactor during a hy-
drothermal extraction process: (a) TOC evolution, (b) molecular weight
evolution (Mw) when both stages are presents and molecular weight
evolution when only stage 2 is present.

higher concentration of ions ([H+] and [OH−]) thus favoring ionic 240 °C) [4–6]. Since hemicelluloses have some potentially acidic groups
reactions vs the radical reactions that are prevalent under SCW condi- (acetyl groups among others), it can be recovered by Kraft pulping.
tions (Fig. 2). However, the use of this technique leads to a degradation of hemi-
This manuscript studies the lignocellulosic biomass fundamentals cellulose, so a different technique is required to obtain it with a high
fractionation in subcritical and supercritical water, in order to improve quality. Thus, hydrothermal extraction would be one of the most pro-
the selectivity of the hydrothermal biomass fractionation. The manu- mising options since it only requires water and mild temperatures
script presents four main sections to presents the fractionation of bio- (160–210 °C) to extract it [7,8]. If the operational temperature is
mass in hemicellulose, sugars and lignin. As the biomass structure plays around 180 °C, 60% of the initial hemicellulose can be recovered as
an important role, a section to study the linked phenols that joint lignin oligomers and sugars [4,9]. Higher yields can be obtained if tempera-
and hemicelluloses is included. ture increases but undesired degradation products appear [10,11].
However, hemicellulose can be recovered also at low temperatures
(90 °C) if the operating time is high enough (days) [12]. Hemicellulose
2. Hemicellulose(s) fractionation fundamentals extraction has been performed in both systems, batch and packed bed
reactors. Therefore, it should be also marked that two different oper-
Hemicellulose is a biopolymer present in lignocellulosic materials ating times can be defined, the solid and liquid time. The former is the
that acts as a connection between the fibbers (cellulose) and the 3-di- time used to treat the solid. The liquid time has the same value as the
mensional structure (lignin), constituting between 25% and 35% of the solid time in batch systems. However, it is fixed by the volumetric flow
whole biomass [1]. It is characterized by their amorphous structure and when semi-batch or continuous system are used, being the residence
by the fact that it is acetylated [2]. Regarding its composition, it is a time (see Appendix A in Supplementary material for more details about
biopolymer mainly composed of pentoses with few hexoses in between, the different residence times). Moreover, hemicellulose hydrothermal
with a maximum length around 200 or 300 monomeric sugars, which fractionation is a complex process that involves several physical phe-
makes it a renewable source of chemicals based on 5-carbon molecules. nomena [6,8,13,14], which are present as in batch as in continuous
The maximum molecular weight is lower than 70 kDa in most cases [3]. systems, and a good knowledge of them is mandatory for designing a
However, there are discrepancies between species. For instance, xylose profitable and sustainable hemicellulose extraction plant. These phe-
is the most common monomer in hemicelluloses of hardwood trees, nomena are:
while softwood trees are principally composed of mannans, like man-
nose [2]. In addition, there are two different types of hemicellulose
• Hemicellulose cleaving into decreasing molecular weight oligomers
from the extraction viewpoint: one hemicellulose easy to extract and
• Hemicellulose deacetylation (autohydrolysis)
another one that is associated with the fibbers of cellulose that can be
recovered only when cellulose is also removed (temperatures above
• Hemicelluloses dissolution and mass transfer between the solid and
552
M.J. Cocero et al. The Journal of Supercritical Fluids 133 (2018) 550–565

the liquid hydrogenation). These so-called “degradation products” can also be the
• Production of sugars & sugars degradation into furfural or other target [8]. For instance, furfural and its derivatives can be used as
substances fungicides or lubricants [21] while lactic acid is a precursor for bio-
• Porosity changes: extraction, swelling and biomass compaction degradable polymers production [22]. Therefore, to avoid/promote
sugar degradation the operating temperature and the volumetric flow
Additionally, and once the phenomenology is explained, a short (the less time in the reactor, the lower degradation [4,7,18,20]) are the
summary about the effect of the main operational variables on hemi- main involved variables. It is worth highlighting that when the reactor
cellulose selectivity is included. is a batch system, the liquid/solid ratio has the same role as residence
time.
2.1. Hemicellulose cleaving into decreasing molecular weight oligomers
2.3. Hemicellulose deacetylation
The following discussion is focused on the behavior observed in a
semi-continuous system since only globalized values can be obtained Hemicellulose deacetylation and cleaving take place in parallel,
from a batch reactor. Hemicellulose cleaving is one of the first phe- which is reflected in a releasing of acetic acid during the hemicellulose
nomena that takes place inside a chip or a particle of biomass. Due to extraction, decreasing the pH of the water. It should be remarked that
the mild operating temperature (e.g. 120–185 °C), the bonds between this acetic acid production mainly happens in the solid phase
the monomeric sugars start breaking randomly, producing progres- [6,14,23–25]. However, acetic acid can be obtained from sugar de-
sively shorter oligomers. This process continues until the moment in gradation in liquid phase too [8]. Similarly, uronic acid can be also
which the oligomer has a length low enough to be extracted from the released during the hydrothermal treatment [26,27]. Nevertheless, it is
solid by solubilization or dragging [6,14,15]. In this moment, both not completely clear if the pH change accelerates extraction or if this
phenomena are present, oligomer dissolution and oligomer cleaving, change is only a consequence of the extraction [10,12]. This phenom-
and two distinct stages can be differentiated: (1) solid oligomer enon is deeply related with the hemicellulose extraction process se-
cleaving, which is present from the beginning, and (2) solid & liquid lectivity since these acids are a source of protons that catalyze the
oligomer cleaving with hemicellulose dissolution. The fact that these cleaving and degradation reactions in liquid phase if the residence time
two phases are present at the same time explain why there is a delay in is high enough [6,14]. A statement that was verified by Song et al. [28],
the extraction profiles (Fig. 3a). Before this first soluble oligomer re- showing that degradation is much lower if the pH is maintained above
leasing, only raw material free sugars and a little number of cleaving 4–5. Moreover, the releasing of acetyl groups also means that the so-
products (small oligomers and monomers) could be removed. Never- lubility of the remained part of the hemicellulose would be lower since
theless, there are cases where no delay is present due to the biomass the capacity of linking by hydrogen bonds with water would be lower.
diversity [16]. This is possible when the initial hemicellulose length is Additionally, the steric hindrance also would be higher. Following this
so low that it is initially soluble or it is so acetylated that only stage 2 is idea, it can be seen in Fig. 4. that the minimum of the pH corresponds to
present. Therefore, if both stages are present, the molecular weight the maximum in the TOC profile. Thus, the oligomers involved in the
evolution during the extraction should have a maximum (the first so- stage (2) defined in Fig. 3 will be more soluble since they are smaller
luble oligomer) near the time (tm) when the concentration in the liquid but, at the same time, their solubility also decreases due to the lack of
reaches the highest value (Fig. 3b). After this molecular weight peak, it acetyl groups, explaining why the extraction is more difficult after the
would continuously decrease due to the cleaving. This behavior was maximum (decreasing slope). Moreover, extraction would also be
already observed in literature [15,16]. However, when only stage 2 is slower because the available amount of hemicellulose is much lower.
present the molecular weight would decrease with time. Thereby, the acetylation degree (and uronic content) is another vari-
To sum up, temperature, the molecular weight and de acetyl con- able to consider.
tents plays an essential role in hemicellulose extraction since they di-
rectly affect hemicellulose solubility. 2.4. Hemicellulose dissolution and mass transfer

2.2. Sugar production from the cleaving processes Hemicellulose solubility has been already discussed in Section 2.1.
External and internal mass transfer resistances can be present in this
As it was explained in the previous section, the cleaving can also process. External mass transfer can be enhanced by a proper mixing for
produce monomeric sugars and, if temperature is high enough, all the batch extraction in a stirred vessel or by high volumetric flows for semi-
hemicellulose could be converted into monomeric sugars. However, the continuous and continuous extraction in tubular beds. Furthermore,
operational conditions required to achieve a complete conversion are so internal diffusion problems can be minimized if the particle diameter is
high that they also imply sugar degradation. Gallina et al. [4] studied low enough. Rissanen et al. [10] indicated that the size of the extracted
the optimal conditions for the hydrothermal fractionation of eucalyptus hemicelluloses is highly dependent on the initial particle diameter.
in a semi-continuous reactor. They found that the optimum monomeric When using particle size below 2–3 mm hemicelluloses are relatively
sugar yield was at 185 °C (67.41%), starting to decrease at higher
temperatures. Yedro et al. [9] assessed the hemicellulose extraction
from holm oak in a batch system, obtaining that the highest monomeric
yield was at 170 °C (60%) and that degradation started at temperatures
as low as 150 °C. Rissanen et al. [10] analyzed the hydrothermal de-
gradation of spruce in the same reactor as Yedro et al. [9], reaching a
similar optimum. Sukhbaatar et al. [11] worked with sugarcane bagase
also in a batch system, being their monomeric yield optimum at 180 °C
and observing a huge degradation above 190 °C. The same biomass was
considered by Vallejos et al. [17] who reached the best monomeric
yield at 180 °C too (70%). Similar result were reported by Thomsen
et al. [18], dos Santos Rocha et al. [19] and Makishima et al. [20] for
wheat straw, sugarcane straw and corncob, respectively.
Focusing on direct sugar production is of interest since they can be
Fig. 4. Relation between the pH and the extracted biomass.
used to produce fuels (bioethanol) or chemicals (like xylitol via

553
M.J. Cocero et al. The Journal of Supercritical Fluids 133 (2018) 550–565

easily extracted from the matrix and higher molecular weights are (C5) decreases although the global yield increases (Yield tot). Ad-
obtained. By contrast, when a bigger particle size is used, e.g. 1 cm ditionally, it is also worth mentioning that if temperature goes above
chips, water comes inside the chip and the hydrolysis starts to take 240 °C, cellulose extraction would take place making it possible to re-
place. If deacetylation occurs, which is probable, then the acetic acid cover the hemicellulose fraction associated with cellulose and ex-
lowers the pH inside the chip accelerating the cleavage. Depending on plaining the increase in the yield. Nevertheless, this higher temperature
the internal and external mass transfer of the protons, the time with low would also mean the releasing of hexoses and cellulose oligomers and
pH inside particle will be different and the final molecular weight will the extraction of some lignin fractions [6,31,32], which would reduce
be also affected. In general, the smaller the particle sizes, the higher the the hemicellulose selectivity. On the other hand, not only does the solid
molecular weights. However, it has been widely demonstrated that time and the temperature control selectivity, but also the liquid re-
temperature and solid operating time have a bigger impact on the yield, sidence time and the operational pH affect it. If residence time is low
promoting extraction when they are increased [4,7,9–13,15–20,29]. enough, sugar degradation can be avoided (Fig. 5b). Regarding pH, it
has been demonstrated that degradation can be reduced up to around
2.5. Porosity changes: extraction, swelling and bed compaction 90% if pH is maintaining upper than 4–5 [33]. To sum up and provided
that a high hemicellulose selectivity is desired, temperatures around
During extraction, it is expected that the porosity of the bed in- 180 °C, pH above 4 and high volumetric flows (residence time lower
creases since a certain amount of mass is being removed. However, if than 4 min) are required.
this extraction was important enough, the bed could collapse, reducing
the porosity and making it impossible to continue the extraction (tre- 3. Bound phenolics
mendous pressures drops). Additionally, biomass also can swell
[6,13,14]. Therefore, a preliminary study about the behavior of the Phenolic compounds are of great interest due to their antioxidant
packed during the extraction should be required to avoid operational properties and potential health benefits that make them useful in nu-
problems. merous commercial applications in the food, cosmetic and pharma-
ceutical industries [34]. They can be present in the matrix in three
forms: free, soluble-bound (esterified to low molecular mass com-
2.6. Main variables on extraction selectivity: oligomers and monomers as pounds) or in an insoluble-bound form (esterified and/or etherified to
target compounds cell wall structural components) [35]. In cereal brans and in some
whole cereals, they are found approximately in the proportion
It can be concluded from the discussion done from Section 2.1–2.5 0.1:1:100, so at the beginning the free and soluble-bound moieties are
that the main variables to promote hemicellulose extraction yield are neglible, but after applying the treatment that release the bound moiety
the solid operational time and the operating temperature. Therefore, from the solid, both soluble forms are of great importance. Moreover,
the higher the temperature/time is, the bigger yield is obtained. How- the free and soluble-bound moieties present initially in the solid are
ever, if these two variables are too high, hemicellulose sugars and oli- removed with the extractives and in this section the release of the
gomers start to degrade, reducing the selectivity [4,30]. For instance, bound moiety is analyzed.
this temperature effect can be easily seen in Fig. 5a, where it can be Phenolic acids are a subgroup of phenolic compounds of great im-
checked that at temperatures higher than 180 °C the yield of pentoses portance in cereal grains, which are mostly found in the insoluble form.
Among them, ferulic and coumaric acid (two hydoxycinnamic acids)
are the most studied ones due to its abundance in the plant kingdom.
They provide rigidity to the cell wall [36] as they crosslink the sugar
moieties and also the lignin. Despite being minor compounds (around
0.5–1% in cereal brans) their high value can enhance the profitability of
the biorefinery and in order to recover them, it is necessary to hydro-
lyze the ester and/or ether bonds that maintain them attached.
Alkaline hydrolysis is the most common procedure used for their
releasing [37] but it is a non-selective method that also alter the whole
matrix. On the other hand, enzymatic hydrolysis can be selective if the
proper enzymes are chosen, but it must be taken into account the ma-
trix as well as the main phenolics present [38]. Feruloyl esterases (EC
3.1.1.73) are a family of enzymes able to cleavage the ester bond be-
tween the hydoxycinnamic acids and the sugar moieties, but in some
cases where the yield is low, the use a mixture of enzymes including
xylanases lead to increase it significantly [39]. In this context, the use of
pressurized water emerge as an interesting technique, which shorten
considerably the extraction time and avoid the use of solvents and/or
expensive enzymes. The optimum temperature, pressure and extraction
time vary depending on the raw material used and the main phenolic
present. As the major part of them are etherified to hemicelluloses and
the ether bonds are labile at 170 °C [35], temperatures higher than that
are commonly used and as consequence, the co-extraction of the
hemicellulose also takes place. After a certain temperature, or long
extraction times, the phenolic compounds start to decompose due to
their thermal degradation. The effects named above can be seen in
Fig. 6 where Pourali et al. [40] studied the effect of temperature on the
Fig. 5. Hydrothermal extraction of eucalyptus in a semi-continuous reactor (solid time of extraction yield of different phenolic acids from defatted rice bran, and
90 min): evolution of the hemicellulose extraction yield (Yield tot), the yield of hexoses obtained different optimum temperatures for each phenolic acid. In the
(C6), pentoses (C5) and degradation products for eucalyptus with temperature (a) and first stage, the ether links are being released and the co-extraction of the
residence time at 185 °C (b) [4].
hemicellulose enhance the extraction, favoring the solvent penetration

554
M.J. Cocero et al. The Journal of Supercritical Fluids 133 (2018) 550–565

Fig. 6. Effect of subcritical water temperature on the extraction of dif-


ferent phenolic compounds from defatted rice bran (residence
time = 10 min).
Obtained from Pourali et al. [40].

and the mass transfer; in the second stage, the degradation of the regarded as a necessary first step [46]. Hydrolysis reaction implies the
phenolic compounds turn out to be higher than the release, and so on, cleavage of a chemical bond by the addition of water [47]. However, as
the extraction yield decrease. mentioned above, cellulose is a water insoluble polymer so that it is not
Therefore, after selecting the optimum conditions to maximize the possible to simply dissolve and hydrolyze cellulose in water at ambient
extraction of the target compound(s), a purification step is necessary to conditions. As a result, the hydrolysis of cellulose in lignocellulosic
separate the phenolic fraction from the hemicellulose fraction. biomass usually involves the use of strong acids as catalysts [46], which
However, this optimum condition should be chosen avoiding or redu- cause a negative impact in the environment and yields a high con-
cing the alteration of the cellulose and lignin fractions, as in the bior- centration of degradation products. However, when using supercritical
efinery process they will be separated in a next step. water cellulose is more effectively converted to oligomers and
monomer sugars instead of yielding mainly degradation products.
Therefore, the objective of this section is to clarify the mechanisms
4. Cellulose
involved in both the dissolution and the hydrolysis of cellulose in water
as well as to discuss the influence of the key parameters which affect
Cellulose is a renewable, cheap and worldwide-distributed polymer
both processes.
with very promising applications for the future industries. Cellulose is a
linear-chain polysaccharide consisting on units of glucose linked by β-
1,4 glucosidic bonds, which structure is shown in Fig. 7. Its formula is 4.1. Cellulose dissolution
(C6H10O5)n, where n is the degree of polymerization and goes from
several hundred to many thousands of glucose units, depending on the The dissolution of cellulose in water have been explained [48–50]
raw material. The hydrogen bond network between OH groups in its from a thermodynamic point of view. The Gibbs free energy is a ther-
structure promotes the aggregation of cellulose chains forming fibrils modynamic magnitude commonly considered to analyze whether a
[41]. These fibrils are tough, water insoluble and forms the foundations chemical process is spontaneous or not. Its variation is expressed as a
of the plant cell walls [42–44]. For mass production, cellulose is ob- combination of the variation of the enthalpy and the variation of the
tained from plants, bacteria, algae and fungi via biosynthesis or in-vitro entropy of the system:
synthesis [45]. Although synthesized cellulose has numerous applica- ΔG = ΔH − T ΔS (1)
tions and uses, it is not the only way to make profit out of residual
biomass. “G” is the Gibbs free energy, “H” the enthalpy, “T” the temperature and
Lignocellulosic biomass is a complicated matrix formed by an in- “S” the entropy. When the variation of the Gibbs free energy is negative,
tricate network where cellulose, hemicellulose and lignin are linked and the process is spontaneous. In the reaction of two different compounds,
interacting together. Conversion of these biomass fractions into valu- the variation of enthalpy represents the heat of reaction or the heat of
able products is the key step for the success of the bio-based future mixing. In the combination of cellulose and water this parameter is
industries, where agricultural or industrial wastes usually regarded as almost negligible since no additional heat is generated or consumed.
worthless would be converted into chemicals, fuels and/or energy. In Therefore, the previous expression is reduced to:
order to get these valuable products from lignocellulosic biomass, de- ΔG = −T ΔS (2)
polymerization and hydrolysis of cellulose to monomer glucose is
Consequently, the dissolution and hydrolysis of cellulose in water is
carried out (spontaneous process) when the entropy variation is posi-
tive. From a structural point of view, the entropy of cellulose increases
when its molecular conformation changes from a rigid structure to a
more flexible one which benefits dissolution. Since cellulose structure is
characterize by its complexity and rigidity, at lower temperatures no
conformational changes will be produced, the entropy will not increase
nor the Gibbs free energy will decrease and therefore no dissolution will
be produced. Only in the cases in which the temperature is considerably
increased and therefore the internal energy of the structure, con-
Fig. 7. Cellulose formula.
formational changes could be produced.

555
M.J. Cocero et al. The Journal of Supercritical Fluids 133 (2018) 550–565

In literature, three main characteristics of the cellulose structure are value products from renewable resources such as biomass. As biomass is
considered of fundamental interest in its dissolution in water: a complex raw material and due to the lack of enough know-how in this
field, the majority of authors have started working with cellulose in-
1) The presence of intra and intermolecular hydrogen bonds stead of with biomass. When cellulose is mixed with water at high
[48,49,51]. Cellulose is constituted by glucose molecules joined temperatures, it is first dissolved and subsequently it reacts with the
together forming long fibbers which are connected by hydrogen water molecules present in the liquid medium producing the cleavage
bonds. This fact results in a rigid and cohesive structure which (hydrolysis) of the glycosidic bonds. As the cleavage of these bonds is
avoids the penetration of water molecules and consequently the not completely simultaneous nor instantaneous, first, oligosaccharides
dissolution of the structure. are generated which are then hydrolyzed to monosaccharides. Finally,
2) Cellulose is considered an amphiphilic molecule [48,49,52]. Its if the hydrolysis reaction is not stopped, the monosaccharides are de-
structure has both hydrophobic and hydrophilic zones as a con- graded to organic compounds such as acids [59,60]. As it has been
sequence of the orientation of its functional groups. While the hy- stated, it is fundamental both in dissolution and in hydrolysis to find the
droxyl groups located in equatorial position create the hydrophilic optimum pair of temperature and reaction time values in order to re-
regions, the axial glycosidic bonds produce hydrophobicity. This is duce the generation of undesired products.
considered the reason why the water molecules are not able to easily The experiments presented in literature are clearly divided in three
create hydrogen bonds with the cellulose which will produce its zones: subcritical region, vicinities of the critical point and supercritical
dissolution. region.
3) Crystallinity: crystallinity has always been considered a key para- In this paper, the subcritical region is considered the one in which
meter when analyzing the dissolution of cellulose in water [48,53]. the temperature remains below 320 °C. In this zone the dissolution and
The high crystallinity of the cellulose molecule is responsible of its subsequent hydrolysis is produced as a result of the consumption of
rigid structure avoiding conformational changes which could facil- superficial cellulose which is able to interact with water molecules
itate its dissolution. [61]. Furthermore, the cellulose which can be easily dissolved is the
one which was present in an amorphous state. Below 280 °C, it is ob-
Considering the lack of a robust model which explains the dissolu- served that the cellulose dissolution rate decreases with time since
tion of cellulose in water, several authors have performed experiments water is not able to dissolve crystalline cellulose once the amorphous
with the objective of analyzing the influence of the process parameters. cellulose has been already dissolved [55]. At temperatures between
From a structural point of view [54,55], studied the influence of the 280 °C and 320 °C increasing either the reaction time or the tempera-
raw cellulose used in the dissolution. They demonstrated that the cel- ture only increases the degradation of the cellulose, mostly amorphous,
lulose allomorph directly affects the dissolution process. For example, which has been already dissolved [55,62]. Therefore, working with low
although cellulose I is the most abundant type in nature, cellulose II is reaction times produces high DP (degree of polymerization) molecules
more stable [49]. Moreover, not only the cellulose type influences the [63]. At temperatures below 250 °C [64] proved that cellulose is dis-
dissolution, also the amount of water has to be considered [48,56]. solved but not hydrolyzed and therefore that it is possible to obtain high
Regarding to the crystallinity of the structure [57], analyzed the dis- DP molecules. However, the process is limited by the amount of
solution of cellulose after milling. Milling produces an amorphous amorphous cellulose available. In these cases reaction times in the order
structure which facilitates the action of water. They stated that the of hours are required which implies the operation in batch and semi-
critical factor is not the reduction of the particle diameter but the continuous reactors. Finally, milling the raw cellulose creates amor-
cleavage of the hydrogen bonds and the consequent generation of phous zones which can be easily dissolved, even at temperatures below
amorphous zones. Amorphous and semi-crystalline zones are easier to 230 °C, generating high DP molecules [57]. At this temperatures, no
be hydrolyzed since water molecules can avoid the hydrophobic zones modifications are observed in the solid residue when the cellulose
which are present in the structure as a consequence of the amphiphilic structure is crystalline instead of amorphous [65].
nature of the cellulose [58]. In the region near the critical point, when the reaction time is in-
From an operating point of view, the majority of experiments ana- creased, the dissolved cellulose is hydrolyzed to glucose and lately to
lyzed the process focusing in the variation of the pressure, the tem- degradation products. It has been experimentally demonstrated [51]
perature and the reaction time. As it has been explained in this section, that at temperatures between 320 °C and 330 °C (25 MPa) a transition
due to the physical structure and the nature of cellulose, its dissolution from a crystalline to an amorphous structure is produced. This transi-
is greatly limited by temperature. Common working temperatures tion explains the rapid dissolution of cellulose in water and the absence
usually range from 200 °C to more than 400 °C. Therefore, in order to of any swelling phenomena [51]. The fact that when cellulose and
maintain water in liquid or supercritical state when the working con- water react at these or higher temperatures during a short reaction time
ditions surpasses its critical point, (Tc = 374 °C, Pc = 22.1 MPa) the the final product obtained is cellulose II when the initial cellulose al-
pressure shall be increased. The analysis of the influence of pressure has lomorph is cellulose I is justified as a consequence of the higher stability
been studied by [53]. They proved that when the pressure is increased of cellulose II. When the temperature is increased above 330 °C cellu-
above 50 MPa (reaching pressures up to 700 MPa), even at relative low lose I is converted into amorphous cellulose. Then, when the tem-
temperatures the water molecules are able to enter inside the cellulose perature decreases, the amorphous cellulose is converted into the more
structure and swell the polymeric matrix which finally collapses. When stable cellulose II allomorph [59,61]. This is also confirmed working at
the pressure is only considered in order to maintain the water in liquid temperatures below 320 °C since only cellulose I is obtained [66].
or supercritical state, its influence is negligible and the fundamental Finally, in the supercritical region the dissolution and hydrolysis of
parameters to be considered are the temperature and the reaction time. cellulose when working at low concentrations is produced simulta-
An increase of temperature clearly benefits the dissolution of cellulose neously, in homogeneous phase and without mass transfer limitations
since it modifies its structure favoring the combination of cellulose and [53,67]. The transition between crystalline cellulose to amorphous
water molecules. However, it also accelerates its hydrolysis consuming cellulose at 330 °C, the high temperatures of reaction which produce
the cellulose which is being dissolved. Consequently, the only possibi- the cleavage of the hydrogen bonds [53,68] and the properties of su-
lity of dissolving cellulose and reduce its hydrolysis rate is selecting an percritical water such as high diffusivity, high density compared with
optimum combination of temperature and reaction time. In literature, water in vapor state and its ability to dissolve organic compounds,
the analysis of cellulose dissolution at high temperatures is generally observing the total dissolution of cellulose [69,70], explain the homo-
combined with hydrolysis studies. Hydrolysis is considered one of the geneity of the process. Recently it has been proved that when the
fundamental processes in green chemistry since it allows obtaining high concentration of cellulose is increased the dissolution and hydrolysis

556
M.J. Cocero et al. The Journal of Supercritical Fluids 133 (2018) 550–565

processes are not completely simultaneous nor homogeneous [71]. under those conditions, the cellulose hydrolysis kinetics are improved
and the glucose hydrolysis kinetics are slow enough so that using the
4.2. Cellulose hydrolysis sudden expansion micro-reactor is possible to stop the reactions after
complete cellulose hydrolysis but before glucose degradation [67]. It
It is noted that in this reaction zone the hydrolysis of biomass has was also proven that cellulose hydrolysis reactions were highly influ-
gained a lot of attention [67,68]. In fact, special attention has been paid enced by temperature, meanwhile pressure did not affected cellulose
to the hydrolysis of cellulose, since it is the major component of lig- hydrolysis rate in the studied range [75,76].
nocellulosic biomass and therefore is the key to better understand the
reaction mechanisms, kinetics and performance of supercritical water 4.2.2. Cellulose hydrolysis kinetics in supercritical water
hydrolysis of real biomass [71,72]. Cellulose was hydrolyzed following the main hydrolysis reaction
pathway in supercritical water which is shown in Fig. 9 [71], where it
can be seen that cellulose is firstly hydrolyzed into oligosaccharides and
4.2.1. Production of sugars from cellulose hydrolysis in supercritical water then into glucose. Once the glucose has been produced, it can be iso-
The conversion of cellulose to sugars in supercritical water has been merized to fructose and then converted into dehydrated (5-HMF) or
extensively studied using different kinds of reactors. The hydrolysis in retro-aldol condensation products (glycolaldehyde, pyruvaldehyde
batch-type reactors is usually carried out with long reaction times, fa- and/or glyceraldehyde). As mentioned above, working at 400 °C and
voring the decomposition of glucose to degradation products [73,74]. very short reaction times, the reaction would be stopped at glucose.
However, the flow-type system makes it possible to reduce the reaction However, if the reaction time is increased, retro-aldol condensation
time and therefore increasing the yields of sugars instead of degrada- products would be produced, yielding aldehydes as glycolaldehyde,
tion products [61,70]. Recently our research group developed an ex- pyruvaldehyde and/or glyceraldehyde. Therefore, the control of reac-
perimental set up to perform the hydrolysis of cellulose suspensions in tion time was the key factor to selectively hydrolyze cellulose in su-
supercritical water by using a continuous micro-reactor, giving as a percritical water.
result a total conversion of cellulose in milliseconds and yielding a The properties of water may vary considerably when changing the
sugars production of 96% w/w [67]. This continuous micro-reactor is conditions form subcritical to supercritical, affecting to the products
shown in Fig. S2, where it can be seen that the reaction section con- yielded from cellulose hydrolysis [72]. Just by changing pressure and
sisted of a tee junction (M) where the cellulose (or biomass) was in- temperature, different reaction mechanisms are favored. Water at
stantaneously heated up by mixing it with a supercritical water stream. 25 MPa and temperatures below 300 °C has a density around 800 kg/
In order to effectively stop the hydrolysis reaction, a sudden de- m3 and an ionic product (as pK) between 11 and 14. Under those
pressurization through a needle valve was carried out, so that the ef- conditions, water is highly dissociated and H+/OH− ions are highly
fluent was immediately cooled down from 400 °C to around 100 °C and available in the reaction medium and therefore ionic reactions are fa-
therefore reaction was over. Then, depending on the dimensions of the vored [77,78]. However, when temperature is increased up to 400 °C at
pipe between the junction and the depressurization valve, the reaction constant pressure, the density considerably decreases (being around
time was calculated as a function of reactor volume and flow to the 150 kg/m3) and the ionic product increases up to 21 [79]. This change
reactor, so that just by changing the dimensions of the reactor of the in the ionic product affects the kinetics of glucose and fructose de-
pumped flow, different reaction times would be provided. In terms of gradation, avoiding the ionic degradation reactions (which are the
sugars yield from cellulose hydrolysis in hydrothermal medium, several governing chemistry when using acid catalysts) and favoring the radical
conditions were tested by changing temperature, pressure and reaction reactions [72]. In fact, it was found that the concentration of H+/OH−
time in the micro-reactor mentioned above. As a result, it was found due to water dissociation was a determining factor in the selectivity of
that the optimal conditions to obtain soluble sugars (up to six units of cellulose hydrolysis in supercritical water [76]. So far, kinetic models
glucose) were achieved at 400 °C with extremely short reaction times for cellulose hydrolysis only considered the concentration of cellulose
(around 0.01 s). If the reaction time was increased, the sugars were and its derived products into the equations, so that first order kinetics
hydrolyzed and the yield decreased, as it can be seen in Fig. 8. The were selected to predict cellulose hydrolysis in supercritical water.
combination of supercritical water medium and the effective method Following those traditional kinetic models Cantero et al. [76] found an
for the reaction time control allowed such a high sugars yield from incongruity for the kinetic constants of fructose dehydration to 5-HMF
cellulose hydrolysis. This fact can be explained taking into account than when carrying out the hydrolysis of cellulose in supercritical water at
temperatures between 300 and 400 °C and 25 MPa. In Fig. 10 it can be
seen the fitted kinetic constants of 5-HMF formation (khmf) versus the
reciprocal temperature, according to Arrhenius law. A break point can
be clearly observed in Fig. 10a, corresponding to the surroundings of
the critical point of water, which represents a deviation from Arrhenius
law. So that, the traditional models where only cellulose concentration
was taken into account in a first order kinetics equation were only
capable to predict the kinetic constants of fructose dehydration to 5-
HMF at subcritical conditions. That suggested that another factor was
not taken into account into the kinetic equation. To solve the problem,
the concentration of protons and hydroxide ions were added to the
kinetic model, turning it into a second order kinetic equation. As a
consequence of that transformation, the kinetic constants followed the
Arrhenius law for the full temperature spectra, meaning that the de-
hydration to 5-HMF under both subcritical and supercritical conditions
was lineally fitted as it can be observed in Fig. 10b. That would suggest
Fig. 8. Sugars yield from cellulose hydrolysis in hydrothermal medium along reaction that the selectivity of the process was strongly affected by the protons
time. Experiment temperature: red = 400 °C; yellow = 350 °C; blue = 300 °C. and hydroxide ions concentration in the reaction medium, so that im-
Experiment pressure (♦) 27 MPa; (■) 25/23 MPa and (▲) 23/18 MPa [76]. (For inter- proving the understanding of the reaction mechanisms of the hydrolysis
pretation of the references to colour in this figure legend, the reader is referred to the web of cellulose in supercritical water. In that way, retroaldol condensation
version of this article.)
reactions from glucose and fructose (to produce aldehydes) are not very

557
M.J. Cocero et al. The Journal of Supercritical Fluids 133 (2018) 550–565

Fig. 9. Reaction pathway for cellulose hydrolysis in supercritical water based on [71].

demanding of ions and therefore they are favored when water is highly supercritical water. As it can be seen in Fig. 11, experimental results of
associated (as it occurs at supercritical state). On the other hand, iso- cellulose hydrolysis in supercritical water at 400 °C and 25 MPa and
merization glucose-fructose and dehydration reactions are not favored different concentrations were fitted to the first order kinetic equation
since these reactions take place forming transition states with OH− and by plotting the logarithm against the reaction time. In all cases, a linear
H+ and thus they are diminished when water is highly dissociated [76]. dependence was found, where the slope represented the kinetic con-
stant, k. In Fig. 11 it can be observed that when increasing the cellulose
inlet concentration the reaction rate is slower, suggesting that mass
4.2.3. Cellulose concentrations as a mass transfer limitation transfer resistances must have an important effect over cellulose hy-
Another factor recently revised concerning cellulose hydrolysis and drolysis kinetics. Also, combining those data with the ones from a
dissolution in supercritical water was the effect of cellulose con- previous work [67] it was possible to calculate the so called mass
centration itself [71]. So far, existing models describing the conversion transfer limit for cellulose hydrolysis in hydrothermal media. Those
rate of cellulose assumed that the hydrolysis of cellulose particles takes calculations are detailed in another work from our research group [71].
place at their surface and therefore the particle size was considered the It was mathematically possible to distinguish between homogeneous
key parameter for the conversion rate. That shrinking-core model im- and heterogeneous reaction medium, just by calculating a new kinetic
plied the use of a nonconventional kinetic equation [66,72]. On the constant (−20.94 s−1), which corresponded to an inlet concentration
other hand, to take into account the reagent concentration, a first order of 3.83% w/w (identified as mass transfer limit). When the concentration
kinetic was assumed to describe the conversion rate of cellulose in

Fig. 10. Kinetic constants Arrhenius fitting for fructose dehydration to 5-HMF at 25 MPa and temperatures between 300 and 400 °C [76]. a) Kinetic evaluation just considering cellulose
and derived products concentration. b) Kinetic evaluation also considering protons and hydroxide ions concentrations as reagents.

558
M.J. Cocero et al. The Journal of Supercritical Fluids 133 (2018) 550–565

formed by all those polymers in plant cell wall makes cellulose and
hemicellulose less accessible for hydrolysis and therefore, higher reac-
tion time are needed in order to hydrolyze them into sugars. That would
explain the need to move optimal conditions from 0.015 s for pure
cellulose hydrolysis to 0.2 s for biomass hydrolysis in supercritical
water. That increase on reaction time promotes the appearance of other
compounds such as glycolaldehyde. Therefore, through supercritical
water hydrolysis of cellulose and biomass it was possible to obtain high
yields of sugars and building blocks for further production of added
value compounds, just by changing the reaction time.

5. Lignin

Lignin, after cellulose, is the second most abundant raw material of


organics [82]. However, lignin is still under use compared to other
biomass products due to its difficult decomposition and the high
Fig. 11. Kinetic analysis for cellulose concentrations of 5, 15 and 20% w/w (corre- amounts of solid residue obtained during its depolymerization [83].
sponding to 1.5, 4.5 and 6% w/w at the reactor inlet). The regression coefficients were: Traditionally, lignin has been considered as a low value by-product
0.90, 0.81 and 0.96, respectively [71].
of the pulping industry. Only 2% of lignin isolated from spent pulping
liquor is used for specialties [84]. In spite of its amorphous and highly
was lower than the mass transfer limit, cellulose was completely solu- branched structure, it is widely accepted that lignin structure comes
bilized in supercritical water and it can be considered that the hydro- from the polymerization of three phenylpropane monomer units,
lysis occurred in a homogeneous phase and thus the conversion rate namely coniferyl, synapyl, and p-coumaryl alcohol [85]. These mono-
was higher. On the contrary, if the concentration was higher than lignols produce guaiacyl, syringyl and p-hydroxyphenyl propanoic units
3.83%, the cellulose behaved as if it was hydrolyzed at subcritical into the lignin polymer. These substituted phenols yield a huge number
conditions. For this subcritical hydrolysis-like, the cellulose was not of functional groups and linkages, which vary from species to species,
totally dissolved and hydrolysis reaction occurred in a heterogeneous tree to tree, and even in woods from different parts of the same tree
phase. [83]. Lignin structure is also influenced by environmental and devel-
opmental cues [86]. For instance, hardwood lignins are primarily
4.2.4. From cellulose hydrolysis to real biomass hydrolysis in supercritical composed by guaiacyl and syringyl units with traces of p-hydro-
water xyphenyl propanoid units, whereas softwood lignins are composed
Then, once all the parameters affecting cellulose hydrolysis/dis- mainly of guaiacyl units with low levels of p-hydroxyphenyl propanoid
solution were revised, it is worth mentioning that the key parameters to units. The theoretical structure of hardwood and softwood lignins, is
selectively hydrolyze cellulose and therefore biomass in supercritical shown in Fig. 12.
water are the effective control of reaction time and the medium prop- Despite its non-well known structure, it suggests that lignin can be a
erties. As the hydrolysis rate of cellulose is higher than the glucose valuable source of chemicals if would be broken into smaller molecular
under supercritical conditions [68], effectively stopping the reaction units. Depolymerization of lignin is an alluring route to an important
after cellulose hydrolysis but before glucose degradation is essential to functionality class. Unfortunately, nowadays this route is deceptive.
obtain high sugar yields. Increasing the reaction time would only pro- Despite a large volume of research, there are very few reports of effi-
duce the degradation of the glucose generated. In order to avoid glucose cient ways of recovering such as high value-added products. Thus, if it
degradation when working above the critical point of water, reaction would be possible to carry out the lignin depolymerization with high
times below 1 s should be selected [67]. It was also found that both yields, lignin would increase its potential as valuable chemicals source,
cellulose [71] and H+/OH− ions [76] concentrations should be taken making more competent the lignocellulosic biorefinery with the effi-
into account in kinetic equations in order to better explain the perfor- cient use of the main three components of biomass, and not only cel-
mance of cellulose hydrolysis in supercritical water. The studies of lulose and hemicellulose as until now. Recently, the hydrolysis of lignin
hydrolysis of cellulose in supercritical water demonstrated that this and its model compounds (vanillic acid, guaiacol, syringol, coniferyl
technology is very promising to obtain mono and oligo-saccharides and synapyl alcohol) in sub and supercritical water is being considered
from cellulose. Working with very short reaction times (lower than 1 s) as a probable pathway in lignin depolymerization. As is well known, the
it was possible to obtain high yields of sugars and low degradation hydrolysis process has some advantages compared to other methods, it
products content. In fact, this technology already proved to be an ef- is performed at lower temperatures, the employed reactants are cheap
fective method not only for pure cellulose hydrolysis but also for and favors higher yields of liquid including monomeric phenols [88].
complex biomass hydrolysis, such as wheat bran [80] and sugar beet
pulp [81]. Working with the continuous sudden expansion micro-re- 5.1. Model compounds
actor mentioned above, it was possible to obtain both sugars and
building blocks (as glycolaldehyde [43,44]) just by changing the reac- The hydrothermolysis of vanillic acid (VA) was studied using a
tion time. When working with wheat bran, the highest recovery of tubular flow reactor, with residence times from 5 to 70 s, between 300
cellulose and hemicellulose as soluble sugars was 73% w/w operating and 375 °C. It was found that below the critical temperature of water,
at 400 °C, 25 MPa and 0.19 s of reaction time. On the other hand, VA was converted exclusively to 2-methoxy-phenol (guaiacol) through
starting with sugar beet pulp as raw material, working at similar con- decarboxylation. At 350 °C the conversion was achieved faster (after
ditions (400 °C, 25 MPa and 0.2 s), a significant amount of glycolalde- only 15 s) than at 300 °C (60 s). At temperatures of 375 °C the con-
hyde (more than 10% w/w) was produced apart from sugars. That ef- version was even more rapid, but the selectivity towards guaiacol was
fluent after supercritical water hydrolysis containing sugars and lower, with catechol being formed as the main secondary product.
glycolaldehyde was then hydrogenated over Ru/MCM-48 catalyst ob- Phenol was also formed, through the free radical decomposition of
taining hexitols and ethylene glycol as products. It is clear that working guaiacol. As the temperature was increased further, more of the sec-
with a real biomass implies not only dealing with cellulose, but also ondary products were formed [89].
with hemicellulose, pectins, lignin, proteins, etc. The intricate matrix Guaiacol decomposition in supercritical water was investigated in

559
M.J. Cocero et al. The Journal of Supercritical Fluids 133 (2018) 550–565

Fig. 12. Typical structure of lignin derived from hardwood (left) and softwood (right) [87].

sealed reactors using both water and in water-salt solutions (NaCl, times were 10–64 min. The reaction products were separated into THF
CaCl2 and FeCl3). The reaction temperature was 383 °C with reaction soluble (TS) and THF insoluble (TIS) fractions. In the absence of phenol
times between 0 and 30 min. It was concluded that at low water den- it was found that the molecular weight distribution of TS products
sity, the reactions led to the formation of phenol, catechol, cresol and shifted to lower molecular weight as the water density increased. The
char, but at higher water densities and with addition of salts the rate of TS products (syringols, guaiacols and catechol) were derived from
hydrolysis and the selectivity towards hydrolysis products, catechol and lignin structure. Authors explained that the conversion of lignin in su-
methanol, was increased [90]. On the other hand, it was also in- percritical water probably proceeded through hydrolysis and deal-
vestigated batch reactions of guaiacol using a 5 ml reactor with re- kylation, which leads to the formation of lighter fraction, such as al-
sidence times between 5 and 180 min (including 3 min of heating up), cohols, aldehydes and their functional groups within the
at 380, 390 and 400 °C. The main products were catechol, phenol and o- macromolecules. The functional groups could form not only as de-
cresol. Catechol was formed quickly over the first 10 min of the reac- composed fragments but also as lignin itself. In the presence of phenol,
tion. Phenol and o-cresol were both formed gradually over the entire there was a lower yield of TIS products, which were also of a lower
reaction time [91]. mass. The yield of TS increased and their average mass also decreased.
Formation of formic and acetic acid from syringol decomposition This can be rationalized by enhanced hydrolysis of lignin at high water
was studied under oxidizing conditions in a batch reactor, with and density, which produces reactive fragments such as formaldehyde. In
without NaOH catalyst. Reaction times of 30–150 s and temperatures the absence of phenol, these reactive fragments act as cross linkers
between 250 and 300 °C were investigated. NaOH had been shown to between depolymerization products such as guaiacol, catechol, and
inhibit the decomposition of organic compounds at these temperatures. larger fragments of lignin, repolymerising the lignin molecule. It is
The phenolic compounds formed were catechol, 1, 2,4-benzenetriol, 1,4 believed that phenol acts as a capping agent by reacting with these
benzenediol and 9 short chain carboxylic acids, ranging from 1 to 6 species to prevent the repolymerization and the formation of the TIS
carbons. The optimal temperature for formic and acetic acid production molecules [95].
was 280 °C though this led to a lower formation of other products [92]. Okuda el al. studied the depolymerization of organosolv lignin in
The formation of organic acids from the decomposition of coniferyl phenol-water mixtures at 400 °C for 6–60 min. As in previous work, the
and synapyl alcohols using a batch reactor at 380 °C, 1000 bar and information about lignin origin is not given. In this study they were
4 min was investigated. Coniferyl and synapyl alcohols formed formic, focused on the production of phenolic chemicals from lignin and ex-
acetic, glycolic and lactic acid. It was suggested that these were formed amined the depolymerization of lignin in a water-phenol mixture with
from decomposition of the aliphatic side chains, being the aromatic higher phenol ratio in order to assess the possibility of complete con-
rings resistant to ring opening reactions under these reaction conditions version of lignin to phenolic chemicals without the formation of char. It
[93]. was found that the yield of TS compounds decreased with time, and
after 1 h the lowest yield was given when just water was used as sol-
vent. The best performing solvent was a water-phenol mixture, which
5.2. Lignin depolymerization
achieved nearly total suppression of char formation with 99% TS mo-
lecules [96].
5.2.1. Reactions in water and co-solvent
Fang et al. followed decomposition of organosolv lignin in water/
Existence of the additional hydrolysis reaction in water at elevated
phenol solution in micro-reactor coupled with optical microscopies at
temperatures and pressures catalyzed by H+ and OH− should cause
temperatures up to 600° C and water densities up to 1165 kg/m3. The
significantly different decomposition from pyrolysis, and the associated
microreactor, diamond anvil cell (DAC) allows for in-situ observations
phase behavior. As a weak-polar solvent with a high value of ion pro-
of samples in the fully-visible chamber via optical microscopy. The DAC
duct supercritical water could be a possible solvent that can dissolve
consisted of a hole and sealed by compression of two opposing anvils
and hydrolyze lignin for potentially production of phenolic chemicals
made of diamond. The chamber was rapidly heated by two electric
or for upgrading lignin for fuels [94].
microheaters by cutting power, which is convenient for the study of
Saisu et al. reported the decomposition of organosolv lignin in a
phase behavior and chemical reactions. Experiments have been done at
batch reactor in supercritical water with and without phenol at 400 °C.
different water densities, heating rate, maximum temperatures and
As properties of lignin depend on the lignin origin, not just on the
lignin concentration. Three different types of products were obtained: a
isolation way, it is important to emphasize that in this work authors did
non-dissolved black residue, a precipitated residue and reddish oil. A
not give additional information about the lignin origin. Both the lignin
homogenous phase was formed for the phenol + lignin system where
to phenol ratios and the volume of water in the reactor were varied in
phenolic char precipitated as the main product. Adding water to this
order to investigate the effect of phenol and water density. Residence

560
M.J. Cocero et al. The Journal of Supercritical Fluids 133 (2018) 550–565

system de-polymerization of lignin was promoted by hydrolysis in a ether bonds from abundant β-aryl ether (β-O-4) linkages in softwood
homogeneous phase and its re-polymerization was inhibited by phenol. lignin [101,102]. The low dissociation enthalpies of the ether bond in
The homogenous phase was not found in the case of lignin + water the β-O-4 linkage initiated the reaction to form a phenoxy radical and a
system. After initial dissolution at above 377 °C lignin underwent hy- secondary alkyl aromatic radical [103]. The Arrhenius behavior shown
drolysis and pyrolysis to phenolic, which are further changed to oil in by lignin decomposition under hydrothermal conditions even in sub-
the aqueous phase. At higher temperatures, solid particles precipitated critical region further supported the conclusion that the initial de-
from the aqueous via homogeneous re-polymerization of the phenolics composition was a radical reaction. TOC yield decreased with tem-
and water soluble compounds to form a phenolic char. At these same perature and the yield was much higher under subcritical condition.
conditions, non-dissolved lignin underwent heterogeneous pyrolysis The TOC yield in subcritical water increased within short residence
and formed polyaromatic char. Higher water density decrease lignin time and remained stable or decrease slowly despite longer residence
dissolution. Therefore, polyaromatic char, with a lighter molecular time. The increasing in the polymerization during the increase of
weight was the main product along with a smaller fraction of phenolic temperature should be reflected in the TOC yield. Low TOC yield under
char. It can be conclude that for water and phenol mixtures, lignin can supercritical condition suggested the occurrence of secondary reaction.
be completely solubilized and undergoes homogeneous hydrolysis and This could be due to the cross-linking between reactive degradation
pyrolysis that prevents further re-polymerization [94]. fragments obtained from the lignin depolymerization to produce frag-
ments with higher molecular weights. Increase in the lower molecular
5.2.2. Water without co-solvent weight compounds during the time resulted with the simultaneous
Sasaki and Goto presented a work in which the chemical conversion formation of the higher molecular weight compounds because of the
of alkali lignin in near and supercritical water at 350 °C and 400 °C and repolymerization. The minimal decrease in TOC yield for subcritical
a pressure of 25–40 MPa using a batch reactor without catalyst, having temperatures implied that the crosslinking reactions between these
5–240 min residence time was studied. The products were separated lower-molecular weight compounds did not take place actively under
into two fractions, methanol soluble (MS) and methanol insoluble (MI). these conditions. This suggested the significance of radical’s involve-
The main products observed in the MS fraction were catechol, phenol, ment in enhancing the reaction [98,99].
and o, m, p- cresols, while MI product was defined as a residual solid. It Char has significantly higher yields in the supercritical region and
was proposed that catechol is formed via hydrolysis of guaiacol which is formation was enhanced at elevated temperatures. Formation of char
the main compound in structure of lignin. In further hydrolysis, phenol, from lignin follows Arrhenius behavior and it is not affected with
m, p and o-cresol were obtained. Dependence of reaction time showed change in water properties under subcritical condition what point ra-
that the yield of catechol rapidly increased with reaction time (till dical reaction. In order to examine the suggested hypothesis of forma-
30 min) and then decreased, especially at 400 ° C, while the yields of tion of low molecular-weight fragments and formation of higher mo-
phenol, m, p and o-cresol increased with reaction time. After 90 min the lecular weight fragments by cross linking of the smaller fragment it was
yields of m, p and o-cresols were almost constant while the yield of determine the yields of the phenolic compounds and aromatic hydro-
phenol slightly increased. At 400 °C after catechol was consumed, the carbons. The main phenolic compound from lignin decomposition is
majority of the reaction most likely terminated. The decreasing of ca- guaiacol, which is followed by minor composition of other phenolic
techol was not followed by the increasing of phenol, m, p and o-cresol compounds such as o, m, p-cresol, catechol and phenol. Formation of
significantly. Water density influence yields of products where the yield guaiacol was higher in supercritical temperature, but rapidly decreased
of catechol was gradually decreased with increasing the water density at longer residence time [98,99]. Guaiacol is an intermediate de-
at 350 °C and dramatically decreased at 400 °C. The yields of phenol, m, gradation product and highly reactive, since the methyl CeO bond is
p and o-cresol increased gradually with increasing the water density at the weakest in the guaiacol unit and is susceptible to undergo cleaving.
350 °C and 400 °C. It was suggested that an increase in water density The aliphatic CeO bond of the methoxyl group is more likely to react
enhanced the hydrolysis rate of ether and carbon-carbon bonds of al- because the bond energy of the aliphatic CeO bond (245 kJ/mol) is
kylphenol in lignin. According to this results it was proposed reaction smaller than that of the aromatic CeO bond (256 kJ/mol). This was
mechanism showed in scheme below (Fig. 13) where lignin was de- concluded in the study of Wahyudiono et al. where also was found that
graded into its derivate compounds by dealkylation and hydrolysis re- guaiacol showed a fast decomposition rate and the formation of high-
action. Under SCW conditions hydrolysis takes place at ether and ester molecular-weight substances reformed to char was important for the
bonds in lignin. Hydrolysis is accelerated by a high ion product of guaiacol decomposition to reach equilibrium [104]. However, high
water. Dealkylation of lignin gives catechol, which is then hydrolyzed yield of guaiacol was also obtained under subcritical conditions. The
into phenol. This reaction pathway suggests that some useful chemical high yield of guaiacol under two separate regions of temperature
intermediates (MS fraction) might be recovered in a rapid and selective (subcritical and supercritical) with different water properties indicated
manner by changing the temperature, reaction time under near and guaiacol formation via two different pathways. The formation of
supercritical water condition. At the same time, re-polymerization of guaiacol from lignin probably proceeded through hydrolysis under
low molecular weight compounds occurs as seen by the formation of subcritical conditions because of the high ionic product and dielectric
char through condensation reaction [97]. constant of water. On the contrary, under supercritical condition and
Lignin conversion was also investigated in the continuous system for high temperature free radical reaction should be enhanced that lead to
short residence time 0.5–10 s, pressure of 25 MPa and different tem- the formation of guaiacol from lignin [98]. In both studies it was
peratures under supercritical conditions at 390 °C and 450 °C and showed that the decomposition of lignin occurred rapidly with re-
subcritical condition at 300 °C and 370 °C [98,99]. Temperature plays sidence time below one second, which indicate that kinetic study
an important factor in deciding the dominant pathway because of the should be done for residence time below 1 s.
existence of the parallel ionic and radical based pathways. Lignin pro-
ducts were divided in char, gas, TOC, phenolic and aromatic hydro- 5.2.3. Water and base catalyst
carbons. Under hydrothermal condition lignin was rapidly converted In order to enhance the obtaining of monomeric phenols, basic
into lower molecular weight products for all temperatures which was compounds such as hydroxides are used as catalyst [105,106] [107].
followed with high yields of TOC, phenolic compounds and aromatic Studies on lignin model compound dihydro-diisoeugenol, showed that
hydrocarbons, while decomposition was accelerated under supercritical the basic agent caused ether and CeC bond cleavage which yielded
condition [99]. Increasing decomposition rate with temperature follows volatile phenols [108]. Furthermore, the analysis of products from
Arrhenius behavior of lignin degradation what was already obtained by model compound reactions revealed that phenyl ether linkages were
Zhang et al. [100]. The rapid depolymerization is cause by cleavage of effectively broken in the base catalyzed hydrolysis reaction while CeC

561
M.J. Cocero et al. The Journal of Supercritical Fluids 133 (2018) 550–565

Fig. 13. Proposed scheme for degradation of lignin


under near and supercritical condition.

linkages were less affected [109]. In another study, it was concluded together with a larger amount of less expensive base (Ca(OH)2) pro-
that in alkaline depolymerisation of lignin, ether bonds are hydrolyzed duced positive results [105].
at random, most likely from the outside of the oligomer and not in the Silva et al. studied the catalytic depolymerization of organosolv
sequence of their bond strengths, forming first large units and then lignin with both NaOH catalyst and with boric acid as a capping agent,
smaller subunits [106]. In addition, it was stated that the formation of aiming to produce oils of monomeric and dimeric products. In the case
monomers is directly proportional to the concentration of sodium hy- of reactions with NaOH and no capping agent, the highest oil yield was
droxide in the aqueous medium. Furthermore, a mechanism for the obtained at 300 °C with a residence time of 4 min. This gave a yield of
NaOH catalyzed breakdown of the ether bonds of lignin is proposed 23% oil and no char formation. Lignin conversion increased steadily
explaining the preferential formation of syringol derivatives, based on with increasing temperature but char was formed as well as oil. In order
the stabilizing effect that the methoxyl groups provides to the transition to increase oil yields, boric acid was used as a capping agent. Without
states of the carbenium ions. It was also concluded that the production base, the boric acid increased the yield of oils to a maximum of 36%
of monomers is limited by the oligomerization and polymerization re- after 40 min at 300 °C, but at longer residence times or higher tem-
actions of the products formed. peratures the yield decreased again. The results showed that the mo-
Miller et al. showed that in the alkali depolymerization of lignin lecular weights of the oils from the boric acid catalyzed reactions were
using water as solvent the most important factor in lignin depolymer- around 500 Da, compared to 300 Da for the base catalyzed depoly-
ization was base concentration. Moreover, it was observed that con- merization [88].
centration excess of a strong base gave better results on lignin depo- In contrast to a basic environment, leading to deprotonation of
lymerization. In addition, a little amount of a strong base (NaOH) phenolic hydroxyl groups and decreased hydrogen bonding, the acidic

562
M.J. Cocero et al. The Journal of Supercritical Fluids 133 (2018) 550–565

environment enhances the degree of internal hydrogen bonding. As [11] B. Sukhbaatar, E.B. Hassan, M. Kim, P. Steele, L. Ingram, Optimization of hot-
compressed water pretreatment of bagasse and characterization of extracted
result, the probability of acid-catalyzed cleavage of ether bonds is re- hemicelluloses, Carbohydr. Polym. 101 (2014) 196–202, http://dx.doi.org/10.
duced compared to base-catalyzed cleavage. Thus, in acid-catalyzed 1016/j.carbpol.2013.09.027.
hydrolysis the primary products produced are larger (dimers to tetra- [12] J.V. Rissanen, D.Y. Murzin, T. Salmi, H. Grénman, Aqueous extraction of hemi-
celluloses from spruce – from hot to warm, Bioresour. Technol. 199 (2016)
mers) than in the base-catalyzed route. For both cases, the primary 279–282, http://dx.doi.org/10.1016/j.biortech.2015.08.116.
products undergo easy addition and condensation reactions leading to [13] W. Reynolds, H. Singer, S. Schug, I. Smirnova, Hydrothermal flow-through treat-
higher molecular weight products [110]. ment of wheat-straw: detailed characterization of fixed-bed properties and axial
dispersion, Chem. Eng. J. 281 (2015) 696–703, http://dx.doi.org/10.1016/j.cej.
Under supercritical and subcritical condition lignin is hydrolyzed 2015.06.117.
and different phenolic and other aromatic compounds could be ob- [14] A. Cabeza, F. Sobrón, F.M. Yedro, J. García-Serna, Two-phase modelling and si-
tained. These hydrolysis reactions occur in the shortest time than the mulation of the hydrothermal fractionation of holm oak in a packed bed reactor
with hot pressurized water, Chem. Eng. Sci. 138 (2015) 59–70, http://dx.doi.org/
residence time that has already been used in literature (more than one
10.1016/j.ces.2015.07.024.
second). It is very important to have better understanding of reaction [15] X. Chen, M. Lawoko, A. van Heiningen, Kinetics and mechanism of autohydrolysis
pathways, intermediate reaction products and reaction products for of hardwoods, Bioresour. Technol. 101 (2010) 7812–7819, http://dx.doi.org/10.
first milliseconds of reaction time, thus the specific weaknesses and 1016/j.biortech.2010.05.006.
[16] X.J. Ma, X.F. Yang, X. Zheng, L. Lin, L.H. Chen, L.L. Huang, S.L. Cao, Degradation
strengths of the polymer and its intermediates – i.e. the substructures and dissolution of hemicelluloses during bamboo hydrothermal pretreatment,
which are the most susceptible to chemical attack. Kinetic models that Bioresour. Technol. 161 (2014) 215–220, http://dx.doi.org/10.1016/j.biortech.
have been obtained until today are justified with the final reaction 2014.03.044.
[17] M.E. Vallejos, F.E. Felissia, J. Kruyeniski, M.C. Area, Kinetic study of the extraction
products, without information about intermediate produced. of hemicellulosic carbohydrates from sugarcane bagasse by hot water treatment,
Considerable effort is still required to address the separation chal- Ind. Crops Prod. 67 (2015) 1–6, http://dx.doi.org/10.1016/j.indcrop.2014.12.
lenges associated with lignin depolymerization. The supercritical water 058.
[18] M.H. Thomsen, A. Thygesen, A.B. Thomsen, Hydrothermal treatment of wheat
ultrafast hydrolysis could open a new way to improve the under- straw at pilot plant scale using a three-step reactor system aiming at high hemi-
standing of lignin depolymerization, as has been done in the cellulose cellulose recovery, high cellulose digestibility and low lignin hydrolysis,
hydrolysis. Bioresour. Technol. 99 (2008) 4221–4228 http://linkinghub.elsevier.com/
retrieve/pii/S0960852407007158.
[19] M.S.R. dos Santos Rocha, B. Pratto, R. de Sousa Júnior, R.M.R.G. Almeida,
Acknowledgements A.J.G. da Cruz, A kinetic model for hydrothermal pretreatment of sugarcane straw,
Bioresour. Technol. 228 (2017) 176–185, http://dx.doi.org/10.1016/j.biortech.
2016.12.087.
The authors thank Ministerio de Economia y Competividad, Junta
[20] S. Makishima, M. Mizuno, N. Sato, K. Shinji, M. Suzuki, K. Nozaki, F. Takahashi,
de Castilla y León, and Feder program for the financial support Projects T. Kanda, Y. Amano, Development of continuous flow type hydrothermal reactor
CTQ2013-44143-R,CTQ2016-79777-R, and VA040U16. for hemicellulose fraction recovery from corncob, Bioresour. Technol. 100 (2009)
2842–2848 http://linkinghub.elsevier.com/retrieve/pii/S0960852408010791.
[21] Eseyin E. Anthonia, H. Steele Philip, An overview of the applications of furfural
Appendix A. Supplementary data and its derivatives, Int. J. Adv. Chem. 3 (2015) 42–47, http://dx.doi.org/10.
14419/ijac.v3i2.5048.
Supplementary data associated with this article can be found, in the [22] F.A. Castillo Martinez, E.M. Balciunas, J.M. Salgado, J.M. Domínguez González,
A. Converti, R.P. de S. Oliveira, Lactic acid properties, applications and produc-
online version, at http://dx.doi.org/10.1016/j.supflu.2017.08.012. tion: a review, Trends Food Sci. Technol. 30 (2013) 70–83, http://dx.doi.org/10.
1016/j.tifs.2012.11.007.
References [23] P. Gao, G. Li, F. Yang, X.-N. Lv, H. Fan, L. Meng, X.-Q. Yu, Preparation of lactic
acid, formic acid and acetic acid from cotton cellulose by the alkaline pre-treat-
ment and hydrothermal degradation, Ind. Crops Prod. 48 (2013) 61–67, http://dx.
[1] P. Bajpai, Structure of lignocellulosic biomass, Pretreat. Lignocellul. Biomass doi.org/10.1016/j.indcrop.2013.04.002.
Biofuel. Prod. Springer, Singapore, 2016, pp. 7–12, http://dx.doi.org/10.1007/ [24] J.C. Parajó, G. Garrote, J.M. Cruz, H. Dominguez, Production of xylooligo-
978-981-10-0687-6_2. saccharides by autohydrolysis of lignocellulosic materials, Trends Food Sci.
[2] O. Bobleter, Hydrothermal degradation of polymers derived from plants, Prog. Technol. 15 (2004) 115–120 http://linkinghub.elsevier.com/retrieve/pii/
Polym. Sci. 19 (1994) 797–841 http://cat.inist.fr/?aModele=afficheN%7B& S0924224403002553.
%7Dcpsidt=4247470. [25] G. Garrote, H. Domı́nguez, J.C. Parajó, Interpretation of deacetylation and hemi-
[3] J.V. Rissanen, H. Grénman, C. Xu, S. Willför, D.Y. Murzin, T. Salmi, Obtaining cellulose hydrolysis during hydrothermal treatments on the basis of the severity
spruce hemicelluloses of desired molar mass by using pressurized hot water ex- factor, Process Biochem. 37 (2002) 1067–1073, http://dx.doi.org/10.1016/
traction, ChemSusChem 7 (2014) 2947–2953, http://dx.doi.org/10.1002/cssc. S0032-9592(01)00315-6.
201402282. [26] M.R. Narkhede, C. Nguyen-Trung, D.A. Palmer, Dissociation quotients of D-ga-
[4] G. Gallina, Á. Cabeza, P. Biasi, J. García-Serna, Optimal conditions for hemi- lacturonic acid in aqueous solution at 0.1 MPa to 1 molal ionic strength and
celluloses extraction from Eucalyptus globulus wood: hydrothermal treatment in a 100 °C, J. Solut. Chem. 23 (1994) 877–888, http://dx.doi.org/10.1007/
semi-continuous reactor, Fuel Process. Technol. 148 (2016) 350–360, http://dx. bf00972751.
doi.org/10.1016/j.fuproc.2016.03.018. [27] M.S. Tunc, A.R.P. van Heiningen, Hemicellulose extraction of mixed southern
[5] C. Wyman, S. Decker, M. Himmel, J. Brady, C. Skopec, L. Viikari, Hydrolysis of hardwood with water at 150 °C: effect of time, Ind. Eng. Chem. Res. 47 (2008)
cellulose and hemicellulose, Polysaccharides, CRC Press, 2004, http://dx.doi.org/ 7031–7037 http://pubs.acs.org/doi/abs/10.1021/ie8007105.
10.1201/9781420030822.ch43. [28] T. Song, A. Pranovich, B. Holmbom, Effects of pH control with phthalate buffers on
[6] A. Cabeza, C.M. Piqueras, F. Sobrón, J. García-Serna, Modeling of biomass frac- hot-water extraction of hemicelluloses from spruce wood, Bioresour. Technol. 102
tionation in a lab-scale biorefinery: solubilization of hemicellulose and cellulose (2011) 10518–10523, http://dx.doi.org/10.1016/j.biortech.2011.08.093.
from holm oak wood using subcritical water, Bioresour. Technol. 200 (2016) [29] H.A. Ruiz, M.A. Cerqueira, H.D. Silva, R.M. Rodríguez-Jasso, A.A. Vicente,
90–102, http://dx.doi.org/10.1016/j.biortech.2015.09.063. J.A. Teixeira, Biorefinery valorization of autohydrolysis wheat straw hemi-
[7] F. Carvalheiro, L.C. Duarte, F. Gírio, P. Moniz, Chapter 14 – hydrothermal/liquid cellulose to be applied in a polymer-blend film, Carbohydr. Polym. 92 (2013)
hot water pretreatment (autohydrolysis): a multipurpose process for biomass up- 2154–2162, http://dx.doi.org/10.1016/j.carbpol.2012.11.054.
grading A2 – Mussatto, Solange I, Biomass Fractionation Technol. A Lignocellul. [30] G. Brunner, Chapter 8 – processing of biomass with hydrothermal and supercritical
Feed. Based Biorefinery, Elsevier, Amsterdam, 2016, pp. 315–347, http://dx.doi. water, in: B. Gerd (Ed.), Supercrit. Fluid Sci. Technol. Elsevier, 2014, pp. 395–509
org/10.1016/B978-0-12-802323-5. http://www.sciencedirect.com/science/article/pii/B978044459413600008X.
[8] C.M. Piqueras, Á. Cabeza, G. Gallina, D.A. Cantero, J. García-Serna, M.J. Cocero, [31] F.M. Yedro, D.A. Cantero, M. Pascual, J. García-serna, M.J. Cocero, Bioresource
Online integrated fractionation-hydrolysis of lignocellulosic biomass using sub- technology hydrothermal fractionation of woody biomass: lignin effect on sugars
and supercritical water, Chem. Eng. J. 308 (2017) 110–125, http://dx.doi.org/10. recovery, Bioresour. Technol. 191 (2015) 124–132, http://dx.doi.org/10.1016/j.
1016/j.cej.2016.09.007. biortech.2015.05.004.
[9] F.M. Yedro, H. Grénman, J.V. Rissanen, T. Salmi, J. García-Serna, M.J. Cocero, [32] H.A. Ruiz, R.M. Rodriguez-Jasso, B.D. Fernandes, A.A. Vicente, J.A. Teixeira,
Chemical composition and extraction kinetics of Holm oak (Quercus ilex) hemi- Hydrothermal processing, as an alternative for upgrading agriculture residues and
celluloses using subcritical water, J. Supercrit. Fluids (2017), http://dx.doi.org/ marine biomass according to the biorefinery concept: a review, Renew. Sustain.
10.1016/j.supflu.2017.01.016 in press. Energy Rev. 21 (2013) 35–51, http://dx.doi.org/10.1016/j.rser.2012.11.069.
[10] J.V. Rissanen, H. Grénman, S. Willför, D.Y. Murzin, T. Salmi, Spruce hemicellulose [33] H.-Q. Li, W. Jiang, J.-X. Jia, J. Xu, pH pre-corrected liquid hot water pretreatment
for chemicals using aqueous extraction: kinetics, mass transfer, and modeling, Ind. on corn stover with high hemicellulose recovery and low inhibitors formation,
Eng. Chem. Res. 53 (2014) 6341–6350, http://dx.doi.org/10.1021/ie500234t. Bioresour. Technol. 153 (2014) 292–299, http://dx.doi.org/10.1016/j.biortech.

563
M.J. Cocero et al. The Journal of Supercritical Fluids 133 (2018) 550–565

2013.11.089. near- and supercritical water solubilization, J. Agric. Food Chem. 51 (2003)
[34] S. Ou, K.-C. Kwok, Ferulic acid: pharmaceutical functions, preparation and ap- 5376–5381, http://dx.doi.org/10.1021/jf025989i.
plications in foods, J. Sci. Food Agric. 84 (2004) 1261–1269, http://dx.doi.org/10. [62] S. Kumar, R.B. Gupta, Hydrolysis of microcrystalline cellulose in subcritical and
1002/jsfa.1873. supercritical water in a continuous flow reactor, Ind. Eng. Chem. Res. 47 (2008)
[35] B.A. Acosta-Estrada, J.A. Gutierrez-Uribe, S.O. Serna-Saldivar, Bound phenolics in 9321–9329, http://dx.doi.org/10.1021/ie801102j.
foods, a review, Food Chem. 152 (2014) 46–55, http://dx.doi.org/10.1016/j. [63] L.K. Tolonen, G. Zuckerstätter, P.A. Penttilä, W. Milacher, W. Habicht, R. Serimaa,
foodchem.2013.11.093. A. Kruse, H. Sixta, Structural changes in microcrystalline cellulose in subcritical
[36] N. Kumar, V. Pruthi, Potential applications of ferulic acid from natural sources, water treatment, Biomacromolecules 12 (2011) 2544–2551, http://dx.doi.org/10.
Biotechnol. Rep. 4 (2014) 86–93, http://dx.doi.org/10.1016/j.btre.2014.09.002. 1021/bm200351y.
[37] K. Krygier, F. Sosulski, L. Hogge, esterified, and insoluble-bound phenolic acids. 1. [64] M. Mohan, R. Timung, N.N. Deshavath, T. Banerjee, V.V. Goud, V.V. Dasu,
Extraction and purification procedure, J. Agric. Food Chem. 30 (1982) 330–334. Optimization and hydrolysis of cellulose under subcritical water treatment for the
[38] I. Benoit, D. Navarro, N. Marnet, N. Rakotomanomana, L. Lesage-Meessen, J.- production of total reducing sugars, RSC Adv. 5 (2015) 103265–103275, http://
C. Sigoillot, M. Asther, M. Asther, Feruloyl esterases as a tool for the release of dx.doi.org/10.1039/C5RA20319H.
phenolic compounds from agro-industrial by-products, Carbohydr. Res. 341 [65] M.S. Seehra, B.V. Popp, F. Goulay, S.K. Pyapalli, T. Gullion, J. Poston,
(2006) 1820–1827, http://dx.doi.org/10.1016/j.carres.2006.04.020. Hydrothermal treatment of microcrystalline cellulose under mild conditions:
[39] H. Barberousse, A. Kamoun, M. Chaabouni, J.-M. Giet, O. Roiseux, M. Paquot, characterization of solid and liquid-phase products, Cellulose 21 (2014)
C. Deroanne, C. Blecker, Optimization of enzymatic extraction of ferulic acid from 4483–4495, http://dx.doi.org/10.1007/s10570-014-0424-y.
wheat bran, using response surface methodology, and characterization of the re- [66] M. Sasaki, T. Adschiri, K. Arai, Kinetics of cellulose conversion at 25 MPa in sub-
sulting fractions, J. Sci. Food Agric. 89 (2009) 1634–1641 http://onlinelibrary. and supercritical water, AIChE J. 50 (2004) 192–202, http://dx.doi.org/10.1002/
wiley.com/doi/10.1002/jsfa.3630/abstract. aic.10018.
[40] O. Pourali, F.S. Asghari, H. Yoshida, Production of phenolic compounds from rice [67] D.A. Cantero, M. Dolores Bermejo, M. José Cocero, High glucose selectivity in
bran biomass under subcritical water conditions, Chem. Eng. J. 160 (2010) pressurized water hydrolysis of cellulose using ultra-fast reactors, Bioresour.
259–266, http://dx.doi.org/10.1016/j.cej.2010.02.057. Technol. 135 (2013) 697–703, http://dx.doi.org/10.1016/j.biortech.2012.09.
[41] R.J. Moon, A. Martini, J. Nairn, J. Simonsen, J. Youngblood, Cellulose nanoma- 035.
terials review: structure, properties and nanocomposites, Chem. Soc. Rev. 40 [68] M. Sasaki, Z. Fang, Y. Fukushima, T. Adschiri, K. Arai, Dissolution and hydrolysis
(2011) 3941–3994, http://dx.doi.org/10.1039/C0CS00108B. of cellulose in subcritical and supercritical water, Ind. Eng. Chem. Res. 39 (2000)
[42] Suhas, V.K. Gupta, P.J.M. Carrott, R. Singh, M. Chaudhary, S. Kushwaha, 2883–2890, http://dx.doi.org/10.1021/ie990690j.
Cellulose: s review as natural, modified and activated carbon adsorbent, Bioresour. [69] L.K. Tolonen, M. Juvonen, K. Niemelä, A. Mikkelson, M. Tenkanen, H. Sixta,
Technol. 216 (2016) 1066–1076, http://dx.doi.org/10.1016/j.biortech.2016.05. Supercritical water treatment for cello-oligosaccharide production from micro-
106. crystalline cellulose, Carbohydr. Res. 401 (2015) 16–23, http://dx.doi.org/10.
[43] A. Wang, T. Zhang, One-pot conversion of cellulose to ethylene glycol with mul- 1016/j.carres.2014.10.012.
tifunctional tungsten-based catalysts, Acc. Chem. Res. 46 (2013) 1377–1386, [70] K. Ehara, S. Saka, Decomposition behavior of cellulose in supercritical water,
http://dx.doi.org/10.1021/ar3002156. subcritical water, and their combined treatments, J. Wood Sci. 51 (2005) 148–153
[44] L. Zhang, C. Xu, P. Champagne, Overview of recent advances in thermo-chemical http://www.springerlink.com/content/w26363n08327171w/.
conversion of biomass, Energy Convers. Manag. 51 (2010) 969–982 http://www. [71] C.M. Martínez, D.A. Cantero, M.D. Bermejo, M.J. Cocero, Hydrolysis of cellulose in
sciencedirect.com/science/article/pii/S0196890409004889. supercritical water: reagent concentration as a selectivity factor, Cellulose 22
[45] D. Klemm, B. Heublein, H. Fink, A. Bohn, Cellulose: fascinating biopolymer and (2015) 2231–2243, http://dx.doi.org/10.1007/s10570-015-0674-3.
sustainable raw material, Angew. Chem. Int. Ed. 44 (2005) 3358–3393 http:// [72] D.A. Cantero, M.D. Bermejo, M.J. Cocero, Kinetic analysis of cellulose depoly-
onlinelibrary.wiley.com/doi/10.1002/anie.200460587/abstract. merization reactions in near critical water, J. Supercrit. Fluids 75 (2013) 48–57,
[46] C.-H. Zhou, X. Xia, C.-X. Lin, D.-S. Tong, J. Beltramini, Catalytic conversion of http://dx.doi.org/10.1016/j.supflu.2012.12.013.
lignocellulosic biomass to fine chemicals and fuels, Chem. Soc. Rev. 40 (2011) [73] K. Ehara, S. Saka, A comparative study on chemical conversion of cellulose be-
5588–5617, http://dx.doi.org/10.1039/C1CS15124J. tween the batch-type and flow-type systems in supercritical water, Cellulose 9
[47] G. Brunner, Near critical and supercritical water. Part I. Hydrolytic and hydro- (2002) 301–311, http://dx.doi.org/10.1023/A:1021192711007.
thermal processes, J. Supercrit. Fluids 47 (2009) 373–381 http://linkinghub. [74] Y. Zhao, W.-J. Lu, H.-T. Wang, Supercritical hydrolysis of cellulose for oligo-
elsevier.com/retrieve/pii/S0896844608002970. saccharide production in combined technology, Chem. Eng. J. 150 (2009)
[48] B. Lindman, G. Karlström, L. Stigsson, On the mechanism of dissolution of cellu- 411–417 http://www.sciencedirect.com/science/article/pii/
lose, J. Mol. Liq. 156 (2010) 76–81, http://dx.doi.org/10.1016/j.molliq.2010.04. S1385894709000564.
016. [75] D.A. Cantero, Á. Sánchez Tapia, M.D. Bermejo, M.J. Cocero, Pressure and tem-
[49] B. Medronho, B. Lindman, Competing forces during cellulose dissolution: from perature effect on cellulose hydrolysis in pressurized water, Chem. Eng. J. 276
solvents to mechanisms, Curr. Opin. Colloid Interface Sci. 19 (2014) 32–40, (2015) 145–154, http://dx.doi.org/10.1016/j.cej.2015.04.076.
http://dx.doi.org/10.1016/j.cocis.2013.12.001. [76] D.A. Cantero, M.D. Bermejo, M.J. Cocero, Governing chemistry of cellulose hy-
[50] L.K. Tolonen, M. Bergenstråhle-Wohlert, H. Sixta, J. Wohlert, Solubility of cellu- drolysis in supercritical water, ChemSusChem 8 (2015) 1026–1033, http://dx.doi.
lose in supercritical water studied by molecular dynamics simulations, J. Phys. org/10.1002/cssc.201403385.
Chem. B 119 (2015) 4739–4748, http://dx.doi.org/10.1021/acs.jpcb.5b01121. [77] N. Akiya, P.E. Savage, Roles of water for chemical reactions in high-temperature
[51] S. Deguchi, K. Tsujii, K. Horikoshi, Crystalline-to-amorphous transformation of water, Chem. Rev. 102 (2002) 2725–2750.
cellulose in hot and compressed water and its implications for hydrothermal [78] A. Kruse, A. Gawlik, Biomass conversion in water at 330–410 °C and 30–50 MPa.
conversion, Green Chem. 10 (2008) 191 http://xlink.rsc.org/?DOI=b713655b. Identification of key compounds for indicating different chemical reaction path-
[52] C. Yamane, T. Aoyagi, M. Ago, K. Sato, K. Okajima, T. Takahashi, Two different ways, Ind. Eng. Chem. Res. 42 (2003) 267–279, http://dx.doi.org/10.1021/
surface properties of regenerated cellulose due to structural anisotropy, Polym. J. ie0202773.
38 (2006) 819–826. [79] C. Promdej, Y. Matsumura, Temperature effect on hydrothermal decomposition of
[53] Y. Ogihara, R.L. Smith Jr., H. Inomata, K. Arai, R.L. Smith, H. Inomata, K. Arai, glucose in sub- and supercritical water, Ind. Eng. Chem. Res. 50 (2011)
Direct observation of cellulose dissolution in subcritical and supercritical water 8492–8497, http://dx.doi.org/10.1021/ie200298c.
over a wide range of water densities (550–1000 kg/m3), Cellulose 12 (2005) [80] D.A. Cantero, C. Martínez, M.D. Bermejo, M.J. Cocero, Simultaneous and selective
595–606 http://link.springer.com/10.1007/s10570-005-9008-1. recovery of cellulose and hemicellulose fractions from wheat bran by supercritical
[54] L.K. Tolonen, P.A. Penttilä, R. Serimaa, H. Sixta, The yield of cellulose precipitate water hydrolysis, Green Chem. 17 (2015) 610–618, http://dx.doi.org/10.1039/
from sub- and supercritical water treatment of various microcrystalline celluloses, c4gc01359j.
Cellulose 22 (2015) 1715–1728, http://dx.doi.org/10.1007/s10570-015-0628-9. [81] A. Romero, D.A. Cantero, A. Nieto-Márquez, C. Martínez, E. Alonso, M.J. Cocero,
[55] R. Abdullah, K. Ueda, S. Saka, Hydrothermal decomposition of various crystalline Supercritical water hydrolysis of cellulosic biomass as effective pretreatment to
celluloses as treated by semi-flow hot-compressed water, J. Wood Sci. 60 (2014) catalytic production of hexitols and ethylene glycol over Ru/MCM-48, Green
278–286, http://dx.doi.org/10.1007/s10086-014-1401-7. Chem. 18 (2016) 4051–4062, http://dx.doi.org/10.1039/C6GC00374E.
[56] Y. Yu, H. Wu, Understanding the primary liquid products of cellulose hydrolysis in [82] R.J.A. Gosselink, E. De Jong, B. Guran, A. Abächerli, Co-ordination network for
hot-compressed water at various reaction temperatures, Energy Fuels 24 (2010) lignin – standardisation, production and applications adapted to market require-
1963–1971, http://dx.doi.org/10.1021/ef9013746. ments (EUROLIGNIN), Ind. Crops Prod. 20 (2004) 121–129, http://dx.doi.org/10.
[57] Y. Yu, H. Wu, Effect of ball milling on the hydrolysis of microcrystalline cellulose 1016/j.indcrop.2004.04.015.
in hot-compressed water, AIChE J. 57 (2011) 793–800, http://dx.doi.org/10. [83] M.P. Pandey, C.S. Kim, Lignin depolymerization and conversion: a review of
1002/aic.12288. thermochemical methods, Chem. Eng. Technol. 34 (2011) 29–41, http://dx.doi.
[58] L.P. Novo, J. Bras, A. García, N. Belgacem, A.A.S. Curvelo, Subcritical water: a org/10.1002/ceat.201000270.
method for green production of cellulose nanocrystals, ACS Sustain. Chem. Eng. 3 [84] J.H. Lora, W.G. Glasser, Recent industrial applications of lignin: a sustainable
(2015) 2839–2846, http://dx.doi.org/10.1021/acssuschemeng.5b00762. alternative to nonrenewable materials, J. Polym. Environ. 10 (2002) 39–48,
[59] L.K. Tolonen, P.A. Penttilä, R. Serimaa, A. Kruse, H. Sixta, The swelling and dis- http://dx.doi.org/10.1023/A:1021070006895.
solution of cellulose crystallites in subcritical and supercritical water, Cellulose 20 [85] J. Ralph, J. Peng, F. Lu, Isochroman structures in lignin: a new β-1 pathway,
(2013) 2731–2744 http://link.springer.com/article/10.1007/s10570-013-0072-7. Tetrahedron Lett. 39 (1998) 4963–4964, http://dx.doi.org/10.1016/S0040-
[60] T. Adschiri, S. Hirose, R. Malaluan, K. Arai, Noncatalytic conversion of cellulose in 4039(98)00968-X.
supercritical and subcritical water, J. Chem. Eng. Jpn. 26 (1993) 676–680, http:// [86] M.M. Campbell, R.R. Sederoff, Variation in lignin content and composition (me-
dx.doi.org/10.1252/jcej.26.676. chanisms of control and implications for the genetic improvement of plants), Plant
[61] M. Sasaki, T. Adschiri, K. Arai, Production of cellulose II from native cellulose by Physiol. 110 (1996) 3–13, http://dx.doi.org/10.1104/pp.110.1.3.

564
M.J. Cocero et al. The Journal of Supercritical Fluids 133 (2018) 550–565

[87] W.-J. Liu, H. Jiang, H.-Q. Yu, Thermochemical conversion of lignin to functional [98] T.L.K. Yong, M. Yukihiko, Kinetic analysis of lignin hydrotermal conversion in sub-
materials: a review and future directions, Green Chem. 17 (2015) 4888–4907, and supercritical water, Ind. Eng. Chem. Res. 52 (2013) 9048–9059.
http://dx.doi.org/10.1039/C5GC01054C. [99] T.L.K. Yong, Y. Matsumura, Reaction kinetics of the lignin conversion in super-
[88] E.A.B. da Silva, M. Zabkova, J.D. Araújo, C.A. Cateto, M.F. Barreiro, critical water, Ind. Eng. Chem. Res 51 (2012) 11975–11988.
M.N. Belgacem, A.E. Rodrigues, An integrated process to produce vanillin and [100] B. Zhang, H.J. Huang, S. Ramaswamy, Reaction kinetics of the hydrothermal
lignin-based polyurethanes from Kraft lignin, Chem. Eng. Res. Des. 87 (2009) treatment of lignin, Appl. Biochem. Biotechnol. 147 (2008) 119–131, http://dx.
1276–1292, http://dx.doi.org/10.1016/j.cherd.2009.05.008. doi.org/10.1007/s12010-007-8070-6.
[89] G. Gonzalez, J. Salvado, D. Montane, Reactions of vanillic acid in sub- and su- [101] J. Li, G. Henriksson, G. Gellerstedt, Lignin depolymerization/repolymerization
percritical water, J. Supercrit. Fluids 31 (2004) 57–66, http://dx.doi.org/10. and its critical role for delignification of aspen wood by steam explosion,
1016/j.supflu.2003.09.015. Bioresour. Technol. 98 (2007) 3061–3068, http://dx.doi.org/10.1016/j.biortech.
[90] G.L. Huppert, B.C. Wu, S.H. Townsend, M.T. Klein, S.C. Paspek, Hydrolysis in 2006.10.018.
supercritical water: identification and implications of a polar transition state, Ind. [102] S. Kang, X. Li, J. Fan, J. Chang, Classified separation of lignin hydrothermal li-
Eng. Chem. Res. 28 (1989) 161–165, http://dx.doi.org/10.1021/ie00086a006. quefied products, Ind. Eng. Chem. Res. 50 (2011) 11288–11296, http://dx.doi.
[91] Wahyudiono, T. Kanetake, M. Sasaki, M. Goto, Decomposition of a lignin model org/10.1021/ie2011356.
compound under hydrothermal conditions, Chem. Eng. Technol. 30 (2007) [103] T. Faravelli, A. Frassoldati, G. Migliavacca, E. Ranzi, Detailed kinetic modeling of
1113–1122, http://dx.doi.org/10.1002/ceat.200700066. the thermal degradation of lignins, Biomass Bioenergy 34 (2010) 290–301, http://
[92] L. Pan, Z. Shen, L. Wu, Y. Zhang, X. Zhou, F. Jin, Hydrothermal production of dx.doi.org/10.1016/j.biombioe.2009.10.018.
formic and acetic acids from syringol, J. Zhejiang Univ. Sci. A 11 (2010) 613–618, [104] Wahyudiono, M. Sasaki, M. Goto, Thermal decomposition of guaiacol in sub- and
http://dx.doi.org/10.1631/jzus.A1000043. supercritical water and its kinetic analysis, J. Mater. Cycles Waste Manag. 13
[93] K. Yoshida, J. Kusaki, K. Ehara, S. Saka, Characterization of low molecular weight (2011) 68–79, http://dx.doi.org/10.1007/s10163-010-0309-6.
organic acids from beech wood treated in supercritical water, Appl. Biochem. [105] J. Miller, L. Evans, J.E. Mudd, K.A. Brown, Batch microreactor studies of lignin
Biotechnol. 121–124 (2005) 795–806. depolymerization by bases. 2. Aqueous solvents, Sandia Natl. Rep. Sand (2002)
[94] Z. Fang, T. Sato, R.L. Smith, H. Inomata, K. Arai, J.A. Kozinski, Reaction chemistry 1318, http://dx.doi.org/10.2172/800964.
and phase behavior of lignin in high-temperature and supercritical water, [106] V.M. Roberts, V. Stein, T. Reiner, A. Lemonidou, X. Li, J.A. Lercher, Towards
Bioresour. Technol. 99 (2008) 3424–3430, http://dx.doi.org/10.1016/j.biortech. quantitative catalytic lignin depolymerization, Chem. A Eur. J. 17 (2011)
2007.08.008. 5939–5948, http://dx.doi.org/10.1002/chem.201002438.
[95] M. Saisu, T. Sato, M. Watanabe, T. Adschiri, K. Arai, Conversion of lignin with [107] A. Toledano, L. Serrano, J. Labidi, Organosolv lignin depolymerization with dif-
supercritical water – phenol mixtures, Energy Fuels (2003) 922–928. ferent base catalysts, J. Chem. Technol. Biotechnol. 87 (2012) 1593–1599, http://
[96] K. Okuda, M. Umetsu, S. Takami, T. Adschiri, Disassembly of lignin and chemical dx.doi.org/10.1002/jctb.3799.
recovery – rapid depolymerization of lignin without char formation in water- [108] B.F. Ward, Tall oil- chemicals from a natural, renewable source, Proc. Eight Cellul.
phenol mixtures, Fuel Process. Technol. 85 (2004) 803–813, http://dx.doi.org/10. Conf. Wood Chem. Chall. 378 (1975) 332–334.
1016/j.fuproc.2003.11.027. [109] J.E. Miller, L. Evans, A. Littlewolf, D.E. Trudell, Batch microreactor studies of
[97] Wahyudiono, M. Sasaki, M. Goto, Recovery of phenolic compounds through the lignin and lignin model compound depolymerization by bases in alcohol solvents,
decomposition of lignin in near and supercritical water, Chem. Eng. Process. Fuel 78 (1999) 1363–1366, http://dx.doi.org/10.1016/S0016-2361(99)00072-1.
Process Intensif. 47 (2008) 1609–1619, http://dx.doi.org/10.1016/j.cep.2007.09. [110] X.E. Iriarte, Study of Lignin as High Added Value Chemical Compounds Source,
001. Thesis, UPV, San Sebastian Spain, 2014.

565

Anda mungkin juga menyukai