You are on page 1of 7

CIS-01444; No of Pages 7

Advances in Colloid and Interface Science xxx (2014) xxx–xxx

Contents lists available at ScienceDirect

Advances in Colloid and Interface Science


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

Brief overview on cellulose dissolution/regeneration interactions


and mechanisms
Bruno Medronho a,⁎, Björn Lindman b,c
a
IBB-CGB, Faculty of Sciences and Technology, Ed. 8, University of Algarve, Campus de Gambelas, Faro 8005-139, Portugal
b
Division of Physical Chemistry, Department of Chemistry, Center for Chemistry and Chemical Engineering, Lund University, Lund SE-221 00, Sweden
c
Materials Science and Engineering, Nanyang Technological University, Singapore 639798, Singapore

a r t i c l e i n f o a b s t r a c t

Available online xxxx The development of cellulose dissolution/regeneration strategies constitutes an increasingly active research field.
These are fundamental aspects of many production processes and applications. A wide variety of suitable solvents
Keywords: for cellulose is already available. Nevertheless, most solvent systems have important limitations, and there is an
Cellulose dissolution intense activity in both industrial and academic research aiming to optimize existing solvents and develop new
Cellulose regeneration ones. Cellulose solvents are of highly different nature giving great challenges in the understanding of the subtle
Hydrophobic interactions
balance between the different interactions. Here, we briefly review the cellulose dissolution and regeneration
Hydrogen bonding
Solvents
mechanisms for some selected solvents. Insolubility is often attributed to strong intermolecular hydrogen bond-
ing between cellulose molecules. However, recent work rather emphasizes the role of cellulose charge and the
concomitant ion entropy effects, as well as hydrophobic interactions.
© 2014 Published by Elsevier B.V.

Contents

1. Cellulose generalities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
2. Solvents used in cellulose dissolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
3. Cellulose dissolution: on the mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4. Cellulose regeneration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
5. Final remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0

1. Cellulose generalities chain stiffness, while, on the other hand, intermolecular hydrogen bonds
allow the linear polymer molecules to assemble in sheet-like struc-
Cellulose is a readily available and renewable biopolymer abundant- tures [3]. A more recent perspective highlights the amphiphilic nature
ly found in nature, typically combined with lignin and hemicelluloses in of cellulose (Fig. 2) [4–6]; the equatorial direction of a glucopyranose
the cell wall of upper parts of plants [1]. This homopolysaccharide is ring has a hydrophilic character because all three hydroxyl groups are lo-
formed by linearly connecting D-glucose units condensed through cated on the equatorial positions of the ring. On the other hand, the axial
β(1–4) glycosidic bonds (Fig. 1). The degree of polymerization (DP) direction of the ring is hydrophobic since the hydrogen atoms of C–H
can vary considerably depending on the source (i.e., DP from 100 up bonds are located on the axial positions of the ring. Thus, cellulose mole-
to 20000) [2]. From the single anhydroglucopyranose unit, AGU, up to cules have an intrinsic structural anisotropy where the rather flat ribbons
the micro and macro fibrils, cellulose organizes in a rather complex present sides with clear differences in polarity [7–9]. Such structural an-
fashion where an extended intra- and intermolecular network of hydro- isotropy is expected to considerably influence both the microscopic and
gen bonds is indicated as the basis of cohesion between cellulose mole- macroscopic properties of cellulose but has largely been neglected in
cules (Fig. 1). It is believed that intramolecular hydrogen bonds provide most discussions.
Cellulose can be considered as a semi-crystalline polymer with amor-
⁎ Corresponding author. Tel.: +351 289800910; fax: +351 289818419. phous regions, of low order, coexisting with higher order crystalline
E-mail address: bfmedronho@ualg.pt (B. Medronho). domains [10]. The origin and pretreatments of a cellulose sample

http://dx.doi.org/10.1016/j.cis.2014.05.004
0001-8686/© 2014 Published by Elsevier B.V.

Please cite this article as: Medronho B, Lindman B, Brief overview on cellulose dissolution/regeneration interactions and mechanisms, Adv Colloid
Interface Sci (2014), http://dx.doi.org/10.1016/j.cis.2014.05.004
2 B. Medronho, B. Lindman / Advances in Colloid and Interface Science xxx (2014) xxx–xxx

Fig. 1. Molecular structure of cellulose where the extended network of intra and inter-hydrogen bonding is represented. The anhydrocellobiose unit (i.e., disaccharide of two glucose
molecules) is highlighted.

determines the degree of crystallinity of cellulose, which typically is cellulose is rather difficult. Cellulose does not melt nor is it soluble in com-
found between 40% and 60%. Several suggestions on how these crystalline mon aqueous and organic solvents [2,17]. However, this polysaccharide is
and non-crystalline regions are intermixed have been developed over the soluble in more exotic media with no apparent common properties [18].
years, such as single crystals or uniform elementary fibrils, but nowadays, The more consensual vision among leaders in the field has been that the
the so-called “fringed fibrillar model” is widely accepted [11–16]. In this key to dissolve cellulose resides in the solvent capacity to break the above
model, the cellulose nanofibril is not regarded a single crystal but rather mentioned intra- and intermolecular hydrogen bond network [19]. Other
as a less structured arrangement of non-uniform crystalline segments interactions among cellulose molecules have been mostly ignored. Re-
accompanied by amorphous parts, both longitudinally and laterally cently, we have reanalyzed this problem and argued against this accepted
displaced [12]. picture [4–6]. Instead, we have concluded that cellulose has clear amphi-
Apart from being used in unmodified forms, such as wood or cotton, philic properties and a careful examination of the interactions involved
cellulose can be extracted from its natural sources and either used in the suggests that hydrophobic interactions play a significant role in governing
paper industry or, in a smaller scale, in some specific applications such cellulose solubility.
as regenerated fibers (e.g., Lyocell or viscose). In fact, it is believed
that forest-based raw materials, such as cellulose, can play a major 2. Solvents used in cellulose dissolution
role in replacing fossil oil-based fibers and cotton by new ecological
man-made fibers in both woven and non-woven end applications. The Why is cellulose dissolution so important? The following reasons
success of this replacement is intimately dependent on the develop of can be enumerated: preparation of regenerated and innovative mate-
new solvents and strategies for dissolution and regeneration. Cellulose rials such as fibers (e.g., textile applications) and films (e.g., packaging
may also be modified by chemical, enzymatic or microbiological applications), production of valuable cellulose derivatives in a homoge-
methods to obtain new valuable derivatives and materials. Large-scale nous environment (note that typical solvents cannot penetrate inside
production of cellulose derivatives (mainly ethers and esters) and crystalline regions of cellulose and heterogeneous modification is re-
regenerated materials (i.e., fibers, films, food casing, membranes, and stricted only to the surface of the crystallites), and finally, to degrade
sponges, among others) find applications in several important commer- cellulose more efficiently (e.g., important for, for instance, biorefinery
cial areas such as the membrane, polymer and paint industries [3]. purposes).
A challenging issue is the fact that many important applications Typically, cellulose dissolution is preceded by polymer swelling,
of cellulose involve its dissolution. Due to the complexity of such a which is defined as a process where the solvent molecules penetrate
biopolymeric network, the partially crystalline structure and the extend- and labilize the cellulose structure to a certain extent, leaving the vol-
ed non-covalent interactions among molecules, chemical processing of ume and physical properties of the biopolymer significantly altered

Fig. 2. Hydrophilic and hydrophobic parts of the cellulose molecule: left, top view of the glucopyranose ring plane highlighting hypothetic hydrogen bonding between the hydroxyl groups
located on the equatorial positions of the intermediate ring; right, side view of the glucopyranose ring plane showing the hydrogen atoms of C–H bonds on the axial positions of the ring.

Please cite this article as: Medronho B, Lindman B, Brief overview on cellulose dissolution/regeneration interactions and mechanisms, Adv Colloid
Interface Sci (2014), http://dx.doi.org/10.1016/j.cis.2014.05.004
B. Medronho, B. Lindman / Advances in Colloid and Interface Science xxx (2014) xxx–xxx 3

while the solid, or semi-solid fractions, remain practically unchanged. cellulose and the highlighted role of hydrophobic interactions and this
On the other hand, complete dissolution is expected to fully destroy has been recently discussed [4–6,29].
the supramolecular structure, resulting in a solution where the polymer Let us now discuss the suggested mechanisms in literature for some
is, at best, molecularly dispersed. Interestingly, the same solvent may of the classical solvents used for cellulose dissolution. In the aqueous
act as a dissolving agent or merely as a swelling agent, depending on complexing systems, such as the cuprammonium hydroxide, “cuam,”
the conditions used in the experiment (i.e., concentrations, tempera- the suggested mechanism assumes that the metal complexes dissolve
ture, etc.) [6]. cellulose via initial deprotonation of the hydroxyls followed by a coordi-
The dissolution of a polymer, such as cellulose, in a solvent is, of native binding to the hydroxyl groups in the C2 and C3 positions of each
course, governed by the free energy of mixing [20]. The mixing process anhydroglucose unit [30].
will occur spontaneously when the free energy change on mixing is In the case of the cold alkali, one of the leading opinions is that NaOH
negative. Otherwise, phase separation may result from the mixing pro- forms hydrates with water capable to break the inter- and intramolecu-
cess. The polymer molecular weight is a key parameter in dissolution; lar hydrogen bonds between cellulose molecules [31]. These hydrates
the higher the molecular weight, the weaker is the entropic driving would have the capacity to bond with one or two hydroxyl groups of
force contribution for dissolution [21]. Under these conditions, the en- each AGU. On the other hand, our vision is that the charging effect, in
thalpy term is crucial in determining the sign of the Gibbs free energy solvent systems such as the alkali, should not be neglected [32]; when
change. One should mention that polymer dissolution is often con- a neutral molecule, such as cellulose, is charged up, its solubility is ex-
trolled by kinetics rather than by thermodynamics and cellulose is a pected to increase and thus cellulose tends to be more soluble/be
very clear example of this. From a thermodynamic point of view, a rea- more penetrated at either high or low pH. This is because the dissociat-
sonable solvent for cellulose dissolution must be able to overcome the ed counterions strongly contribute to the translational entropy of
low entropy gain by favorable solvent/polymer interactions. In practice, mixing. We note that this system is limited to moderately low DPs
finding a non-derivatizing solvent that does not reduce the DP nor does and actually this has been used to explain the difficulties faced in dis-
react with the cellulose is one the most challenging parts in cellulose ruption of the long-range order inherent to the higher DPs. Another par-
dissolution. From the traditional viscose route, which is generally a ticularity of the cold alkali track, is that the concentration range where
slow process and environmentally hostile (e.g., discharge of toxic NaOH works as a direct solvent is very narrow. This has been discussed
gases), up to state-of-art ionic liquid systems, several derivatizing and on the basis of a concentration dependent size of the above mentioned
non-derivatizing (aqueous and non- aqueous) solvents, with strikingly NaOH/water hydrates [33]; at lower alkali concentrations, these hy-
unrelated properties, have been developed [6]. The list is vast and drates are speculated to have hydrodynamic diameters too large to pen-
includes quite unusual combinations and experimental conditions: etrate and diffuse into the very densely packed crystalline regions of
simple or multicomponent mixtures, aqueous and organic media, inor- cellulose. The performance of these alkali systems can be increased
ganic and organic salts, high and low temperatures, high and low pHs, when certain additives, such as zinc oxide (ZnO), urea and thiourea
etc. A few classical examples of non-derivatizing solvents can be men- are present even if their role is still unclear. ZnO has been suggested to
tioned, such as aqueous inorganic complexes (e.g., cuprammonium hy- form strong hydrogen bonds between cellulose and Zn(OH)2− 4 , a stable
droxide), concentrated salt solutions (e.g., zinc chloride, ammonium, zinc hydroxide species in solution [34]. Regarding urea or thiourea,
calcium and sodium thiocyanate solutions), salts dissolved in organic some authors argue that urea hydrates possibly self-assemble at the
solvents (e.g., lithium chloride/N,N-dimethylacetamide, ammonia/ surface of the NaOH hydrogen-bonded cellulose to form an inclusion
ammonium salt, tetrabutylammonium fluoride/dimethyl sulfoxide), complex (IC), which would lead to full dissolution [35–39]. Zhou et al.
aqueous alkali (lithium hydroxide or sodium hydroxide solutions), etc. state that the role of urea and thiourea is to function as hydrogen
[22]. A more detailed and updated list can be find elsewhere [1,6,23]. bond donors and acceptors to prevent the reassociation of cellulose
What most of these solvents have in common, for different reasons, is molecules, therefore conferring stability to the solution [40]. Molecular
the fact that they are not easily scaled up. Recently, developed processes dynamic (MD) simulations indicate that, in contrast to water molecules,
for producing regenerated cellulose fibers are emerging as possible the cellulose chain prefers to interact with urea molecules [41]. A stron-
alternatives to the viscose process. Among them, the non-derivatizing ger and more stable interaction of cellulose with urea would decrease
N-methylmorpholine-N-oxide (NMMO) process [24,25] has been the self-interaction of cellulose chains and promote dissolution of cellu-
successful on the industrial scale, even if the recovery of the NMMO sol- lose in urea-containing solvent mixtures. Kunze and Fink [42] argued
vent is complicated, energy demanding, and costly. These are reasons that NaOH and urea hydrates work in synergy to separate the cellulose
enough to believe that this system does not meet all the requirements chains from each other. They are also supposed to prevent cellulose
for a complete replacement of the viscose technology. Ionic liquids from regenerating its intermolecular hydrogen bonding by forming a
(ILs) have been extensively studied in recent years since these non- stable “hydrate coat” on their surface. The same conclusion is shared
volatile organic solvents are capable of dissolving large amounts of by Isobe et al. [43], which have recently reanalyzed the role of urea in
cellulose and other carbohydrates [26]. However, there are some im- NaOH solvent system. These authors have concluded that urea has no
portant limiting factors for large-scale applications of ILs such as their direct interaction with cellulose but helps the alkali to penetrate into
high cost of production, high viscosity, sensitivity to moisture content cellulose crystalline regions, stabilizing the swollen cellulose molecules.
and poorly developed appropriate purification processes [27,28]. In addition, the authors propose that the stabilizing effect of urea de-
rives from the fact that urea hinders the hydrophobic association of cel-
lulose [43]. These ideas follow the established role of urea in protein
3. Cellulose dissolution: on the mechanisms denaturation [44]. Recent theoretical work, based on molecular model-
ing [45] and Kirkwood–Buff solution theory [46], suggests a preferential
Regarding the mechanism of dissolution, we already mentioned association of urea to the hydrophobic sites of proteins. This alternative
above that the widely accepted picture considers, almost exclusively, idea highlights the amphiphilic properties of cellulose, and therefore, it
that cellulose dissolution results from the solvent ability to eliminate is believed that urea may weaken the entropic effect by accumulating in
the inter- and intramolecular hydrogen bonds among biopolymer mol- the proximity of the hydrophobic surface of cellulose leading to a higher
ecules. On the other hand, cellulose can be regarded as an amphiphilic thermal stability of cellulose in aqueous alkali solvent [43]. These
molecule and thus research aiming in creating a basis for the develop- suggestions clearly support our recent discussions, which emphasize
ment of new solvents should focus, not only in eliminating hydrogen the role of hydrophobic interactions in cellulose insolubility [4–6,29].
bonding but, more importantly, on eliminating hydrophobic interac- Another important variable in NaOH-based solvents is the relatively
tions. There are several arguments to believe in an amphiphilicity of low temperature needed to efficiently dissolve cellulose in opposition to

Please cite this article as: Medronho B, Lindman B, Brief overview on cellulose dissolution/regeneration interactions and mechanisms, Adv Colloid
Interface Sci (2014), http://dx.doi.org/10.1016/j.cis.2014.05.004
4 B. Medronho, B. Lindman / Advances in Colloid and Interface Science xxx (2014) xxx–xxx

the standard picture of solubility, which implies that the entropic system where the solubilization of the biopolymer is expected to be en-
driving force is strengthened with increasing temperature. Typically, hanced by the electrostatic repulsion between the negatively-charged
sub-zero temperatures are required for dissolution in soda [47]. This cellulose chains (due to the condensation of F−) [53].
phenomenon has been analyzed thermodynamically by the total en- Regarding ionic liquids, although great experimental and computa-
thalpy of cellulose dissolution [48]. It was suggested that the only endo- tional progress has been made, there is no clear understanding of the
thermic term is associated with breaking the hydrogen bonds in the role of individual ionic species in dissolution. Up to date, two main
crystalline regions. All other terms relate to interactions between cellu- views of the interactions between ILs and cellulose prevail: (1) the dis-
lose hydroxyl groups and the solvent system and are exothermic. There- solution process is governed by the interactions between the anion and
fore, the conclusion is that the overall process of cellulose dissolution is the biopolymer, with no specific role for the cation; and (2) the major
exothermic and is thus favored by a lower temperature. Additionally, driving force for cellulose dissolution comes from the H-bond interac-
the network of solvent hydrates (i.e., Na+ and OH− ions surrounded tions of the cellulose hydroxyls with the anion and the cation of ILs. In
by a “cage” of water molecules) is suggested to be temperature depen- literature, we find several examples where either changing the cation
dent probably due to the increasing hydrogen bonding strength [49]. As or the anion produces remarkable effects on dissolution efficiency. In-
mentioned above, some authors believe that this network of solvent terestingly, some of the good hydrogen-bond-accepting anions that
hydrates disrupts mainly the closely chain packed cellulose through have been found to dissolve cellulose can be rendered ineffective
the formation of new hydrogen bonds between cellulose and NaOH through pairing with certain cations [54]. One of the structural features
hydrates. Thus, when the temperature is raised, the hydrogen bonds that makes ILs so interesting is their strong asymmetry. Typically, the cat-
are assumed to be weaker and the network of hydrates is gradually ions are bulky species with amphiphilic properties. Such amphiphilicity is
destroyed [49]. Regarding the temperature dependence, our description normally not considered when discussing the mechanism of dissolution
of the phenomenon is based on temperature-dependent conformation- of cellulose. This is particularly relevant since crystalline cellulose has an
al changes in the cellulose chain, which would make the polymer less amphipathic-like structure (Fig. 1b): hydrophobic surfaces consisting
polar at higher temperatures, thus reducing the attractive interactions of pyranose-ring hydrogens and hydrophilic regions arising from the
with the polar solvent and vice-versa [5]. Pinkert et al. [50] offers a hydroxyl groups directed toward the sides of the ring. The non-polar
different view based on the idea that the inability of NaOH-based sol- surfaces can be organized into hydrophobic sheets paired against one
vent systems to dissolve cellulose at increased temperatures strongly another, rather than structuring large amounts of water in solution.
suggests that an ordered solvent state of some kind is required to (Note that in this type of interaction, the driving force for association is
achieve dissolution. not simply van der Waals' interactions but rather hydrophobic associa-
In the case of the NMMO solvent, the active moiety is the NO group tion driven by the liberation of structured water molecules [55].) We be-
with its strongly dipolar character. It is believed that the oxygen in this lieve that the dissolution of an amphiphilic polymer, such as cellulose,
group is able to form one or two hydrogen bonds with AGU of cellulose. would be facilitated in amphiphilic solvents. Therefore, the amphiphilic
The proposed mechanism assumes the cleavage of intermolecular hy- properties of all cations in ILs fit in our suggestion. In fact, this also follows
drogen bonds of cellulose and the formation of a soluble complex of from the earlier discussion on the effect of additives such as PEG, urea,
stronger hydrogen bonds between the cellulosic hydroxyls and the NO and thiourea on dissolution in alkali solutions.
group of NMMO [18].
For mixed inorganic/organic systems such as lithium chloride/
dimethylacetamide (LiCl/DMAc), several dissolution mechanisms have 4. Cellulose regeneration
been suggested [51,52]. Generally, the dissolving mechanism is believed
to go via an intermediate involving the interaction of Cl− (due to its ba- Typically, the regeneration of cellulose occurs when contacting the
sicity) with cellulose in addition to an exchange of DMAc in the lithium cellulose solution with a coagulation bath. The polymer profile at the
co-ordination sphere by cellulose hydroxyl groups (Fig. 3). Accumula- point of precipitation exhibits a very high interfacial concentration,
tion of C1− along the cellulose chain produces a negatively charged thus favoring the formation of a dense polymer “skin.” The bulk of the
polymer with the macrocation, [Li-DMAc], as the counterion. This sample is at near the initial concentration and is in a fluid state. Thus,
process can be regarded as a polyelectrolyte effect, where polymer a rapid inflow of the coagulant can take place through the weak points
molecules are forced apart due to charge repulsion. Moreover, the at the skin interface. Rapid growth of finger-like voids in the fluid region
high osmotic pressure drives a continuous influx of solvent resulting is expected to occur due to the moving interface created by the coagu-
in further disruption of the cellulose binding forces until the polymer lant (less viscous) and the solution (more viscous). The kinetics of
is totally solvated. A similar polyelectrolyte effect is suggested for regeneration is mainly controlled by the relative velocities of the
the tetrabutylammonium fluoride/dimethyl sulfoxide (TBAF/DMSO) counter-diffusion process—diffusion of the solvent from the solution

Fig. 3. Example of a hydrogen-bond breaking mechanism for the cellulose dissolution in the lithium chloride/dimethylacetamide (LiCl/DMAc) solvent system. Cellulose hydroxyl groups
interact with an intermediate (Li+-DMAC macrocation) via hydrogen bonding bridged by the chloride anion. The lithium cation interacts with the carbonyl oxygen via an ion–dipole
interaction.

Please cite this article as: Medronho B, Lindman B, Brief overview on cellulose dissolution/regeneration interactions and mechanisms, Adv Colloid
Interface Sci (2014), http://dx.doi.org/10.1016/j.cis.2014.05.004
B. Medronho, B. Lindman / Advances in Colloid and Interface Science xxx (2014) xxx–xxx 5

into the coagulation bath and the non-solvent from the bath into the so- bonding to form Na–cellulose type IV crystallites, a hydrate form of cel-
lution [56]. lulose II (Fig. 4).
The exchange of solvent with non-solvent leads to a desolvation of This constitutes the first experimental evidence of the development
the cellulose molecules and to the supposed reformation of the intra- of hydrophobically stacked monomolecular sheets which was firstly hy-
and inter-molecular hydrogen bonds [57]. The regeneration of cellulose pothesized by Hermans [64] and later by Hayashi et al. [65]. These au-
from, for instance, NMMO solutions using water as the coagulation sys- thors found that structural disorders lie mainly in the hydrogen-
tem results from a phase separation process. Typically, two mechanisms bonded intermolecular region and assumed the primary structure as a
of phase separation can take place during liquid-liquid demixing of monomolecular sheet formed by stacking of glucopyranoside planes
polymer solutions; either nucleation and growth (i.e., the nuclei of by van der Waals forces. Recently, the theoretical work of Miyamoto
one phase grow in the mixture) or spinodal decomposition (i.e., a peri- et al. [9] simulated the regeneration of cellulose by MD, reproducing
odic variation of concentration leads to the final phase separation). In the hypothesis of Hermans and Hayashi. While these studies gave a con-
the NMMO-water system, if there is a clear difference between these sistent picture of the primary structure formation of regenerated cellu-
two mechanisms in their first stages (nucleation and growth shows iso- lose, they were either inferred from the resulting structure or based on
lated entities while spinodal decomposition has a 3D network-like mor- computer simulations. Therefore, the work of Isobe et al. [59] not only
phology), typically both tend to the same morphology at the end due to shines light into the regeneration mechanism of cellulose but also con-
surface tension effects [58]. stitutes an alternative vision to the typical regeneration mechanism
The mechanical and surface chemical properties of regenerated cel- found in literature which, as we have seen, essentially focuses exclu-
lulose are known to depend strongly on the type of cellulose solvent and sively on the reformation of the broken inter- and intramolecular hy-
coagulant. For a bibliography, the reader is directed to the work of Isobe drogen bonds among cellulose molecules.
et al. [59] and references therein. For instance, in molten salts systems,
the regenerated morphologies can vary significantly depending on the 5. Final remarks
salt used; in thiocyanate melts, fiber-like samples (similar to the raw
cellulose) have been obtained while lamellar-like samples and layered From this brief review on cellulose dissolution and regeneration, it is
structures were regenerated from LiClO4 · 3H2O and chlorides, respec- clear that quite contrasting starting points have been chosen through
tively [60]. the years for the analysis of the mechanisms. As can be inferred, the
In IL systems, the presence of water affects many of the solvent prop- nature of the solvents that can efficiently dissolve cellulose covers an
erties. To date, experimental studies on cellulose/IL/non-solvent ternary enormous range making it difficult to provide a single consistent mech-
systems have not been conclusive, and a mechanistic understanding of anism. While we recognize that there may be different mechanisms for
the interactions between water, IL, and cellulose remains unknown. different cases, it seems important to make a classification based on
For instance, MD results are taken to indicate that the ordered cation- some simple views on solubility, in general, and polymer solubility, in
anion polar interaction network is disrupted by water that forms a particular.
network of interactions with the anion and with the cation [61]. Solubility is governed by entropy, notably translational and configu-
More specifically, the suggested mechanism assumes that while rational entropy, the enthalpy in general, being unfavorable. Regarding
water diffuses inside the first solvation shell of cellulose, the number translational entropy, this will be very much higher for ionic polymers
of hydrogen bonds among water molecules and the polymer in- with dissociating counterions than for non-ionic polymers. This is
creases and the number of H-bonds between the anion and the amply demonstrated for cellulose where protonation or deprotonation,
sugar decreases. Meanwhile, the cations are considered to be main- due to addition of simple acids and bases, offers efficient ways of swell-
tained in the second solvation shell of cellulose due to their strong ing and dissolution of cellulose. An effective ionization can be obtained
electrostatic interactions with the anions present. The formation of not only from changes in the state of protonation but also from associa-
this network displaces the cation out of the first solvation shell lead- tion/adsorption of an ionic species. Thus, it is well known that the solu-
ing to cellulose precipitation. In this model, phase separation of the bility of non-ionic polymers, such as poly(ethylene glycol) and cellulose
cellulose as a function of water results from competition between derivatives, can be strongly increased by association of ions. This is par-
the water–anion, cellulose–anion, cellulose–water, cellulose–cation, ticularly efficient for ionic surfactants but also ions with a high polariz-
and anion–cation interactions [61]. ability can act as pseudo-surfactants. Actually, this is the basis for the
The regeneration mechanism for alkali systems suggests that the in- lyotropic series, often referred to as the Hofmeister series. Thus, it is
clusion complex associated with cellulose, NaOH, and, in some cases, found that large anions, such as iodide, can increase the solubility of
urea or thiourea hydrates is disrupted by adding a non-solvent such as non-ionic macromolecules, whereas small ions with a high charge den-
water, leading to the self-association of cellulose. The regenerated cellu- sity, have the opposite effect. The large polarizable ions are effectively
lose film is said to be formed through a rearrangement of the hydrogen hydrophobic and tend to have a low solubility in a polar solvent and/
bonds [62,63]. Moreover, in acidic non-solvents, the H+ assumes a key or being enriched at a hydrophobic surface. Generally speaking, we
role to trigger cellulose regeneration by neutralizing the alkaline con- can thus expect that a highly unsymmetrical electrolyte would provide
tent. Additionally, as in the Lenzing viscose production process, the ad- good solubility conditions for a non-ionic polymer such as cellulose,
dition of a strong electrolyte to H2SO4, such as Na2SO4 and ZnSO4, can since the ions would be non-uniformly distributed in the solution.
reduce the H+ concentration in the coagulant, leading to a counter- This is the case for molten salt systems with small, strong polarizing cat-
diffusion rate and slower acid–alkali neutralization process than by ions (i.e., Li+ and Zn2+) and large polarizable anions but, from the broad
just H2SO4 alone. This will considerably influence the final properties overview above, looks like a rather general trend since many of the effi-
of the regenerated cellulose material [10]. cient solvents of cellulose follow this pattern, including many of the
Recently, a very interesting study has been presented by Isobe et al. ionic liquids.
[59]. The regeneration of cellulose, either using a coagulant or upon As we have previously argued, cellulose manifests strong hydropho-
heating, was followed in an aqueous alkali-urea solvent by monitoring bic properties and is thus amphiphilic [4,5]. As arguments for amphi-
the process by time-resolved synchrotron X-ray scattering. The authors philic properties we can mention the ability of other polyglucoses,
suggested that when the medium surrounding the cellulose molecules such as cyclodextrin and amylose, to solubilize non-polar and amphi-
becomes energetically unfavorable for molecular dispersion, regenera- philic compounds in aqueous solution. It is also strongly supported by
tion starts, and the initial process would consist in stacking the hydro- simulations [66]. Another clear support for amphiphilic properties can
phobic glucopyranoside rings (driven by hydrophobic interactions) to be found from studies of surfactant binding to cellulose derivatives.
form monomolecular sheets, which then would line up by hydrogen Thus, surfactants bind to cellulose derivatives, soluble due to

Please cite this article as: Medronho B, Lindman B, Brief overview on cellulose dissolution/regeneration interactions and mechanisms, Adv Colloid
Interface Sci (2014), http://dx.doi.org/10.1016/j.cis.2014.05.004
6 B. Medronho, B. Lindman / Advances in Colloid and Interface Science xxx (2014) xxx–xxx

Complete
regeneration

Intensity (a.u.)

Time
control

q (Å-1)

Fig. 4. On top, the time evolution of synchrotron x-ray diffraction profiles of a cellulose solution under regeneration in a 5 wt% Na2SO4 aqueous solution is shown. In the control (10 wt.%
cellulose solution without coagulant) no Bragg reflections are observed. After a certain time, a first Bragg reflection (110) is observed while the second one (020) is only visible after com-
plete regeneration. On bottom, a schematic illustration of cellulose regeneration deduced from the X-ray data, where first the hydrophobic stacking of cellulose molecules occurs and then
is followed by their binding via hydrogen bonding (adapted from reference [55]).

hydrophilic groups, which do not associate with surfactants. An illustra- Acknowledgments


tive example is that of cationic hydroxyethyl cellulose, which binds an-
ionic surfactant in excess of charge stoichiometry; obviously, this can The authors acknowledge support from the Portuguese Foundation
only be due to hydrophobic interactions between the cellulose back- for Science and Technology (FCT, project PTDC/AGR-TEC/4049/2012
bone and the alkyl chain of the surfactant [67]. In view of this, eliminat- and a postdoc grant assigned to B. Medronho, SFRH/BPD/74540/2010).
ing the hydrophobic interactions between cellulose molecules must be Financial support was also received from Södra Skogsägarnas Stiftelse,
very important for the aqueous solubility of cellulose. We note that Stiftelsen Nils och Dorthi Troëdssons forskningsfond and the CelluNova
urea can facilitate cellulose dissolution. Urea has a much lower polarity consortium.
than water and is well known to eliminate hydrophobic association in
water. As alluded to above, it acts as protein denaturant by reducing in-
tramolecular hydrophobic association. Furthermore, it inhibits hydro- References
phobic association of surfactants as can be seen from the increase in
critical micelle concentration on urea addition. Other substances of [1] Olsson C, Westman G. Direct Dissolution of Cellulose: Background, Means and
Applications; 2013.
intermediate polarity, such as urea derivatives and poly(ethylene [2] Marsh JT. An introduction to the chemistry of cellulose. 2nd ed. London: Chapman &
glycol), have similar effects and are well-known enhancers of the Hall; 1942.
aqueous solubility of cellulose. The alternative notion to hydrophobic in- [3] Klemm D, Heublein B, Fink HP, Bohn A. Cellulose: Fascinating biopolymer and
sustainable raw material. Angew Chem Int Ed 2005;44:3358–93.
teractions as the limiting factor of dissolution of cellulose in water is [4] Medronho B, Romano A, Miguel MG, Stigsson L, Lindman B. Rationalizing cellulose
hydrogen-bonding. However, this would require that the carbohydrate- (in)solubility: reviewing basic physicochemical aspects and role of hydrophobic
carbohydrate hydrogen-bonding would be much stronger than water- interactions. Cellulose 2012;19:581–7.
[5] Lindman B, Karlstrom G, Stigsson L. On the mechanism of dissolution of
carbohydrate and water-water hydrogen bonding, which is not the cellulose. J Mol Liq 2010;156:76–81.
case; regarding a solo hydrogen–bonding mechanism, it is striking that [6] Medronho B, Lindman B. Competing forces during cellulose dissolution: From
all these interactions are very similar in magnitude; ca. 5 kcal/mol. Fur- solvents to mechanisms. Curr Opin Colloid Interface Sci 2014;19:32–40.
[7] Biermann O, Hadicke E, Koltzenburg S, Muller-Plathe F. Hydrophilicity and lipophi-
thermore, a hydrogen-bonding mechanism is inconsistent with a high licity of cellulose crystal surfaces. Angew Chem Int Ed 2001;40:3822−+.
aqueous solubility of polysaccharides, which have as high hydrogen- [8] Yamane C, Aoyagi T, Ago M, Sato K, Okajima K, Takahashi T. Two different surface prop-
bonding capacity as cellulose. erties of regenerated cellulose due to structural anisotropy. Polym J 2006;38:819–26.

Please cite this article as: Medronho B, Lindman B, Brief overview on cellulose dissolution/regeneration interactions and mechanisms, Adv Colloid
Interface Sci (2014), http://dx.doi.org/10.1016/j.cis.2014.05.004
B. Medronho, B. Lindman / Advances in Colloid and Interface Science xxx (2014) xxx–xxx 7

[9] Miyamoto H, Umemura M, Aoyagi T, Yamane C, Ueda K, Takahashi K. Structural [38] Lue A, Liu YT, Zhang L, Potthas A. Light scattering study on the dynamic behaviour
reorganization of molecular sheets derived from cellulose II by molecular dynamics of cellulose inclusion complex in LiOH/urea aqueous solution. Polymer
simulations. Carbohydr Res 2009;344:1085–94. 2011;52:3857–64.
[10] Klemm D, Philipp B, Heinze T, Heinze U, Wagenknecht W. Fundamentals and [39] Cai J, Zhang L. Unique gelation behavior of cellulose in NaOH/Urea aqueous solution.
analytical methods. Comprehensive Cellulose Chemistry, vol 1. Weinheim: Wiley - Biomacromolecules 2006;7:183–9.
VCH; 1998. [40] Zhou JP, Zhang LN. Solubility of cellulose in NaOH urea aqueous solution. Polym J
[11] Fink HP, Hofmann D, Philipp B. Some aspects of lateral chain order in cellulosics 2000;32:866–70.
from x-ray-scattering. Cellulose 1995;2:51–70. [41] Cai L, Liu Y, Liang HJ. Impact of hydrogen bonding on inclusion layer of urea to cel-
[12] Hearle JWS. A fringed fibril theory of structure in crystalline polymers. J Polym Sci lulose: study of molecular dynamics simulation. Polymer 2012;53:1124–30.
1958;28:432–5. [42] Kunze J, Fink H-P. Structural changes and activation of cellulose by caustic soda so-
[13] Fink HP, Ganster J. Relations between structure and mechanical properties of lution with urea. Macromol Symp 2005;223:175–87.
cellulosic man-made fibres. Chemistry Seminar Proceedings: Stockholm, Sweden; [43] Isobe N, Noguchi K, Nishiyama Y, Kimura S, Wada M, Kuga S. Role of urea in alkaline
1994. dissolution of cellulose. Cellulose 2013;20:97–103.
[14] Lenz J, Schurz J, Wrentschur E. Properties and structure of solvent-spun and viscose- [44] Tanford C. Isothermal unfolding of globular proteins in aqueous urea solutions. J Am
type fibers in the swollen state. Colloid Polym Sci 1993;271:460–8. Chem Soc 1964;86 [2050-&.].
[15] Krassig HA. Struktur und reaktivität von cellulosefasern. Das Papier 1984;38:571–82. [45] Zangi R, Zhou RH, Berne BJ. Urea's action on hydrophobic interactions. J Am Chem
[16] Fink HP, Philipp B. Models of cellulose physical structure from the viewpoint of the Soc 2009;131:1535–41.
cellulose-I→Cellulose-II transition. J Appl Polym Sci 1985;30:3779–90. [46] Shimizu S. The effect of urea on hydrophobic hydration: preferential interaction and
[17] Krassig HA. Cellulose: structure, accessibility and reactivity. 1st ed. Amsterdam: the enthalpy of transfer. Chem Phys Lett 2011;517:76–9.
Gordon and Breach Science Publishers; 1993. [47] Sobue H. Kiessig HH. K Z Phys Chem 1939;43:309.
[18] Hudson SM, Cuculo JA. The solubility of unmodified cellulose—a critique of the [48] Wang Y, Deng YL. The kinetics of cellulose dissolution in sodium hydroxide solution
literature. J Macromol Sci R M C 1980;C18:1–82. at low temperatures. Biotechnol Bioeng 2009;102:1398–405.
[19] Zhang LN, Ruan D, Gao SJ. Dissolution and regeneration of cellulose in NaOH/ [49] Egal M, Budtova T, Navard P. Structure of aqueous solutions of microcrystalline
thiourea aqueous solution. J Polym Sci Polym Phys 2002;40:1521–9. cellulose/sodium hydroxide below 0 degrees C and the limit of cellulose dissolution.
[20] Grulke EA. Solubility parameter values. In: Brandrup J, Immergut EH, Grulke EA, ed- Biomacromolecules 2007;8:2282–7.
itors. Polymer handbook, 4th ed., vol. 7. New York, NY: Wiley; 1999. [50] Pinkert A, Marsh KN, Pang SS. Reflections on the solubility of cellulose. Ind Eng Chem
[21] Holmberg K, Jonsson B, Kronberg B, Lindman B. Surfactants and polymers in aqueous Res 2010;49:11121–30.
solution. Hoboken: Wiley; 2002. [51] Dawsey TR, Mccormick CL. The lithium chloride/dimethylacetamide solvent for
[22] Heinze T, Koschella A. Solvents applied in the field of cellulose chemistry—a mini cellulose—a literature-review. J Macromol SciR M C 1990;C30:405–40.
review. Polímeros: Cienc Tecnol 2005;15:84–90. [52] Sen S, Martin JD, Argyropoulos DS. Review of cellulose non-derivatizing solvent in-
[23] Liebert T. Cellulose Solvents - Remarkable History, Bright Future. In: Liebert TF, teractions with emphasis on activity in inorganic molten salt hydrates. Acs Sustain
Heinze TJ, Edgar KJ, editors. Acs Sym Ser., 1033; 2009. p. 3–54. Chem Eng 2013;1:858–70.
[24] Fink HP, Weigel P, Purz HJ, Ganster J. Structure formation of regenerated cellulose [53] Ostlund A, Lundberg D, Nordstierna L, Holmberg K, Nyden M. Dissolution and gela-
materials from NMMO-solutions. Prog Polym Sci 2001;26:1473–524. tion of cellulose in TBAF/DMSO solutions: the roles of fluoride ions and water.
[25] Perepelkin KE. Lyocell fibres based on direct dissolution of cellulose in N- Biomacromolecules 2009;10:2401–7.
methylmorpholine N-oxide: development and prospects. Fibre Chem 2007;39:163–72. [54] Pinkert A, Marsh KN, Pang SS, Staiger MP. Ionic liquids and their interaction with cel-
[26] Swatloski RP, Spear SK, Holbrey JD, Rogers RD. Dissolution of cellulose with ionic lulose. Chem Rev 2009;109:6712–28.
liquids. J Am Chem Soc 2002;124:4974–5. [55] Cousins SK, Brown RM. Polym Bull 1995;36:3885–8.
[27] Zavrel M, Bross D, Funke M, Buchs J, Spiess AC. High-throughput screening for ionic [56] Shen TC, Cabasso I. Macromolecular solutions — solvent-property relationships in
liquids dissolving (ligno-)cellulose. Bioresour Technol 2009;100:2580–7. polymers. In: Seymour RB, Stahl GA, editors. New York: Pergamon Press; 1982
[28] Mazza M, Catana DA, Vaca-Garcia C, Cecutti C. Influence of water on the dissolution ISBN 0-8-026337-2.
of cellulose in selected ionic liquids. Cellulose 2009;16:207–15. [57] Zhang S, Fu CF, Li FX, Yu JY, Gu LX. Direct preparation of a novel membrane from
[29] Glasser WG, Atalla RH, Blackwell J, Brown RM, Burchard W, French AD, et al. About unsubstituted cellulose in NaOH complex solution. Iran Polym J 2009;18:767–76.
the structure of cellulose: debating the Lindman hypothesis. Cellulose [58] Biganska O, Navard P. Morphology of cellulose objects regenerated from cellulose-N-
2012;19:589–98. methylmorpholine N-oxide-water solutions. Cellulose 2009;16:179–88.
[30] Kluefers P, Kunte T. Angew Chem Int Ed 2001;40:4210. [59] Isobe N, Kimura S, Wada M, Kuga S. Mechanism of cellulose gelation from aqueous
[31] Cai J, Zhang L, Liu SL, Liu YT, Xu XJ, Chen XM, et al. Dynamic self-assembly induced alkali-urea solution. Carbohydr Polym 2012;89:1298–300.
rapid dissolution of cellulose at low temperatures. Macromolecules 2008;41:9345–51. [60] Leipner H, Fischer S, Brendler E, Voigt W. Structural changes of cellulose dissolved in
[32] Kihlman M, Medronho BF, Romano AL, Germgard U, Lindman B. Cellulose Dissolution molten salt hydrates. Macromol Chem Phys 2000;201:2041–9.
in an Alkali Based Solvent: Influence of Additives and Pretreatments. J Braz Chem Soc [61] Liu HB, Sale KL, Simmons BA, Singh S. Molecular dynamics study of polysaccharides in
2013;24:295–303. binary solvent mixtures of an ionic liquid and water. J Phys Chem B 2011;115:10251–8.
[33] Yamashiki T, Kamide K, Okajima K, Kowsaka K, Matsui T, Fukase H. Some character- [62] Li R, Zhang LN, Xu M. Novel regenerated cellulose films prepared by coagulating
istic features of dilute aqueous alkali solutions of specific alkali concentration with water: structure and properties. Carbohydr Polym 2012;87:95–100.
(2.5mol l −1) which possess maximum solubility power against cellulose. Polym J [63] Zhang S, Li FX, Yu JY. Kinetics of cellulose regeneration from cellulose-NaOH/
1988;20:447–57. thiourea/urea/H2O system. Cellul Chem Technol 2011;45:593–604.
[34] Liu WQ, Budtova T, Navard P. Influence of ZnO on the properties of dilute and semi- [64] Hermans PH. Degree of lateral order in various rayons as deduced from x-ray
dilute cellulose-NaOH-water solutions. Cellulose 2011;18:911–20. measurements. J Polym Sci 1949;4:145–51.
[35] Cai J, Zhang LN, Chang CY, Cheng GZ, Chen XM, Chu B. Hydrogen-bond-induced in- [65] Hayashi J, Masuda S, Watanabe S. Plane lattice structure in amorphous region of
clusion complex in aqueous cellulose/LiOH/urea solution at low temperature. cellulose fibers. Nippon Kagaku Kaishi; 1974 948–54.
Chemphyschem 2007;8:1572–9. [66] Bergenstrahle M, Wohlert J, Himmel ME, Brady JW. Simulation studies of the insol-
[36] Cai J, Kimura S, Wada M, Kuga S, Zhang L. Cellulose aerogels from aqueous alkali ubility of cellulose. Carbohydr Res 2010;345:2060–6.
hydroxide-urea solution. Chemsuschem 2008;1:149–54. [67] Sjostrom J, Piculell L. Interactions between cationically modified hydroxyethyl cellu-
[37] Lu A, Liu YT, Zhang LN, Potthast A. Investigation on metastable solution of cellulose lose and oppositely charged surfactants studied by gel swelling experiments—effects
dissolved in NaOH/urea aqueous system at low temperature. J Phys Chem B of surfactant type, hydrophobic modification and added salt. Colloid Surf A
2011;115:12801–8. 2001;183:429–48.

Please cite this article as: Medronho B, Lindman B, Brief overview on cellulose dissolution/regeneration interactions and mechanisms, Adv Colloid
Interface Sci (2014), http://dx.doi.org/10.1016/j.cis.2014.05.004

You might also like