NC State
BioResources
Hubbe, M. A., Sjöstrand, B., Lestelius, M., Håkansson, H., Swerin, A., and Henriksson, G. (2024). “Swelling of cellulosic fibers in aqueous systems: A Review of chemical and mechanistic factors,” BioResources 19(3), Page numbers to be added.

Abstract

Factors affecting the swelling of cellulosic fibers are considered in this review.  Emphasis is placed on aqueous systems and papermaking fibers, but the review also considers cellulose solvent systems, nanocellulose research, and the behavior of cellulosic hydrogels.  The topic of swelling of cellulosic fibers ranges from effects of humid air, continuing through water immersion, and extends to hydrogels and the dissolution of cellulose, as well as some of its derivatives.  The degree of swelling of cellulose fibers can be understood as involving a balance between forces of expansion (especially osmotic pressure) vs. various restraining forces, some of which involve the detailed structure of layers within the fibril structure of the fibers.  The review also considers hornification and its effects related to swelling. The expansive forces are highly dependent on ionizable groups, pH, and the ionic strength of solution.  The restraining forces depend on the nature of lignin, cellulose, and their detailed structural arrangements.


Download PDF

Full Article

Swelling of Cellulosic Fibers in Aqueous Systems: A Review of Chemical and Mechanistic Factors

Martin A. Hubbe,a Björn Sjöstrand,b Magnus Lestelius,b Helena Håkansson,b Agne Swerin,b and Gunnar Henriksson b,c

Factors affecting the swelling of cellulosic fibers are considered in this review. Emphasis is placed on aqueous systems and papermaking fibers, but the review also considers cellulose solvent systems, nanocellulose research, and the behavior of cellulosic hydrogels. The topic of swelling of cellulosic fibers ranges from effects of humid air, continuing through water immersion, and extends to hydrogels and the dissolution of cellulose, as well as some of its derivatives. The degree of swelling of cellulose fibers can be understood as involving a balance between forces of expansion (especially osmotic pressure) vs. various restraining forces, some of which involve the detailed structure of layers within the fibril structure of the fibers. The review also considers hornification and its effects related to swelling. The expansive forces are highly dependent on ionizable groups, pH, and the ionic strength of solution. The restraining forces depend on the nature of lignin, cellulose, and their detailed structural arrangements.

DOI: 10.15376/biores.19.3.Hubbe

Keywords: Osmotic pressure; Donnan equilibrium; Dissociation; Hydrophilicity; Crystallinity; Water retention value

Contact information: a: Department of Forest Biomaterials, North Carolina State University, Campus Box 8005, Raleigh, NC 27696-8005, USA; b: Pro2BE, The research environment for Processes and Products for a circular Biobased Economy, Department of Engineering and Chemical Sciences, Karlstad University, SE-65188, Karlstad, Sweden; c: KTH Royal Institute of Technology, School of Engineering Sciences in Chemistry, Biotechnology and Health, Department of Fiber and Polymer Technology, Division of Wood Chemistry and Pulp Technology., SE-10044 Stockholm, Sweden; *Corresponding author: hubbe@ncsu.edu

INTRODUCTION

The swelling of cellulosic fibers can have profound effects on such applications as absorbent paper products (Botkova et al. 2013; Hubbe et al. 2013; Ma et al. 2015), nanocellulose products (Torstensen et al. 2022), and products made from dissolving grades of cellulosic fibers (Okugawa et al. 2020, 2021). Less dramatic but still very important effects are observed in other paper products (Larsson and Wågberg 2008; Figueiredo et al. 2016). During the manufacture of paper, there is often a positive correlation between the swelling of cellulosic fibers (especially for kraft fibers) and the development of paper strength (Jayme and Büttel 1968). This phenomenon is mainly connected to the increased fiber flexibility that is associated with the swelled structure and consequently the increased number of fiber-fiber bonds in the fiber network. On the other hand, with increasing swelling of cellulosic fibers, the amount of water carried past the press section of a paper machine is increased (Hubbe and Heitmann 2007; Hubbe et al. 2020). The need to evaporate such water can slow down the rate of production and increase the net expenditure of energy.

Background: Structural Character of Cellulose and Matrix Polymers

The swelling of cellulosic fibers and their networks is normally quantified as liquid (water) mass uptake per mass of fiber, rather than quantified by a length or volume increase. A main reason is that several important steps in fiber swelling do not involve geometric or dimensional changes. To be able to discuss swelling behavior of cellulose and pulp fibers, one needs to take account of the structure of not only cellulose, but also cell wall matrix polysaccharides, i.e., hemicelluloses and pectins. Cellulose has a simple primary structure consisting of anhydrous β-glucopyranosides connected with 1,4 glycosidic bonds, and a degree of polymerization that can reach several thousands (Heinze 2015). The structure of this monomer can be described as being like a discus, in which there are hydroxyl groups peripherally oriented, whereas there are apolar groups facing the tops and bottoms (Fig. 1) (Seeberger 2015).

Fig. 1. Representation of the repeat unit in a cellulose chain: (a) β-glucopyranoside, (b) 1,4 glycosidic bonds, depicting also intra-chain hydrogen bonds

The glycosidic bond in the cellulose chain allows the possibility to form two stabilizing hydrogen bonds on each side, and these stabilize the structure into a two-fold helix, i.e., a flat and almost straight structure, with hydroxyls exposed to the sides and a hydrophobic character facing upwards and downwards (Kovalenko 2010). In natural cellulose, i.e. cellulose I, chains with the same direction, i.e., having the reducing endo towards the same direction, are lined up beside each other. They are connected with multiple hydrogen bonds, forming sheets in a secondary structure. These sheets are located on the top of each other and held together by van der Waals bonds in a tertiary structure. For steric reasons, one residue is not located exactly over another but displaced both in chain direction and side direction. Thus, cellulose is forming a crystalline structure (Glasser et al. 2012), which is important for the development of superior mechanical properties of the polysaccharide (Ioelovich 2016).

The exact shape of the crystal, the quaternary structure, is a matter of debate; cellulose is synthesized by an enzyme complex located in the cell membrane, and the chains are present in aggregates, meaning that several chains are synthesized in parallel, thereafter forming extended crystals, which together constitute microfibrils (Saxena et al. 2005). The size, i.e., the number of cellulose chains, and maybe also the shape of the microfibril, therefore depend on the size and shape of the aggregate of cellulose synthases. In some non-plant sources of cellulose, i.e., bacteria, tunicates, and certain algae, the microfibrils can be very large and sometimes have a rather flat structure (Moon et al. 2011). In higher plants, the microfibrils are believed to be rather small; recent results suggest 18 chains (Jarvis 2018), but earlier both 28 and 36 chains have been suggested (Cosgrove 2014).

What then is the cross-sectional shape of a fibril in higher plants? The answer is not known with certainty, but a common suggestion is the honeycomb shapes presented in Fig. 2. Although other shapes also have been suggested, it is inevitable that the microfibril will have both hydrophobic and hydrophilic surfaces exposed (Mazeau 2011). These two surfaces have distinctly different properties, and it seems as if the hydrophobic side is attractive for adsorption of hemicelluloses, as will be further discussed below (Heinonen et al. 2022; Kong et al. 2022). Proteins that specifically bind to cellulose, such as the cellulose binding module of fungal aerobic cellulases, seem to bind to the hydrophobic surface by forming sandwich structures between the hydrophobic top of glucopyranside residues and aromatic amino acids, i.e. tyrosine and tryptophan (Arola and Linder 2016). Issues related to the differing affinities of different crystal planes of cellulose will be considered in more depth later in the article.

Fig. 2. Likely cross-sectional shapes that have been proposed for the microfibrils of cellulose within higher plants. a) Honeycomb (diamond) model. This model is favored by many scholars; note that both hydrophilic and hydrophobic surfaces are exposed. b) An alternative “double fibril” model. Here the hydrophobic surfaces are smaller, but instead “pockets” are created.

The contrasting hypothetical cross-sectional shapes of the initial cellulose microfibrils that form during biosynthesis can be expected to have implications related to how the units self-assemble into larger structures at later stages of biosynthesis. A recent review of the fundamentals of self-assembly of lignocellulosic materials showed how idealized shapes, not unlike those shown in Fig. 2, might be expected to govern patterns of self-assembly (Hubbe et al. 2023). For instance, one might speculate that the “pocket” depicted in Part B of the figure might provide some kind of evolutionary advantage in the ability of plants to form strong cell walls. However, many fundamental questions remain unresolved concerning the mechanisms of cell wall formation in plants.

The cellulose microfibril should not be understood as being a stiff, static, and perfect crystal; rather, cellulose consists of a mixture of highly crystalline and more unordered cellulose domains. The unordered cellulose is partly located on the surface of the fibril. In the presence of water or similar liquids, there is probably some movability of the superficial cellulose chains on the fibril – but there are indications that there also are segments of the fibrils in which the unordered structure goes deeper into the fibril than the single layer indicated in Fig. 3 (Brestkin and Frenkel 1971; Nishiyama et al. 2003; Fernandez et al. 2011). The unordered cellulose is generally more reactive towards hydrolysis with acid or cellulolytic enzymes (cellulases) and probably also for chemical derivatization (Wang et al. 2014a; Ioelovich 2021). Most likely, an unordered segment in a fibril might allow the fibril to bend more easily, and the movability of the superficial cellulose chains might play an important role for interaction with water or another component. A honeycomb organization of the microfibril with moving superficial cellulose chains will give the microfibril a cylindrical appearance – something that seems to be supported by electron microscopic pictures of cell walls.

Fig. 3. The central part of the microfibrils might in general be rather crystalline, whereas surficial cellulose might be more flexible and unordered. There are most likely also short segments where the unordered cellulose goes deeper (Brestkin and Frenkel 1971; Fernandes et al. 2011).

With some exceptions, such as the seed hairs of cotton, cellulose does not appear as a single polysaccharide component in higher plants; the plant cell wall generally contains other polymeric carbohydrates, which are often called matrix polysaccharides. These are divided into two groups, the hemicelluloses and the pectins (Scheller and Ulvskov 2010). Hemicelluloses have a main chain in which the monosaccharide residues are connected with β-1,4 glycosidic bonds. The hydroxyl group on the C4 is peripherally located (as in glucose, Fig. 1). This allows the polysaccharide to adapt to the same two-fold structure as in cellulose, although other conformations also are possible. Examples are glucomannans, arabinoxylans, and xyloglucans. Pectins are cell wall polysaccharides that cannot adapt to such a two-fold helix. Examples include β-1,4 galactans, -1,5 galacturonans, and rhamnogalacturonans (Alkorta et al. 1998; Ebringerová et al. 2005; Scheller and Ulvskov 2010).

Recent data suggest that both hemicelluloses and pectin can interact, such that they bind to cellulose surfaces, but that they do so in different ways; hemicelluloses can align themselves both parallel and antiparallel with cellulose chains in the microfibrils. In a sense, they may appear to “crystalize” onto the cellulose domains. By contrast, pectins bind in more nonspecific way. Not surprisingly, the hemicelluloses appear to be more firmly attached to the cellulose microfibrils. Generally, matrix polysaccharides appear to prefer to bind to the hydrophobic surfaces of fibrils (Fig. 2) (Heinonen et al. 2022; Jarvis 2023).

The composition of matrix polysaccharides differs greatly between different phylogenic groups, but also between different cell wall layers. These differences start with the secondary wall, which is dominant in terms of both mass and volume (Harris 2006). The most studied of these are the woody eudicolyledons (hardwoods, i.e. “eucots”), monocotyledons, and conifers (softwoods, i.e. the dominating phylum of gymnosperms) (Daly et al. 2001). In woody eucots, xylans are the most common secondary cell wall polysaccharides, but glucomannans also are present in significant amounts (Salmén 2022). In monocots, xylans dominate almost totally, whereas in conifers, galactoglucomannans are the most common, followed by rather large amounts of arabinoxylans (Moreira and Filho 2008; Peng and She 2014). The primary cell walls have a more complex composition, and in addition to xylans, also xyloglucans and pectins (rhamnogalacturonan and polygalacturonic acid) are common (Hayashi 1989; Brett et al. 2005; Caffall and Mohnen 2009). The latter polysaccharides can also be present in the middle lamella. Berglund et al. (2020) presented evidence that different hemicelluloses can influence the elastic modulus and stretch to breakage of cellulose-based films in contrasting ways.

In summary, the cellulose itself probably mainly exposes surfaces rich in hydroxyls, thereby opening up for the ability to interact with water and other molecules, often leading to hydrogen bonding. In addition to this, there are possibilities for hydrophobic interactions and van der Waals interactions on the hydrophobic sides of a fibril (Fig. 2). A more unordered cellulose, when it is in a swollen state, will is likely to have proportionally more hydrophobic surfaces exposed. Charges, on the other hand, do not occur on pure natural cellulose (which is a homoglucan), so salt bridges do not occur directly between natural cellulose molecules (Kontturi et al. 2006). Anionic structures, i.e., carboxylic acids, hypothetically may be induced on the cellulose polymer, by technical processes, such as bleaching. The matrix polysaccharides (the hemicelluloses) will most likely be forming layers around microfibrils, connecting them, or embedded in them, adding more of unordered structures with richer possibilities for hydrophobic interactions and van der Waals forces. Furthermore, certain matrix polysaccharides, i.e., xylans, polygalacturonic acid, and rhamnogalacturonan, contain anionic carboxylic acids, thereby making plant fibers negatively charged. Xylans have been suggested to act as a kind of dispersant for cellulose microfibrils by adding charge to them by associating with the surfaces (Berglund et al. 2020). A similar role has been observed for glycerol (Moser et al. 2018b). Especially the cationic groups in pectins often form salt bridges to calcium ions, thereby crosslinking chains by means of “eggbox structures” (Fig. 4) (Cao et al. 2020). These structures function as crosslinkers to polysaccharide chains and have importance for the rigidity of pectin-rich tissues, such as middle lamella of the phloem in certain herbs, e.g. flax and hemp. Removal of calcium from pectin leads to a swelling and weakening of pectin rich tissues, and it increases their accessibility for enzymatic degradation; this approach has been considered for chemical retting of bast fibers (Henriksson et al. 1997). Addition of calcium salts may on the other hand decrease swelling (Thakur et al. 1967). In line with this, Eucalyptus wood with high content of calcium is more recalcitrant to kraft pulping (Vegunta et al. 2022).

Fig. 4. Complexes formed between carboxylic anions and calcium(II) ions are known as “egg boxes”. They are believed to play a central role in the rigidity of pectin-rich tissues (Heinze 2016). a) Chemical structure of an eggbox, i.e., complex between two uronic acids and a calcium ion, thereby crosslinking two polysaccharide chains; b) Schematic presentation of the role of egg boxes in pectin structures; the calcium ions crosslink the “smooth” regions rich in galacturonic acids (Cao et al. 2020)

Steps in Swelling

The process of swelling of cellulosic fibers can be envisioned as a continuum of overlapping steps. To begin, equilibration of dry cellulosic fibers, paper products, or other such materials with increasing levels of relative humidity results in increasing moisture uptake and slight changes in macroscopic dimensions (Larsson and Wågberg 2008; Gamstedt et al. 2015; Joffre et al. 2016). Greater proportional effects of relative humidity on dimensions have been found when examining the thickness dimension of thin films of nanocellulose (Rehfeldt and Tanaka 2003; Shrestha et al. 2017; Torstensen et al. 2018). Wetting by pure water results in much larger mass uptake and swelling of dimensions (Scallan 1983; Solhi et al. 2023). A key focus of this review article will be on ways in which the pH and ion contents of such solutions affect the degree of swelling and how such effects can be understood based on chemistry, physics, and colloid science. Further aspects of the swelling mechanisms are revealed by considering evidence from more extreme situations, including the behavior of cellulose-based hydrogels (Ganji et al. 2010; Chang and Zhang 2011; Ma et al. 2015) and solvent systems for cellulose (Budtova and Navard 2016). For instance, some of the greatest observed degrees of swelling of cellulosic material are associated with the last steps before dissolution in the presence of specialized solvent systems (Cuissinat and Navard 2006; Zhang et al. 2013; Budtova and Navard 2016).

The Riddle of Cellulose’s Insolubility in Water

In addition to the practical importance of swelling, as already described, the results of investigations into the swelling of cellulose fibers also can provide evidence into fundamentals of how the cellulosic material is being held together. As will be described further in this article, hydrogen bonding plays a central role not only in the bonding within cellulosic fibers, but also with respect to their interactions with water (Medronho et al. 2012; Norgren et al. 2023; Sjöstrand et al. 2023). If that were the only type of bonding, however, then it would be hard to explain the insolubility of cellulose in most aqueous conditions (Bergenstråhle et al. 2010; Budtova and Navard 2016). As been discussed above, cellulose chains and microfibrils display hydrophobic surfaces at specific planes, and these might interact with each other by van der Waals forces. When water is present, the cellulose is kept together, in part, by hydrophobic interactions, which involves an entropic effect of the solvent (Ghosh et al. 2002; Glasser et al. 2012). Thus, this article will consider effects such as the orientation of cellulose macromolecular chains, which have differing affinity for water on different sides. This amphiphilic character is demonstrated by the differing wettability characteristics of the different faces of cellulose crystals (Yamane et al. 2006). In addition, the contrasting microfibril angles in different sublayers of a cellulosic fiber, as will be shown later, sometimes tend to restrain the swelling of other layers within the same fibers.

Sections of this Article and Recognition of Earlier Review Articles

This review article is organized into the major sections of quantification of swelling, factors affecting the swelling of cellulosic matter, theoretical aspects of swelling of cellulosic matter, applications of swelling of cellulosic material, and some closing comments. From a developmental and materials science perspective, the articles to be cited point to two main strategies that have been used to modify the swelling degree of cellulose-based materials intended for different applications. Chemical approaches, including the forming or attachment of different ionizable functional groups on cellulosic surface, can be used to increase the osmotic pressure that tends to increase swelling, subject to various aqueous conditions. Chemical delignification, along with physical treatments such as mechanical fibrillation (i.e. refining), can be used to alter or decrease physical restraints to the swelling of the same materials. However, as will be discussed, the restraining tendencies of cellulosic structures are also highly dependent on any history of drying, which can decrease the swelling ability, i.e. effects of hornification (Fernandes Diniz et al. 2004; Salmén and Stevanic 2018; Sjöstrand et al. 2023; Sellman et al. 2023).

Aspects of the swelling of cellulosic fibers have been considered in some earlier review articles, and these are listed in Table 1. Based on the items in this list, there appears to have been an acceleration of interest within the last two years.

Table 1. Earlier Review Articles Dealing with Aspects of Swelling of Cellulosic Fibers and Related Materials

QUANTIFICATION OF SWELLING

The swelling of cellulosic fibers and networks is often defined as the amount of liquid (often water) mass uptake per mass of fiber. Another way to define swelling is based on dimensional changes. However, such dimensional changes may be insignificant or hard to quantify in many situations of interest. Thus, in the literature, various different definitions and evaluations methods have been employed.

Commonly used methods to quantify the swelling of cellulosic fibers and related materials can be placed into the categories of water retention value (and related tests), methods involving measurements of the concentrations of soluble polymers in solution, methods based on observed changes in dimensions of objects, methods based on permeability, and a variety of lesser-used approaches (Gallay 1950). These are described in the subsections below. As a general summary, each of the existing methods to be considered for evaluation of swelling involves areas of uncertainty and approximation. Methods that rely upon the dimensions of swollen cellulosic material typically suffer from uncertainty in defining the boundary between the material and the surrounding aqueous (or other) medium. Methods that depend on the sizes of selected probe molecules tend to be blind to the presence of some pores that may be larger than the probe. And some, such as the WRV tests to be described next, can be regarded as practical, but not firmly grounded in theory.

Water Content Assessment

Water retention value

The water retention value (WRV) test (Welo et al. 1952; Marsh et al. 1953; Jayme and Büttel 1964, 1968; Cheng et al. 2010), which involves centrifugation of a damp plug of fibers, is specified in standard methods (SCAN-C 62:00, 2000; TAPPI UM 256 1981; ISO 23714:2014). The basic assumption is that when a specified centrifugal acceleration is applied to the damp plug of fibers, the acceleration will remove water that was either between the fibers or within the fiber lumens. The idea is that the effective forces acting on the fibers will squeeze each of the fibers, such that any water contained within the lumen space will come out via the pit openings. An analogy can be made to the squeezing of a tube of toothpaste. By contrast, it is assumed that any water contained within the mesopores or micropores within the cell walls of fibers will be prevented from escaping by sufficiently large capillary forces (Welo et al. 1952). Different fiber starting materials will of course showcase varying fiber properties, and the remaining water will be located differently based on fiber morphology.

As a general rule, WRV tests have a reputation of providing useful comparisons between chemically delignified pulp samples that have been refined to different levels. However, high levels of external fibrillation can contribute to the WRV values in a way that does not fit with the usual definition of swelling. For example, a well-fibrillated thermomechanical pulp (TMP) fiber may have a WRV value that suggests about twice the level of swelling in comparison to the results of soluble exclusions tests (see later discussion). In addition, TMP tends to maintain its WRV over cycles of papermaking and repulping (Alanko et al. 1995). Because TMP has a composition similar to wood, its cell walls lack the porosity and swelling ability of kraft fibers, especially after the latter have been refined.

A related test method of water removal is drainage, which can be measured in Degrees Schopper Riegler (°SR) or Canadian Standard Freeness (CSF). Drainage is more related to liquid water passing between fibers but has been shown to correlate linearly with WRV. However, it is important to note that slopes of linear correlations are different for different pulp compositions (Sjöstrand et al. 2019).

To carry out a WRV test, a well-defined amount of pulp fibers is formed into a plug, which is then centrifuged under specified conditions of centrifugal force, time, and temperature. One weighs the material immediately after centrifugation. The procedure is done in such a way as to allow water released from the fibers to pass through a screen and not return. A second weight is determined after the same plug of fibers has been completely dried, preferably according to ISO 638 (2008) or a similar method.

Different results can be obtained from WRV tests, depending on the level of centrifugal acceleration (Cheng et al. 2010). The WRV values tend to decrease with increasing centrifugal acceleration and time of spinning. As described in the cited article, smaller particles of cellulose gave rise to larger WRV values, which is consistent with the presence of generally smaller capillary opening within the material. Based on this principle, WRV has been employed as a way to characterize the degree of fibrillation of highly fibrillated cellulose materials, i.e. micro- or nanocellulose (Gu et al. 2018). Effects of carboxymethylation (Laine et al. 2003) and TEMPO-oxidation (Maloney 2015) of cellulosic fibers also have been evaluated with WRV.

Although the WRV test has been used for many years as a practical way to estimate the amount of water held within the cells walls of ordinary cellulosic fibers, especially when they are being used to manufacture typical grades of paper, there is evidence that some other categories of water can be involved, especially when polymeric chemical additives are being employed. For example, if one were to assume that the WRV test senses only water present in mesopores, then one would expect that relatively small cationic polymers, having easy access into such pores, would have a big effect on the WRV results. In principle, the small cationic polymers would be expected to diffuse into tiny pores in the cell walls and neutralize surface charges, leading to pore contraction due to the decreased osmotic pressure. However, the opposite was found by Swerin et al. (1990), and Ström and Kunnas (1991). They observed that a high-molecular-mass cationic polymer product had a much bigger effect on WRV than a low-mass version. It has been argued that such results are consistent with a charged patch type of interactions among external fibrils and polyelectrolytes at the fiber surfaces (Hubbe and Panczyk 2007b; Hubbe et al. 2008). Alternatively, the same effects might be attributed to complexation among oppositely charged polyelectrolytes, where the cellulosic fibrils are regarded as if they were anionic polyelectrolyte strands (Hubbe 2019). In either case, the WRV results appear to be sensitive to details concerning such charged polymer additives.

Some modified versions of the WRV test have been demonstrated, seeking to broaden the range of information that can be obtained. For example, Hubbe and Panczyk (2007a) developed a modified WRV method (MWRV) in which a mass was placed on top of the damp plug of fibers before the start of centrifugation. Such a modification makes the test results somewhat more relevant to such situations as water removal in a wet-pressing operation. Notably, however, variations in pH and salt concentration were not found to have major effects on the results of the MWRV tests. Rather, increased refining increased the MWRV results, whereas drying history of the fibers decreased the MWRV values.

Some researchers have obtained WRV values when evaluating the properties of nanocellulose suspensions, but the justification and meaning of such tests can be questioned. To carry out such tests, a membrane is used to support the cellulosic material during centrifugation (Cheng et al. 2010; Gu et al. 2018). This approach makes it possible to evaluate nanocellulose materials having small enough particle size to pass through the usual supports used in WRV tests. Since nanocellulose suspensions do not have any lumen spaces or inter-fiber spaces, one cannot claim that the applied centrifugation is meant to remove water from such spaces. Rather, it seems likely that the WRV results from testing of nanocellulose suspensions will be related to surface area and fineness of the material, which are expected to contribute to both capillary forces and resistance to flow.

Solute Concentration Tests

Another basic way to gain information about the amounts of water that are contained within the mesopore structure of cellulose fiber cell walls involves the evaluation of concentrations of probe macromolecules in solution (Stone and Scallan 1966, 1967, 1968; Scallan and Carles 1972; Alince 1991). The idea is based on an assumption that a water-soluble polymer of a certain molecular mass will be prevented from entering a pore if it is smaller than a critical size, perhaps related to the radius of gyration of that macromolecule. Three ways of carrying out such tests are possible, as described below.

Fiber saturation point

To determine the fiber saturation points of cellulosic materials in aqueous suspension, one selects a water-soluble polymer having a lack of attraction to cellulosic surfaces and a suitably high and narrow distribution of molecular mass such that it can be assumed that little of the macromolecule will be present in the mesopores of a typical chemical pulp (kraft, sulfite, etc.) specimen (Stone and Scallan 1966, 1967; Alince 1991). Dextran molecules having molecular mass of about one million Daltons or greater have been used for such tests. Rather than providing a physical barrier to entrance of the probe polymers, one relies upon the fact that the probability that such a polymer will diffuse into a given space depends strongly on whether or not the freedom of motion of polymer segments will become constrained, thus affecting the entropy content and the free energy (Alince 2002). The concentration of the probe polymer in the bulk of solution will tend to be increased, compared to what can be calculated based on the known added quantity. The calculation involves the total amount of solution present, including solution that happens to be within mesopores and micropores of the cell walls of fibers. One then calculates the volume that is associated with the pores too small to permit access. As noted by Scallan and Carles (1972), the obtained fiber saturation value results are often in reasonable agreement with WRV test results, at least when testing ordinary papermaking fibers within typical ranges of mechanical refining.

Inverse size-exclusion chromatography

Size exclusion chromatograph (SEC) has emerged as an effective and practical way to evaluate the molecular mass distributions of polymers in solution (Burgess 2018). The method is based on the principle that the time required for a given macromolecule to elute through a packed column will depend on whether or not it is small enough to spend time within the pores of a selected packing material. Very small probe molecules will require the longest elution times, since they will be spending more time, on average, within parts of the liquid that are not available for larger macromolecules. Inverse size exclusion chromatography (ISEC) measurements are similar, except that one employs a known probe macromolecule, having a narrow molecular mass distribution, and the quantity to be determined is the pore size or pore size distribution of a solid material, e.g., a swollen cellulosic substance (Yao and Lenhoff 2004). The ISEC method was employed, using a series of different molecular mass fractions, by Berthold and Salmén (1997a) to help understand how the process of kraft pulping and bleaching affect the pore size distribution. In general, it was found that there was a substantial increase in the volume of mesopores (i.e., pores from 2 to 50 nm in diameter) following chemical pulping, with even greater porosity after both pulping and bleaching. Follow-up work showed how such distributions were shifted due to partial closure of intermediate-sized pores (mainly mesopores) in the course of drying and rewetting (Berthold and Salmén 1997b).

Polyelectrolyte adsorption

Because cellulosic materials typically have a negative charge – due to the presence of axabinoxylan or pectin or by chemical modification of cellulose (natural cellulose is, as discussed above, uncharged) – cationic polyelectrolytes will be expected to interact strongly with them. It has been proposed to use cationic polyelectrolytes of known molecular mass to probe the pore structures of cellulosic fibers (Alince and van de Ven 1997; Alince 2002). By employing a highly branched polymer, the cited work reached the conclusion that mesopores within the swollen cell walls of bleached kraft fibers were relatively uniform and about 80 nm or larger in diameter. However, much smaller sizes have been determined by other methods, such as NMR (Li et al. 1993). Studies have shown that factors affecting permeation of cationic polyelectrolytes from aqueous into porous materials are complex (Wu et al. 2009). As described in the cited review article, cationic polymers at first may be drawn into small pores in a snake-like fashion; however, the progress of cationic polymers into such pores is likely to be strongly affected by kinetic factors, including essentially trapped non-equilibrium conditions. Further information about cellulose porosity was obtained indirectly by sensing electrokinetic effects that originate from those areas of bleached kraft fibers (Hubbe et al. 2007a).

Table 2. Methods to Evaluate Swelling Based on Dimensional Changes

Methods Based on Dimensional Measurements

Reported swelling methods involving dimensional measurements of cellulose-based materials have tended to be diverse, rather than following a standard. It appears that different kinds of specimens, or possibly different orientations of those specimens, will require different procedures. Table 2 summarizes such work, with attention to the nature of the specimens and the methods. Note that some of these methods were designed in such a way as to capture rapid changes in dimensions upon introduction of changes in humidity or immersion (Qing et al. 2013; Jablonsky et al. 2014; Shrestha et al. 2017). In addition, it has been possible to capture dramatic and to a large extent reversible changes in dimension by wetting and redrying by suitable methods (Kontturi et al. 2011).

Permeability

Though the degree of correlation between permeability and porosity (or fractional volume content of pores in a material) needs to be established in each case, permeability is sometimes used as a proxy for porosity measurements. In paper specimens, there is a general expectation that air permeability will decrease with increasing apparent density. However, the effect is not linear. This is shown in Fig. 5, which for the first time plots data reported by Vänskä et al. (2016). The cited authors varied the lignin content as well as the extent of mechanical refining. As shown in the figure, the logarithm of the time required for permeation of 100 cm3 of air was found to be roughly proportional to apparent density. However, all of the results shown in the figure for the highest Gurley seconds values were beyond the upper limit of measurement. The densest sheets, representing high levels of refining, tended to act as near-perfect seals, suggesting molecular contact between adjacent fibers in the structure. In other cases, the correlation between permeability and density may be poor due to detailed differences in pore structure (Hantel et al. 2017). Nonlinear and complex relationships between paper permeability and density are also suggested by studies of the relative ease of removal of water during the paper forming process (Hubbe et al. 2020). The cited review article showed, for instance, that densified surface layers may have a dominant effect on permeation behavior.

Fig. 5. Plot prepared here for the first time from the data of Hantel et al. (2017), showing the logarithm of air permeation times for paper prepared with different delignification degree and mechanical refining levels, leading to the apparent density values shown for the resulting paper

Surface Area and Porosity Using Specialized Drying Approaches

One of the most trusted approaches to determining the surface area and pore sizes of materials is by adsorption of very cold nitrogen or argon gases at low pressures (Bardestani et al. 2019). When the goal is to study the pore structures of cellulosic materials, one of the key challenges is the fact that such tests require the usage of completely dry samples. It is well known, however, that substantial and partly irreversible closure of pores (i.e. hornification) occurs when cellulosic materials are dried using conventional evaporation, especially in cases where the lignin has been removed (Fernandez Diniz et al. 2004; Hubbe et al. 2007b). Geffertova et al. (2013) found that repeated cycles of paper forming, drying, and reslurrying of the pulp in preparation for the next cycle without additional refining tended to increase the air permeability of sheets. This is consistent with the increasing stiffness and lack of compliance of hornified fibers. To minimize such effects during determination of pore sizes in wet specimens, it is recommended to replace the water in the specimens with another solvent and then employ critical point drying (Kang et al. 2018). In other words, evaporation takes place at the location within a phase diagram where solid, liquid, and gas phases meet each other. Such approaches can avoid development of high surface tension forces and extensive hydrogen bonding in the course of drying. Once the specimens have been dried, then one can use the Brunauer-Emmett-Teller (BET, 1938) method to obtain the surface area, based on gas adsorption isotherms (Sing 2001; Bardestani et al. 2019). Related methods can provide estimates of pore-size distributions (Barrett et al. 1951; Bardestani et al. 2019), though such determinations require simplifying assumptions (Groen et al. 2003).

These BET-related methods require that the sample is dried. Even when using specialized drying methods, there still will be some risk of pore collapse, which will influence the result. Therefore, there is interest in methods for measuring the cellulose surface area without drying. One such method is based on adsorption of the colorants methylene blue and Congo red (Inglesby and Zeronian 1996). This method requires a high content of sulphate ions to make the adsorption possible. This is problematic, since it can lead to aggregation of fibers and collapse of pores. Moser et al. (2018a) developed a method based on adsorption of the polysaccharide xyloglucan on the cellulose, which does not require increased ionic strength. The content of non-bound xyloglucan can easily be detected, since it forms a strongly colored complex with iodine.

FACTORS AFFECTING THE SWELLING OF CELLULOSIC MATTER

The emphasis of this section is to consider evidence about what parameters affect swelling of cellulosic fibers. Though some theoretical aspects may be unavoidable, in light of such emphasis, the goal will be to consider articles that employed an empirical approach, rather than to focus on theories, which will be covered in the subsequent main section.

Aspects related to experimental design and selection of conditions need to be considered carefully when attempting to understand published work related to the swelling of cellulosic fibers. With many exceptions, typical experiments related to the swelling of cellulosic materials have involved one-at-a-time variation of certain parameters e.g., pH. The conditions that are held constant, in such experiments, can be numerous and quite specific to the interests of the researchers. Those conditions may not necessarily match those employed by other researchers. As will be illustrated by some of the cases considered in this article, some findings will appear to have general validity, whereas other findings may be limited to special circumstances.

In general, as will be described in this section, there has been great progress in understanding different aspects that contribute to swelling behavior (Benselfelt et al. 2023; Sellman et al. 2023; Sjöstrand et al. 2023; Solhi et al. 2023). Factors that influence the extent of swelling include not only the chemical composition and aqueous conditions, but also the details of the layered structures that make up a cellulosic fiber.

Fiber Structural Details Resulting from its Biosynthesis

Many key attributes affecting the swelling behavior of cellulosic fibers have their origin in the nanostructures and microstructures that are established during biosynthesis. For a full understanding of those details, readers are referred to other sources (Brown and Saxena 2000; McNamara et al. 2015; Manan et al. 2022).

Some important aspects, relative to swelling behavior, can be emphasized. First, it has been shown that cellulose chains during their biosynthesis are essentially extruded towards the outside of cell membranes in parallel, essentially continuous manner (Brown and Saxena 2000; Tobias et al. 2020). As mentioned earlier, they almost immediately join together in nascent crystalline form in groups of about six, thus becoming microfibrils. Although the term “microfibrils” does not contain the term “nano,” it has become well established in the literature (Doblin et al. 2002) as a descriptor for these features that have nano-sized thicknesses. The process is depicted schematically in Fig. 6, part A. The nanofibrils may then join in parallel with other such groups, thus becoming larger fibrillar structures. In close succession, hemicellulose is also formed, and then lignin, thus surrounding and connecting the cellulose nanofibrils or microfibrils. In that manner, nature forms a nanocomposite structure in which the relatively stiff cellulose fibrillar elements are embedded in a double matrix of hemicellulose and lignin. Figure 6, part B illustrates the diffusion of monolignol compounds into the nascent cellulosic substance during the biosynthesis process.

Fig. 6. Schematic diagram of key biosynthetic events in the formation of lignocellulosic fibers. A: cellulose biosynthesis; B: lignin biosynthesis occurring within spaces of the just-formed cellulose nanofibrils or microfibrils.

Another key aspect that will affect the swelling behavior of cellulosic fibers, related to their biosynthesis, is the layered nature of the cell walls, and in particular the microfibril angles within the cell walls. Figure 7 depicts a typical cellulosic fiber, as it might be present in a wood specimen. As shown, each fiber has a thin primary (P) layer, and within that are three secondary (S) sublayers, of which the S2 sublayer is by far the thickest, such that it tends to dominate the structural aspects of fibers. The microfibril angle within the S2 sublayer often is within the range of 2 to 25 degrees, i.e., close to being aligned with the fiber and with the axis of the plant (Barnett and Bonham 2004; Donaldson 2008). Substantially higher values of the S2 sublayer fibril angle are found in such wood species as juniper (Hanninen et al. 2012); such wood species tend to be much more flexible and resistant to breakage, but less able to grow high.

Fig. 7. Sketch of lignocellulosic fiber, including layered structure, typical microfibril angles, and lignin distribution

The thinner S1 sublayer wraps itself around the S2 sublayer, such that it tends to restrict outward swelling of the S2 sublayer (Barnett and Bonham 2004). On the inside of the structure, adjacent to the lumen space, the S3 sublayer is likely oriented such that it tends to prevent inward collapse of the S2 sublayer. Between adjacent fibers, the “middle lamella” region is rich in lignin, thus providing a stiff, decay-resistant matrix to the composite structure. Studies suggest that at least part of the hemicellulose within a typical cellulosic fiber acts as a covalently bonded connector between lignin and cellulose (Lawoko et al. 2003).

Disassembly of Woody Material

In view of the theme of restraining factors and forces that tend to oppose the swelling of cellulosic fibers, it is important to keep in mind some of the widely employed technologies that are being used to prepare such fibers for papermaking, nanocellulose, and other applications. Many such processes can be regarded as disassembly processes, whereby either some components are removed or the fibers are mechanically separated from each other. Such treatments may weaken the ability of structures within the fiber to restrain swelling tendencies.

Delignification

Terms including “pulping” and “chemical pulping” are used to denote the widely employed methods of removing lignin (often with partial removal of hemicellulose) as part of the preparation of cellulosic fibers for papermaking and other applications (Mboowa 2021). It has been shown that removal of lignin from a cellulosic fiber, still in its undried state, leaves behind a structure having substantial mesoporous character (Stone and Scallan 1967; Berthold and Salmén 1997a,b). In other words, there are pores having diameters in the range of about 2 to 50 nm. Berthold and Salmén (1997a), who employed inverse size exclusion chromatography, found that delignification resulted in development of about 1.1 to 1.2 mL/g of new pore volume for water-swollen never-dried bleached and unbleached kraft pulps. Confirmatory evidence of increased pore volume after chemical pulping has been obtained by WRV testing (Jayme and Büttel 1968; Laivins and Scallan 1993). For example, Carlsson et al. (1983) showed that the WRV, a measure of swelling, increased from an initial value near to 115% (mass water per mass solids) up to the range 160 to about 250 %, depending on the pH. Scallan and Tigerström (1992) showed that such results could be interpreted as being due to a decreased elastic modulus of the cellulosic fiber material with the removal of the lignin, thus decreasing the ability of the fiber structures to resist swelling.

Presumably this development of mesopore spaces in wood fiber cell wall as a result of chemical pulping could be attributed to the presence of lignin-rich nano-domains having such dimensions in the original wood. Such domains are illustrated in Fig. 8, which is a redrawn version inspired by Fengel and Wegener (1989). Note that lignin is envisioned in the sketch as surrounding a microfibril (i.e., an assemblage of nanofibrils), the center portion being mostly cellulose, and with hemicellulose acting as an intermediary between the cellulose and the lignin. As mentioned in the Introduction, it should be understood that pectins also are expected to play a role (Alkorta et al. 1998; Ebringerová et al. 2005; Scheller and Ulvskov 2010). The idea of lignin occupying nano-sized domains is supported by the relatively large size of the lignin macromolecules that have been isolated by alkaline pulping methods (Gupta and Goring 1960).

Fig. 8. Depicture of woody material at the nano-scale, whereby cellulose and lignin tend to occupy different nano-domains, and the hemicellulose tends to be located between the cellulose and lignin

Lignin-carbohydrate complexes in intact wood

Another important feature that is shown in Fig. 8 is a type of linkage that can be expected to play a contributing role in inhibiting swelling of intact wood specimens. Note the label “LP-Linkage” that appears at the lower right of the figure. This label refers to covalent attachments between hemicellulose and lignin, i.e., lignin-carbohydrate complexes (LCCs) (Lawoko et al. 2003). The presence of such bonds helps to explain how it is possible for the hemicelluloses, which are the most hydrophilic main component of wood, to remain securely adherent, at a molecular level, to lignin, which is much more hydrophobic. Thus, wood retains its inherent strength and integrity, even when it is soaked during rainy seasons or immersion in water. In intact wood, lignin forms connections with different matrix polysaccharide molecules in a way that crosslinks the structure. This prevents wood fibers from extensive swelling in water (Henriksson 2017). Such limitations to swelling can be regarded as an attribute of “woody” matter, i.e. secondary xylem. During chemical pulping, the lignin-polysaccharide networks are in large part dissolved (Lawoko et al. 2005).

The moisture content of intact wood specimens has been shown to rise in approximately linear fashion with increasing relative humidity (RH) from about zero to about 0.3 mass fraction at about 95% RH (Fredriksson 2019). Further exposure of the material to liquid water gave rise to a dramatic rise to about 2.5 mass units of water per mass unit of solid. It is reasonable to expect that such values are constrained by such factors as the pore volume within cell lumens, vessels, some swelling of hemicellulose, and the presence of LCCs that prevent what otherwise might be a form of delamination at the nanoscale of wet wood. As noted by Fredriksson (2019), under dry conditions, much of the water is associated with wood material via hydrogen bonding, whereas under saturated conditions capillary forces draw water into the structure.

Mechanical refining of pulp

To prepare cellulosic fibers for papermaking, their aqueous suspensions are typically passed through mechanical devices called refiners (Gharehkhani et al. 2015). Often this operation is carried out at consistency (filterable solids) levels of 4 to 10%. In a modern pulp refiner, flocs of fibers experience repeated compression and shearing events as the rectangular “bars” of a rotor repeatedly cross a corresponding set of bars on a stator.

There are two main classes of refining practices – refining of mechanical pulps and refining (sometimes called “beating”) of chemical pulps such as kraft pulps. A key difference lies in the fact that mechanical pulps are generally fed into the refining system in the form of wood chips, whereas the chemical pulps already have been liberated as individual fibers due to the breakdown and dissolution of the lignin. The equipment, processes, and various options for pulp refining have been discussed elsewhere (Li et al. 2011; Gharehkhani et al. 2015; Kerekes 2015). As noted by Lahtinen et al. (2014), both main classes of pulp refining generally result in a fibrillated appearance of the material. In other words, some unraveling takes place of the outer fibrillar layers due to the repeated compression and shearing event. Kraft pulps were found to fibrillate relatively easily and quickly, yielding large values of WRV (e.g. 380%). This effect is consistent with the breakdown and removal of cellulosic strands encircling the dominant S2 sublayer of pulp fibers. Increases in WRV due to refining can be largely attributed to delamination within the cell walls, i.e., internal fibrillation (Przybysz et al. 2017). By contrast, mechanical pulps did not reach as high levels of WRV and there was substantial fragmentation of fibers, in addition to visible fibrillation (Lahtinen et al. 2014). The fines fraction of refined pulp has been shown to hold onto a disproportionate amount of the water detected by WRV and related tests (Laivins and Scallan 1996; Olejnik et al. 2017).

Figure 9 presents a schematic view of how the cross-section of a kraft pulp fiber is expected to change over the course of extensive mechanical refining. The manner in which fibers are packed together in the woody tissue of a tree gives rise to the rectangular shape of unrefined wood-derived fibers, as depicted on the left side of the figure.

Fig. 9. Schematic view of (left) the cross-section of an unrefined kraft pulp fiber, showing the layered structure of the cell wall and (right) an intentionally exaggerated view of the same fiber’s cross-section after mechanical refining, featuring swelling of the partly delaminated S2 sublayer, mechanical breakdown and either removal of fibrillation of the outer layers, and partial or complete collapse and closure of the lumen space. Figure previously used by Debnath et al. (2022)

Repeated shearing and compression events as the fiber passes through one or more mechanical refiner stages is expected to impart swelling. In this illustration, the bright green S2 sublayer is depicted as becoming thicker, presumably in the course of internal partial delamination, with the intake of water. The breakdown of the outer layers of the kraft fiber, as well as some of the S2 sublayer, may partly result in attached cellulosic fibrils extending outward from the main part of the fiber. When such fibril material becomes detached during the course of refining, they become part of the fines component of the fibrous slurry. These narrow cellulosic materials will be in contact with large amounts of water, leading to the question as to whether this can be regarded as another contribution to swelling.

Preparation of nanofibrillated cellulose

Considerably greater increases in apparent swelling can be achieved if mechanical compression and shearing of cellulosic material is continued way beyond the levels associated with ordinary papermaking, i.e., with the production of microfibrillated (MFC) or nanofibrillated cellulose (NFC) (Eichhorn et al. 2010; Lavoine et al. 2012; Hubbe et al. 2017). The resulting materials, often consisting of highly branched and diverse strands of cellulose having diameters less than about 100 nm, typically have no restraints on their swelling. When the amount of water is kept in a suitable range, hydrogels are formed (Hubbe et al. 2013). However, in violation of the usual expectation for a proper hydrogel, such mixtures can continue to take up more water until they become viscous non-gelled suspensions (Hubbe et al. 2017). By contrast, a proper hydrogel will be composed of hydrophilic (usually negatively charged) polymer segments with sufficient covalent cross-linking to prevent their dissolution (Ganji et al. 2010; Hubbe et al. 2013).

A variety of mechanical devices, as well as different chemical and enzymatic treatments, can be employed to prepare nanofibrillated cellulose and related products from a variety of cellulosic materials. From a practical standpoint, there can be advantages of simply using ordinary pulp refining equipment, which is a mature technology employed at paper mills throughout the world (Gharehkhani et al. 2015). By carrying out very large numbers of repeated passages through such refiners, gel-like MFC suspensions can be obtained (Chen et al. 2016; Shafiei-Sabet et al. 2016). At a laboratory scale, various homogenizer devices can be used effectively to achieve different degrees of fibrillation by varying the number of passes. For instance, Pääkkö et al. (2007) used a high-pressure homogenizer to convert oxidized cellulose into nanocellulose suspensions having exceptionally high viscosity at a given solids content. Others have used devices in which a pair of jets collide directly, causing suspended cellulose particles to become fibrillated (Dimic-Misic et al. 2016). Finally, friction-grinding equipment can be used as a way to prepare highly fibrillated cellulose suspensions at the lab scale (Hassan et al. 2011). Products of such shearing devices, as long as they are kept in their wet state, can swell with water without any upper limit. However, such mixtures may pass from being gel-like to being liquid-like at some point, since typically there are no permanent attachments between the adjacent particles while they remain in a wet condition.

Aqueous Conditions

Colloidal science considers the forces of interaction between materials at very short distances in aqueous media (Hubbe and Rojas 2008; Benselfelt et al. 2023). Thus, many of the factors related to water that affect swelling of cellulosic materials fall within that field. Here the attention will be on the evidence, considering in what ways the swelling has been found to depend on pH, the nature of charged groups on cellulosic materials, the concentrations of salts, and effects due to various multivalent ions.

Table 3. Reported Effects of pH on the Swelling of Cellulosic Materials

pH and fiber charge

Most effects of pH on the swelling of cellulosic materials can be attributed to the degree of dissociation of ionizable groups, especially carboxylic acid groups, which will be considered in more detail later. Table 3 summarizes some key findings related to observed pH effects. As shown, it has generally been found that swelling increases with increasing pH. The results are in agreement with the relationship between pH and dissociation of carboxylic acid groups (Herrington and Petzold 1992).

As mentioned in Table 3, the findings of Lindström and Kolman (1982) provide a good reference point, since they involve ordinary unbleached and bleached kraft pulps. Their main findings are shown in Fig. 10. As indicated in part A of the figure, only minor pH effects were observed in the case of bleach kraft pulp. When comparing pH effects during refining in the absence of salt (distilled water), an increase in WRV from about 172 to about 182% was observed when the pH was raised from about 3 to 5. But when the WRV was retested in pH=4 buffer solution, all of the values were about the same, regardless of the pH. In addition, no effect of pH on WRV was detected when refining was carried out in the presence of salt (0.1 N NaCl). Part B of Fig. 10 shows a different story when evaluating the corresponding unbleached kraft pulp. Increasing pH during refining yielded relatively large increases in WRV, at least up to pH=10. The decrease observed above pH=10 in the absence of salt is consistent with the higher ionic strength that is required to reach those high pH values. Ionic strength effects will be discussed next. Note that the effects of pH on WRV were still apparent for the unbleached kraft pulp even after the pulp had been rinsed and adjusted to pH=4 with a buffer.

Fig. 10. Redrawn figures from Lindström and Kolman (1982) showing effects of pH during refining of (A) bleached kraft pulp and (B) unbleached kraft pulp. Refining was at the pH values shown either with no salt of 0.1 N NaCl. WRV measurements were done either in the same solution or after rinsing with pH-4 buffer solution

Salts

Simple monovalent salts, such as NaCl, are known to suppress osmotic effects that otherwise would promote swelling, as will be discussed later in the article. Empirical evidence of such effects is widely available. For instance, Fig. 10 shows a case in which the presence of 0.1 N NaCl decreased the ability of a certain amount of mechanical refining to increase the WRV of both bleached and unbleached kraft pulps. In the case of the bleached kraft pulp, the effect of salt disappeared after the pulp specimens had been rinsed and resuspended in mildly acidic buffer solution, but in the case of the unbleached kraft pulp, which tends to bear a higher density of carboxylic groups (Lloyd and Horne 1993; Lindgren et al. 2002; Hubbe et al. 2012), the effects of salt during refining were still apparent in the pH=4 buffer solution.

Effects of salt can be expected to be more dramatic in cases where the cellulosic materials have been treated to increase its charge density of carboxylic acid groups. Thus, large decreases in WRV were observed upon salt addition to pulp fibers that had been treated with carboxymethyl cellulose (CMC) (Laine et al. 2002). In the cited study it is notable that the charged groups were associated with the cellulose merely by an adsorption process of the CMC onto the fibers, with no covalent attachment. The effect was greater, per unit of adsorbed CMC, when the molecular mass of the CMC was large. The authors attributed those results to an extended, bulky conformation of CMC chains attached externally to the cellulosic fibers. Related effects were found by Karlsson et al. (2018), who studied carbohydrate gels that had been carboxymethylated to different degrees. Salt effects became important at pH values of about 7 and higher. Salt-free gels having high carboxylate content reached WRV values up to about 140%, whereas the corresponding WRV values in the presence of 10 mM NaCl were about 73%.

Cation valence

The ability of positively charged ions to suppress the swelling of negatively charged cellulosic materials increases strongly with increasing valence (Scallan 1983; Lindström 1992; Maloney 2015; Kummala et al. 2018). Such effects were shown in particular by Scallan and Grignon (1979). Higher-valence metal ions suppressed the fiber saturation point (FSP) values of sulfite and kraft pulp fibers to a greater extent. The suppression of swelling was highly correlated with decreases in the strength of paper sheets formed form the respective fibers. Similar effects of cationic valence on swelling and fiber strength were reported by Katz et al. (1981). Notably, the hydrogen ion, associated with low pH, had effects similar to those of the trivalent aluminum ion. This is consistent with the findings of Scallan and Tigerström (1992), who showed that exchanging sodium ions in place of hydrogen ions led to large increases in the swelling of pulp fibers.

Benselfelt et al. (2019) found that the ions that were most effective for lowering the swelling of negatively charged nanofibrillated cellulose were those that developed strong complexation. Of the divalent ions, the magnesium ion had the minimum deswelling effect, whereas barium (a highly polarizable divalent ion) and copper (a much less polarizable ion than magnesium) had greater deswelling effects. A similar relationship was found in the case of monovalent cations (Benselfelt et al. 2023), such that the greatest swelling was observed for Li+, somewhat less for Na+, and the least for Cs+. These effects were observed in the case of negatively charged nanocellulose films.

Cationic polyelectrolytes

Cationic polyelectrolytes likewise have been shown to decrease the swelling of cellulosic materials in suspension (Swerin et al. 1990; Ström and Kunnas 1991; Zhang et al. 2002; Aarne et al. 2012). Thus, it was found that a low-mass quaternary ammonium polymer (polybrene) was able to suppress the swelling of hardwood kraft fibers to lower WRV values than treatment with high-mass poly(diallyldimethylammonium chloride) (polyDADMAC) (Aarne et al. 2012). The polyDADMAC was able to achieve a relatively high decrease in WRV (from about 355% to about 250% in the case of highly refined fibers) at a lower added amount, making it a more efficient deswelling agent. Ström and Kunnas (1991) found, likewise, that higher-mass polyethylenimine (PEI) was both more efficient (requiring a lower dosage) and more effective (reaching a lower WRV) in comparison to a lower-mass PEI product. Aarne et al. (2012) found that adsorption of polybrene onto bleached hardwood kraft pulps mainly decreased the swelling of the larger pores in the cell walls, having initial diameters in the range of about 27 to 220 nm.

Anions acting as swelling agents

Though it has received less attention in published articles, certain ions having the same (i.e., negative) charge as the cellulosic material have been shown to increase swelling, in contrast to, for instance, the chloride ion. Thus, Fält and Wågberg (2003) observed significantly higher WRV values of unbleached kraft pulp in the presence of sodium sulfate at a 0.1 M concentration, in comparison to NaCl. The effects were shown to correlate to the results of tests with quartz crystal microbalance with dissipation (QCM-D) (Fält et al. 2003). Thus, the Na2SO4 solution gave rise to a looser, presumably more swollen cellulosic film. Sodium bicarbonate solutions have shown related effects (Kahar et al. 2013), but effect of the pH of those solutions makes such findings harder to interpret.

Bendzalova et al. (1996) reported that treatment of wood chips with “swelling agents” including amines, carbonates, and chlorides before high-yield pulping at high temperature led to higher swelling. The WRV of the resulting pulp was increased from about 100% in the case of water to almost 500% in the case of sodium carbonate. Related results were found with measurements of the fiber saturation point, but only a fair correlation was observed between the two different measurements of swelling. It is possible that such effects bear a relationship to the ionic liquid solvent systems to be described later; thus, higher swelling may be associated to combinations of positive and negative ions that do not have a tendency to form crystals.

Attributes of the Cellulosic Material

Many factors that affect the swelling of cellulosic materials fall into the category of attributes of the cellulosic materials themselves. Some of these attributes, being so central to the discussion, have already arisen when considering aspects of biosynthesis, delignification, mechanical refining, and the effects of aqueous conditions. The purpose of this subsection is to consider such aspects as chemical composition, aspects of microstructure, chemical modifications, effects related to delamination, effects related to the crystalline nature of cellulose, and effects on swellability related to drying.

Cellulose fibrillar orientation

The form of an especially dramatic manifestation of swelling, called ballooning (Cuissinat and Navard 2006; Zhang et al. 2013; Budtova and Navard 2016), has been attributed to the differing orientations of cellulose microfibrils in the different layers of a cellulose fiber cell wall. Aspects of ballooning will be considered in more detail later in this article when considering solvent systems for cellulose. Figure 11 provides a schematic illustration of the ballooned structure of a typical kraft pulp fiber that has been subjected to swelling in cold, moderately concentrated (7.6%) NaOH solution (Cuissinat and Navard 2006), especially in the presence of thiourea and urea (Zhang et al. 2013). Because most of the mass of the cell wall is contained within the S2 sublayer of the fiber, that sublayer accounts for most of the observed swelling. But the S2 sublayer is wrapped by the very thin primary (P) layer and the S1 sublayer, both of which have at least some cellulose fibrils oriented to encircle the fiber. In particular, the predominant microfibril angle in the S1 sublayer of Norway spruce wood species has been measured as 89 degrees, implying that it wraps the fiber perpendicular to the fiber axis (Andersson et al. 2000). Apparently, the swelling of the thick S2 sublayer causes it to burst through parts of the P and S1 layers, and that outer cellulosic material slides and is forced into narrow bands (Cuissinat and Navard 2006). Le Moigne and Navard (2010) described the “rolling up” of the outer cellulose layers into collars.

Fig. 11. Features of fir kraft fibers highly swollen (“ballooned”) in NaOH-water solution, at different stages of the swelling process (redrawn in simplified form from Cuissinat and Navard 2006)

Confirmation of the mechanism just described comes from the fact that the ballooning phenomenon does not arise in some other notable forms of cellulose. For instance, regenerated cellulose fibers and filaments prepared after complete dissolution do not show ballooning phenomena. On the other hand, ballooning is still seen when studying intact plant-based cellulosic fibers that have been chemically derivatized (nitrocellulose, cyanoethylcellulose, cellulose xanthate fibers) (Cuissinat et al. 2008).

The orientation of crystals in a film of cellulose nanocrystals (CNCs) has been shown to affect the nature of swelling (Shrestha et al. 2017). Only isotropic swelling was observed when self-organized cellulose nanocrystal (CNC) films were exposed to high humidity conditions. By contrast, films composed of shear-oriented CNC particles showed increases in film dimensions of 0.02 and 0.30% in the parallel and perpendicular directions, respectively (Shrestha et al. 2017). This is consistent with the fact that plant fibers typically swell very little in their length directions, while they show various degrees of swelling perpendicular to their axes (Solhi et al. 2023).

Lignin content

As was shown already in the context of conventional delignification (i.e. chemical pulping) by breaking down and dissolving lignin from plant material, the resulting wet material contains substantial pore volume. To further explore such effects, Bai et al. (2022) used a combination of phosphoric acid and hydrogen peroxide to treat sugarcane bagasse. The treatment was found to remove essentially all of the hemicellulose and about 98% of the lignin. The void volume was increased by a factor of about ten relative to the untreated bagasse.

Such effects suggest two contributing mechanisms, the first of which is the fact that spaces are left behind after the removal of the solubilized material. The second mechanism is related to the relatively stiff, hydrophobic nature of lignin, in comparison to the other main components of woody material. Presumably after that stiff material is gone, the remaining structure will be more compliant and thus able to swell. It is also possible that the breakage of covalent bonds associated with LCC may play a role.

Hemicellulose content

Various work has shown that removal of both lignin and different levels of hemicellulose tends to increase swelling of the resulting fibers. The situation considered by Bai et al. (2022), as discussed above, was special, since the treatment had removed not only the lignin, but also all of the hemicellulose. A number of studies have considered effects in which the level of hemicellulose remaining in the fiber material was systematically varied. For example, Katz et al. (1981) progressively removed more hemicellulose from spruce and aspen mechanical pulps by NaOH treatments of differing severity.

The fiber saturation point was found to increase in a linear fashion in each case with the increasing content of acidic groups, which are mainly associated with the hemicellulose component. Likewise, work by Pejic et al. (2008) showed that removal of hemicellulose, along with some lignin from hemp fibers increased the capillary rise but decreased the WRV.

These WRV findings are consistent with the hydrophilic nature of hemicellulose and its tendency to swell in water. Another way to estimate effects of hemicellulose in pulp specimens is to compare the carboxylic acid group contents, of which the hemicellulose xylan is typically the main contributor. Zanuttini and Marzocchi (2003) showed that the WRV of chemi-mechanical pulps increased to an accelerating degree with increasing acidic group content.

Palasingh et al. (2021) discovered a unique effect with xylans, a hemicellulose component from hardwood pulps, were added to other polysaccharides as combined binders in the preparation of nanocellulose films. Though both components of the binder mixture were hydrophilic, the combination proved less susceptible to swelling. The effects were tentatively attributed to nanostructural effects.

Cellulose chemical modification with charged groups

Studies have shown that increasing amounts of ionizable groups in cellulose lead to greater swelling in water. Such findings are highlighted in Table 4. Such findings have been demonstrated, for instance, when cellulose is oxidized, giving rise to carboxyl groups. Hashemzehi and Sjöstrand (2022, 2023) prepared gel-like suspensions of extremely swelled fibers from commercial pulps by a combination of chemical treatments of oxidation and deep eutectic solvents, and even if the structure became extremely swollen, fibers were still visible in polarized light microscopy. Similarly, Table 5 highlights the findings of studies in which covalent reactions were employed to attach charged groups to cellulose, with evaluation of effects on swelling.

Table 4. Oxidation of Cellulose and its Effects on Swelling

Note: TEMPO = 2,2,6,6-tetramethyl-1-piperidinyloxy

Table 5. Derivatization or Mixing of Cellulose with Negatively Charged Groups or Adsorption of Such Groups and its Effects on Swelling

Cellulose chemical modification with uncharged groups

Though the reasons are somewhat different, derivatization of cellulosic materials with uncharged substituent groups also has been shown to affect swelling in water. Such findings are highlighted in Table 6. Based on such findings, it appears that a moderate level of substitution with uncharged groups can interfere with possible hydrogen bonding interactions between adjacent cellulosic surfaces. However, at higher levels of substitution, the hydrophobic nature of the attached groups can have a dominant effect, thus rendering the material significantly more hydrophobic and thereby decreasing swelling in water at higher levels of substitution.

Table 6. Derivatization of Cellulose with Uncharged Groups and its Effects on Swelling

Hornification’s effect on swelling

Effects of drying of cellulosic materials on subsequent swelling in water already have been mentioned in the course of discussing some other factors (Fernandes Diniz et al. 2004; Salmén and Stevanic 2018). The term hornification, as mentioned earlier, is widely used to refer to such loss of swelling ability, and the most commonly used method to quantify hornification is water retention value (WRV) tests (SCAN-C 62:00, 2000; TAPPI UM 256 1981; ISO 23714:2014), which involve centrifugation. The most severe effects of this nature are associated with conventional air-drying of cellulosic material from which lignin and much of the hemicellulose has been removed (Hubbe et al. 2007b). Such susceptibility is consistent with the compliant nature of kraft fibers, especially after mechanical refining, the presence of pores in the mesopore size range, and the fact that freshly exposed cellulosic surfaces are no longer isolated from each other by layers of lignin. Drying of paper also gives rise to strong capillary forces, which draw cellulosic surfaces into molecular contact (Page 1993; Wohlert et al. 2022), thus promoting the establishment of dense areas of hydrogen bonding.

The mechanism of this phenomenon was proposed to be driven by chains of hydrogen bonds (Sjöstrand et al. 2023), although this does not account for the temperature dependence reported in several publications (Luo and Zhu 2011; Salmén and Stevanic 2018; Sellman et al. 2023). In fact, there are some disagreements in the literature on the mechanisms for hornification, probably due to the fact that multiple mechanisms can be simultaneously contributing to the reduced swelling during drying cycles. The hornification effects have been shown to decrease when using liquids other than water (Sjöstrand et al. 2023; Hashemzehi et al. 2024), and in some cases even negative hornification was achieved (Hashemzehi et al. 2024). Negative hornification is when the fiber swelling is higher when fibers are dried from a solvent compared with a never-dried fiber reference. It has been demonstrated that negative hornification may occur during the drying of cellulose from specific solvents, which is particularly evident with less polar solvents. The underlying idea behind this phenomenon is that the presence of the solvent could lead to increased cellulose swelling by disrupting cellulose-cellulose bonds during the drying process. This disruption might expose and expand the inter- and intrafibrillar structure more than that found in a never-dried pulp.

Cellulose is typically bonded together in layers by hydrogen bonds, with hydrophobic interactions and van der Waals forces maintaining the layers’ cohesion. Similar interactions could occur between cellulose fibrils (Medronho et al. 2012). The hypothesis from Hashemzehi et al. (2024) suggests that when a non-polar solvent infiltrates cellulose surface (Norgren et al. 2023), it potentially weakens the hydrophobic interactions, thus leading to possible cellulose swelling, ultimately resulting in negative hornification. Supporting this idea, the weakening of non-polar interactions has been proposed to play a role in cellulose dissolution (Lindman et al. 2010). Additionally, other factors might contribute to this phenomenon, such as the solvent’s interaction with the polymer itself. For instance, the cellulose chains might become stiffer upon contact with certain solvents, further contributing to a more open structure.

Higher temperature during drying of pulps from water has been shown to result in stronger hornification (Welf et al. 2005). This phenomenon was especially stronger for drying temperatures over 100 °C. Sjöstrand et al. (2023) suggested that this is due to two different mechanisms. For the highest temperature, dehydration might give rise to covalent crosslinking of fibers. This is supported by a yellowing of these fibers. The temperature dependence under 100 °C is more puzzling, since hydrogen bond chains and thereby osmotic pressure are expected to decrease at higher temperatures; a possible explanation is that higher mobility on superficial cellulose chains (Fig. 3), as introduced by Salmén and Stevanic (2018). This allows cellulose surfaces to interact more firmly and with more hydrophobic interaction and van der Waals interactions. That the latter interactions might be important in hornification is supported by the results of von Schreeb et al. (2024); in that study microcrystalline cellulose was highly swollen by a partial dissolution in alkali, followed by precipitation, thereby creating a very swollen and amorous form of cellulose. The highly swollen cellulose became much more strongly hornified than the original material, and it was suggested that this was due to that the more amorphous cellulose had more hydrophobic surfaces exposed (von Schreeb et al. 2024). Table 7 highlights some key findings of studies showing how the drying of cellulosic materials has been found to affect subsequent swelling in aqueous media.

Table 7. Findings Related to the Hornification of Cellulosic Materials, i.e. the Loss of Aqueous Swelling Ability Due to a History of Having Been Dried

Crystal form of cellulose and related effects

By means of preparing regenerated cellulose products, technologists have been able to change the crystal form and other aspects of the material. Table 8 shows studies in which such differences in crystal type suggest that it may have played a role in swelling. It is important, when viewing such findings, to keep in mind that such transformation in crystallinity leads to profound changes that are not limited to just the crystalline zones. For this reason, caution should be exercised in drawing conclusions.

Table 8. Cellulose Crystal Type as a Possible Factor Contributing to Swelling in Aqueous Media

Transformation of cellulose I to cellulose II occurs upon dissolution and regeneration of cellulose. The change also takes place after treatment with sodium hydroxide in mercerization of cellulose and subsequent removal by washing, even though the fibers remain intact during such processing. In the mercerization, sodium ions are incorporated in the structure as “alkali cellulose”. Studies have shown that alkali concentrations of 8 to 10% may transform a substantial part of the cellulose to cellulose II (Dinand et al. 2002; Schenzel et al. 2009). It has also been discussed that some steps in the pulping and/or bleaching processes may result in a higher amount of cellulose II and a higher degree of hornification (Ferreira et al. 2020).

THEORETICAL ASPECTS OF SWELLING OF CELLULOSIC MATTER

Adsorption, Wetting, and Permeation

Having considered, in the previous section, various evidence related to factors that affect swelling, the focus of this section will be on finding general explanations. These can be broadly assigned to the categories of thermodynamics, physical constraints, chemical effects, and kinetic effects. Thermodynamics considers ways in which physical systems arrange themselves to reach a minimum state of free energy, taking account of heats of interaction and a preference for freedom, i.e. maximization of entropy (Ganji 2010; Alexandersson and Ristinmaa 2018). Physical constraints can be considered in terms of the mechanics of materials, including the modulus of elasticity and other materials properties. Chemical effects include osmotic pressure (Grignon and Scallan 1980). Chemical effects also include consideration of solvent systems that have been optimized for the swelling of cellulose, up to its complete dissolution. Kinetic effects include consideration of diffusion-related processes and capillary wicking rates, in addition to considering trapped non-equilibrium states.

At a basic level, the interaction of water and cellulose can be regarded as a type of adsorption (Brunauer et al. 1938). As described in the cited article, even without considering the possibility of swelling, it is possible to account for many features of such interactions, including multilayer adsorption at very high levels of relative humidity. A further advance in understanding was achieved by thinking in terms of the interactions within clusters of water molecules (Frank and Wen 1957). When considering the polar nature of the -OH groups at cellulosic surfaces, Berthold et al. (1996) considered the relationship between such groups and the number of adsorbed water molecules.

The basic thermodynamics of water adsorption onto cellulosic materials has been reviewed (Eklund and Lindström 1991). Briefly stated, thermodynamics considers the equilibrium processes of water and cellulosic materials. Such processes will proceed in such a way as to minimize the Gibbs free energy of the system, which contains an enthalpy (heat) term and an entropy (randomness) term. Thus, at equilibrium, the chemical potential (Gibbs free energy per monomeric unit) of an adsorbed water molecule will equal that of a non-adsorbed one. The approach has been developed further to predict the pressures and relative rates of evaporation from curved menisci, such as when a pore is fully or partly filled with water (Thomson 1871; Fisher et al. 1981; Galvin 2005). Eklund and Lindström (1991) show how such concepts can be used to explain or predict adsorption behavior onto porous materials, including paper.

In general, for interaction between water or humid air and fibers, a change in Gibbs free energy is the driving force, as given by:

(1)

In Eq. 1, nw and nf are the molar quantities of water and fiber, respectively and ∆µw and ∆µf are the change in chemical potential for water and fiber, respectively. The swelling of cellulosic fibers and fibrous networks, the change in Gibbs free energy ∆Gswell consists of several contributions and can be summarized as,

(2)

in which ∆Gsorption is from both specific and non-specific moisture and liquid sorption. This contribution is from general moisture sorption to surfaces, as from moisture adsorption and desorption isotherms (Eklund and Lindström 1991), as described above. ∆Gmixing is the change in Gibbs free energy from mixing, with contributions from osmotic pressure and Gibbs-Donnan equilibria, and in general to which Flory-Huggins theory can be applied (Burchard 2003). ∆Gionic is from electrostatic interactions, such as depicted by the DLVO theory and specific ionic interactions (Derjaguin and Landau 1941; Verwey and Overbeek 1948; Benselfelt et al. 2023). The ∆Gnetwork term comes from restrictions to swelling from the fiber lamellae structure (see above), whereas ∆Gcapillarity comes from wetting along and inside fibers and networks as well as from capillary work, in which a ∆P term, in most cases, is the Laplace capillary pressure but it can also comprise an external pressure. The latter two contributions are discussed below as well as osmotic contributions. However, for detailed discussions of Eq. 2, readers are encouraged to study the references cited above. Recently, an approach based on calorimetry data and MD simulations has been suggested by Benselfelt et al. 2024, in which cellulosic fibers could be considered as polyion-polydipoles showing considerable entropy-driven interactions in water.

The individual or overall change in Gibbs free energy can be negative or positive. A negative value indicates a spontaneous process, whereas a positive value implies that a process would cost energy to occur. In a larger context, swelling is governed not only by thermodynamic considerations but also kinetic ones, and it further depends on domain sizes and interacting distances.

More and more, the understanding of the forces that govern how biological structures are assembled has developed. The term self-assembly of biological systems is by no means young (Kushner 1969), but it has grown in importance as self-assembly more and more serves as inspiration when materials and products are innovated and developed. The assembly of lignocellulosic materials is now studied with great effort. Plant organisms that are constituted by such materials show a vast number of properties that are of interest.

Self-assembly, from a physical and chemical point of view, is governed by thermodynamics, which involves entropy and enthalpy (Hubbe et al. 2023). In self-assembly, a metastable state is preferable from an energy point of view. From a metastable state, the system can then overcome the entropic drive to maximize degrees of freedom. This circumstance is observed in nature in several examples, such as cell membranes, RNA/DNA, and cellulose (Etale et al. 2023).

In many cases, water acts as a mediator for the assembly and stabilization of biological molecular systems (Dargaville and Hutmacher 2022). Thermodynamics is essential here. Water produces an environment that can drive metastable configurations and long-range molecular order, encompassing not only the hydrogen bonding, but also van der Waals forces. These become more important as the hydrogen bonding moieties of molecules force the more apolar parts closer together, animal cell membranes being one example (Cresswell et al. 2021; Furman et al. 2021).

The thermodynamics of the complete system becomes vast when considering numbers of molecules involved, their interaction, and all structures engaged. Thus, the task of modelling such systems in detail becomes very challenging (Shen and Gnanakaran 2020). The number of molecules that can be modeled to a sufficient degree of accuracy, when using more detailed modeling methods such as molecular dynamics, are often quite small (Araujo et al. 2018; Chen et al. 2022; Wohlert et al. 2022). This means that very long range and large systems effects can be difficult to capture, in particular their summative effects on a higher structural scale (Sinko et al. 2015; Salem et al. 2022; Thybring et al. 2021, 2022).

Crystal/microfibril level

Some important questions arise with respect to events that occur within the size ranges of crystals and microfibrils. These include, how far do the order effects reach (Jarvis 2023; Lindman et al. 2017). With this respect, what is the relevance of structural features of cellulose? With respect to heterogeneity – the structural heterogeneity should disrupt the reach of order effects. But how large are the entropic gains? What could be overcome?

Removing water and putting it back again

The OH-groups present on cellulose will have to bind to something else. If they bond to other OH-groups in an orderly manner, this will tend towards increasing order and thus larger sizes of crystals. The bonds involved in such self-assembly are not easily opened again. Structure prevails. Long range order induces strain, stiffening the structure. These ideas have been supported by various research (Maloney et al. 1998a,b; Salmén and Stenovic 2018; Grunin et al. 2020).

Chemical effects

With respect to chemical heterogeneity, a question arises as to whether in a similar manner, structural heterogeneity should induce disturbance. Chemical pulping is in some sense a method to remove the heterogeneity, leaving room for continued crystallization of the cellulose (Pönni et al. 2012; Brännvall et al. 2021; Hult et al. 2003; Kihlman et al. 2013; Pönni et al. 2014). Examples of where chemical disturbance can be overcome have been described (Östblom et al. 2006).

Kinetic effects

A further question is whether crystallization and morphological changes (on a nano scale) can be observed in situ. It is known that crystallinity can be developed during pulping, and it can be characterized using various methods. Water bonding changes can be observed. Deuterated water has been used, since vibrations, measured as peaks in IR spectroscopy, can be assigned to bulk water and surface water (Rowland et al. 1995; Engquist et al. 1997; Östblom et al. 2006; Habibi et al. 2010; Suchy et al. 2010; O Neill et al 2017; Sinko et al. 2015; Han et al. 2019; Nishiyama et al. 2019; Borrero-Lopez et al. 2023; Greca et al. 2023). Stresses tend to induce a long-range order on a fibrillar level and eventually impact fiber strength (increase stiffness).

Osmotic Pressure

Ionic strength and charge density

The osmotic pressure that develops within hydrophilic polymeric material immersed in water is a major contributing factor in swelling (Ganji et al. 2010). Proctor (1914) may have been the first to clearly explain osmotic effects as controlling the swelling behavior of hydrogels. Hydrogels generally can be described as loosely crosslinked polymer chains that contain ionizable groups. A historical breakthrough in the understanding of many swelling behaviors in aqueous systems with cellulosic materials can be attributed to a decision to regard those systems as versions of hydrogels (Grignon and Scallan 1980; Scallan and Tigerström 1992). As described by these authors, the extent of swelling then can be estimated by considering a balance between osmotic pressure, which favors swelling, and restraining forces, which are due to the properties of the material, crosslinking effects, and various related factors. As noted by Scallan (1983), it is mostly the carboxylic acid groups within cellulosic materials that contribute to such swelling effects, though one needs to keep in mind that other ionized groups, such as sulfonate groups, can likewise be expected to contribute. A testament to the strength of osmotic forces is the fact that a sufficiently high substitution of cellulose with carboxyl groups can cause it to become converted to nano-sized cellulose fibrils, i.e. NFC, without need for intensive shearing (Sjöstedt et al. 2015). Osmotic forces also play a role in steric stabilization, a mechanism by which hydrophilic polymer chains extending from a surface keep those surfaces from colliding and adhering to each other (Zauscher and Klingenberg 2000).

The osmotic pressure can be calculated based on the following equation (Ganji et al. 2010),

(3)

where Φ is the osmotic coefficient of the gel phase and φ is the corresponding quantity for the external aqueous solution. Likewise, is the concentration of the ith ionic species in the gel phase and is the corresponding value in the external solution phase. At equilibrium, one can expect a balance between the osmotic pressure and any elastic forces tending to limit swelling in the material (Grignon and Scallan 1980). The osmotic coefficient of the gel phase is given by Eq. 4 (Ganji et al. 2010),

(4)

where np is the molarity of the polymer, nm is the molarity of the monomer, and α is the degree of ionization as a fraction. Various expected dependencies of swelling effects follow from these basic equations. Thus, swelling is expected to increase with increasing density of the ionic groups on the polymer chains or lamellar surfaces, etc.

Charge Effects

Effects of pH can be explained by the fact that a sufficiently low pH will cause protonation of carboxylate groups, thus decreasing the expressed charge density. The pKa value, which is the pH at which there are an equal amount of charged and uncharged groups, is typically about 3.3 for carboxyl groups associated with hemicelluloses (Laine et al. 1996; Hubbe et al. 2012). Higher values of pKa are associated with carboxyl groups with other neighbor groups, e.g., those associated with lignin, as well as fatty acids and resin acids in the pulp. As the pH is adjusted to become increasingly higher than the pKa value of the predominant acidic groups, an increasingly higher proportion of those groups will be in their deprotonated, negatively charged form.

The effects of pH typically are spread out over a relatively wide interval. On an external surface or on a small molecule, each successive unit of pH increase above the pKa value results in a tenfold decrease in the remaining uncharged carboxylic acid groups. Thus, at about pH=6 or higher, one can regard such carboxyl groups in typical cellulosic materials as being mainly in their negatively charged form. But swelling often continues to increase over a wider pH range (Lindström and Carlsson 1978; Lindström and Kolman 1982; Carlsson et al. 1983; Gellerstedt et al. 2000; Karlsson et al. 2018). The explanation lies in the fact that pH values can be markedly different in the interior of a material that behaves like a hydrogel (Grignon and Scallan 1980). The reason for such a difference can be explained based on a Donnan distribution of ions (Laine and Stenius 1997). Within the hydrogel, the ionic groups include those that are immobile – attached to the polymeric material. Those groups do not exist outside of the hydrogel. Differences in the concentrations of specific ions inside and outside of structure resembling a hydrogel are required because nature will not allow a buildup of net charge at any location. Thus, within the hydrogel there will be an equal number of positive and negative ions present, which will include the bound ions. As a practical consequence of the Donnan equilibrium, the pH is often two or three points lower within a cellulose-based hydrogel than outside, though the difference becomes less with increasing salt addition (Grignon and Scallan 1980).

The increasing concentrations of simple salts, such as NaCl, strongly decreases osmotic pressure values, thus decreasing swelling effects. As noted already, such effects have been widely reported for cellulosic materials (Grignon and Scallan 1980; Laine et al. 2002; Karlsson et al. 2018). Corresponding effects have been shown in swollen nanocellulose films (Ahola et al. 2008; Reid et al. 2016, 2017; Benselfelt et al. 2023), and in ultrathin cellulose films regenerated onto the surfaces of quartz crystal microbalance surfaces (Fält et al. 2003).

An alternative, and possibly equivalent way to account for the effects of surface charge and salt concentrations, with respect to the swelling of cellulosic materials, is based on repulsive forces between like-charged surfaces. These forces were first theoretically calculated by Dejaguin and Landau (1941) and Verwey and Overbeek (1948), who together are recognized for developing the so-called DLVO theory. The notable feature of this theory is the principle that the repulsion effect arises not directly from the charges on the facing surfaces, but rather due to the overlap of the adjacent layers of counter-ions, i.e. clouds of sodium or potassium ions, etc., which balance the net charge on the surfaces.

Such an effect is illustrated pictorially in Fig. 12. In the figure, the pink area represents cellulosic material, and the gap in that material is supposedly filled with aqueous solution. The forces are expected to be strongest when the physical distance between the surfaces is approximately 0.5 to 2 nm, i.e. wide enough to physically accommodate a double layer, but not too much larger than that. The DLVO theory predicts a decreasing range of electrostatic repulsion with increasing salt in systems related to cellulosic materials (Hubbe and Rojas 2008), which is in line with predictions based on osmotic pressure calculations.

Fig. 12. Simplified qualitative representation of the potential swelling effect due to repulsive forces between facing cellulosic surfaces having the same sign of net charge, shown here as originating from the presence of carboxylic acid groups, where the counter-ions are shown as Na+, and where the co-ions are shown as Cl

Complexation

As was discussed earlier in the article, multivalent cations, including various metal ions and cationic polyelectrolytes, have been shown to suppress swelling of cellulosic materials to a greater extent than monovalent cations, such as Na+ (Scallan 1983; Swerin et al. 1990; Lindström 1992; Aarne et al. 2012; Maloney 2015; Kummala et al. 2018). For example, it has been noted that such forming such complexes in nanocellulose gels can greatly suppress swelling (Benselfelt et al. 2023). In addition, although separately each polyelectrolyte may be water-soluble, a well-formed film composed of complexed polyelectrolytes can be water-resistant with relatively low permeability to liquids and gases (Hubbe 2021).

Many of these effects can be placed within the category of polyelectrolyte complexation (Ström and Kunnas 1991; Chang et al. 2011). One of the key findings when two charged polyelectrolytes are mixed with each other is that the minimum swelling coincides with a state of equal interactions between opposite charges on the two kinds of polyelectrolytes (Chang et al. 2011). Notably, the stoichiometry of interaction between a strong polyacid and strong polybase in solution approaches a strict 1:1 ratio of charged groups in the relative absence of salt ions (Chen et al. 2003). PECs are known to undergo an unusual “antipolyelectrolyte effect” in which the degree of swelling increases with increasing salt addition (Dautzenberg and Jaeger 2002; Valencia and Pierola 2007). This effect can be attributed to the fact that the salt, within a certain range of concentration, weakens the complexation between the chains before sometimes causing the PEC to come apart. Such systems can have the unique ability to absorb and remove salt from brine solutions (Ayoub et al. 2013).

The mechanism by which polyelectrolyte complexation takes place is illustrated in Fig. 13. At left in the figure, one visualizes two separate polyelectrolytes, presumably being combined as dilute solutions from different containers. Each of these separate kinds of polyelectrolytes, in salt-free solution, has mainly two types of ions – those bound to the chain and a set of oppositely charged monovalent ions (the counter-ions) that have more freedom of motion. As depicted in the figure, when a pair of oppositely charged ions forms a polyelectrolyte complex (PEC), a direct pairing takes place between ions attached to the contrasting macromolecular chains. Because the polyelectrolyte chains, in addition to the charged groups attached to them, are already highly constrained in terms of possible movements, only a relatively small amount of freedom of motion (entropy) is lost by such pairing. On the other hand, the former counter-ions of both signs of charge now are able to circulate freely within essentially all of the aqueous solution in the system. Thus, the PEC process tends to be driven to completion by a combination of two energy terms. These are namely the increased entropy mainly due to the more fully liberated former counter-ions and the heat of interaction associated with forming relatively stable ion pairings along the chains.

Fig. 13. Schematic representation of the interaction between two initially separate solutions of oppositely charge polyelectrolytes to form polyelectrolyte complexes, with the pairing of chain-bound charge groups and increased freedom of the fully liberated former counter-ions

It has been proposed that certain chemical additive programs that are widely employed during the industrial production of paper involve deswelling mechanism related to the polyelectrolyte complexation just described (Hubbe 2005). During the papermaking process, water needs to be removed from a layer of fiber slurry as it travels over a continuous loop of a mesh screen (or in the gap between two such screens). It had become well known that the dewatering process could be promoted by the additional of cationic polymers, with optimum results associated with adding the needed amount to just neutralize the net colloidal charge (Horn and Melzer 1975). Such observations are consistent with some effects of cationic polyelectrolytes on swelling (Swerin et al. 1990; Ström and Kunnas 1991; Zhang et al. 2002; Aarne et al. 2012), as mentioned earlier in this article. However, even greater dewatering effects have been observed in papermaking systems when treatment of the pulp furnish with a cationic polymer (usually either a very-high-mass cationic acrylamide product or cationic starch) is followed by either the sodium form of montmorillonite (commonly called bentonite) or various forms of colloidal silica (Andersson et al. 1986; Langley and Litchfield 1986; Andersson and Lindgren 1996; Hubbe 2005). As depicted in Fig. 14, the loops of cationic polymer are expected to collapse onto the surfaces of the small negatively charged particles, thus giving rise to a contraction and deswelling of the polyelectrolyte coils, i.e. a self-wringing sponge effect (Debnath et al. 2022).

Fig. 14. Schematic illustration of a proposed deswelling effect when a fibrillated cellulosic fiber (shown in cross-sectional view) has just been treated with cationic acrylamide copolymer of very high molecular weight (left view) and subsequently is treated with an optimal level of a negatively charged particulate entity with very high surface area, e.g. colloidal silica, or a sodium montmorillonite (bentonite) product

Structural Restraining and Loosening of Restraints

The word “structure” can be regarded as a starting point for understanding and possibly quantifying the forces that restrain the swelling of cellulosic materials, for instance in opposition to the osmotic pressure effects just described. The cellulose macromolecule, often in the form of fibrils composed of those macromolecules, has a very strong resistance to stretching in its lengthwise direction (Iwamoto et al. 2009). It has been proposed that the elastic modulus of cellulose can be modeled in terms of the force-distance relationships that include a hydrogen bonding term (Kroon-Batenburg et al. 1986).

The high resistance to stretching of cellulose fibrils can provide strong restraint of swelling, depending on structural arrangements. Thus, Dufresne (2012) showed that the presence of stiff, relatively long cellulose nanocrystals tended to decrease the swelling of hydrophilic matrix polymers in a composite film exposed to water. The crystalline nature of cellulose and its degree of crystallinity can be expected to govern its swelling and its mechanical properties that may serve to resist swelling (El Seoud et al. 2008). An indirect demonstration of the forces within cellulose fibers that serve to limit swelling comes from measurements of the swelling of nanocellulose films (Rehfeldt and Tanaka 2003; Moriwaki and Hanasaki 2023). The cited authors pointed out that nanopaper films, composed of NFC, swell in the direction perpendicular to the fibrils to a much greater extent than ordinary cellulosic fibers. The explanation is that such films often lack extensive connections already established between the cellulose fibrils, including intertwined crystalline zones of the cellulose, covalent bonding to adjacent lignin domains (Lawoko et al. 2003), and an intertwined, relatively dense structure involving collaboration among the cellulose, hemicellulose, and lignin, as will be considered next.

Role of lignin

The lignin component of wood and many cellulosic fibers has a reputation as being relatively stiff. It follows that lignin, to the extent that it has not been removed by a chemical pulping operation (Fardim and Tikka 2011), can play a major restraining role with respect to the swelling of the polysaccharide components of the fiber. Evidence of such a role comes from studies showing how the thermal softening of lignin can enable fiber swelling (Eriksson et al. 1991). On the other hand, Wang et al. (2021) showed that incorporation of lignin into a nanopaper structure by hot-pressing was able to decrease swelling by 94%. Likewise, high levels of modulus of elasticity were achieved when lignin was used as a binder in particleboards, along with a reduction in water swelling (Mathiasson and Kubat 1994).

Another factor that can contribute to lignin’s general negative effect on fiber swelling is its generally hydrophobic character, especially in its natural form (Notley and Norgren 2010; Borrega et al. 2020; Lisy et al. 2022). Ekeberg et al. (2006) showed that it is possible to separate kraft lignin into fractions having differing degrees of hydrophobicity, which indicates its heterogeneous nature in high-yield kraft fibers. As noted by Borrega et al. (2020), the water-wettability of lignin surface can be strongly affected by the often nano-particulate nature of lignin. Reduction in the pH, which occurs during the process of washing unbleached kraft pulp, can precipitate small lignin particles back onto the fiber surfaces, leading to nano-scale roughness. It has been shown that roughness on a very fine scale tends to amplify differences in wettability (Wenzel 1949; Hubbe et al. 2015).

The elastic modulus of lignin tends to decrease with increasing content of water, which appears to function as a plasticizing agent (Cousins 1976; Back and Salmén 1982; Eriksson et al. 1991). The Young’s modulus of different lignin specimens can range from about 2 to 7 GPa (Cousins 1976), depending on the moisture content and various changes that may occur during isolation or processing. These values are in the general range that has been predicted by quantum chemical calculations (Elder 2007). Back and Salmén (1982) showed that the softening temperature of native lignin decreased from about 200 to 115 °C as water content was increased in the range of zero to about 2.5%. Further increase in water content was found to merely remain external to the native lignin domains. However, it was found that derivatization of lignin, especially sulfonation, increased its swelling ability, which continued to depress the softening point of the lignin with much higher moisture contents of the modified lignin (Back and Salmén 1982).

Polysaccharides and hydrogen bonding issues

The mechanisms by which the polysaccharide components of woody materials can contribute to restraint of swelling often depend on hydrogen bonding in some way. It follows that such contributions will depend on such factors as the relative humidity or state of immersion in aqueous media. Assaf et al. (1944) used the term “avid” to describe the attraction of water molecules from the air onto and into dry cellulose, due to hydrogen bonding opportunities. Already at that time it was understood that the non-crystalline portions of cellulose, in addition to the hemicellulose, which is fully amorphous, take up most of that water. When forming nano-paper from NFC under ideal conditions, it is possible to achieve films that are sufficiently dense and defect-free that they can serve as superior barriers to oxygen and oils (Dufresne 2012). However, such barrier properties become progressively degraded with increasing relative humidity (Aulin et al. 2010).

A key principle regarding hydrogen bonds is that they generally have about the same energy content in a range of different circumstances. These include their presence in liquid water, in interactions between water molecules and cellulosic surfaces, and in air-dried or strongly dried cellulose material, including paper (Medronho et al. 2012). All such hydrogen bonds have an energy content of about 5 kcal/mole (21 kJ/mole).

Considerable progress has been achieved in differentiating between various states of water that may exist within cellulose-based materials over a range of moisture contents. These categories of water include bulk water, freezing bound water, and non-freezing bound water (Tait et al. 1972; Froix and Nelson 1975; Maloney et al. 1998a,b; Capitani et al. 1999; Park et al. 2007; Gao et al. 2015). In addition, some water molecules could, in principle, be present in cellulosic materials in the form of chemical complexes (Joubert et al. 1959); however, such a category might be regarded as the same thing as non-freezing bound water. Differential scanning calorimetry has revealed that some of the more tightly bound water requires a higher expenditure of heat to induce evaporation (Park et al. 2007). As noted by Paajanen et al. (2022), coalescence due to hydrogen bonding or other interactions between adjacent cellulose chains may sometimes have the effect of opening up larger domains of bulk water within a swollen cellulosic material. Under some drying conditions, increased amounts of the remaining water may become less mobile, thus showing lower rates of diffusion in comparison to bulk water (Salmén and Stevanic 2018). The unique behavior of bound water can include changes in mobility, as detected by NMR methods (Lindh et al. 2017).

Crystallinity as an impediment to swelling

The crystalline domains within cellulosic materials are understood to be non-swelling, and they also may restrain swelling of adjacent cellulosic structures (Roberts 1996; Solhi et al. 2023). Large differences in swelling have been reported when comparing amorphous vs. crystalline thin films of cellulose (Aulin et al. 2009). Ottesen and Syverud (2020) found a correlation between crystalline content and aqueous swelling of NFC. Paajanen et al. (2022) reported that the swelling within wood microfibril bundles involved increased thickness of water layers between the adjacent cellulosic fibrils.

Related observations have been made for various regenerated cellulose materials. Chaudemanche and Navard (2011) observed much greater swelling of outer parts of lyocell cellulose filaments. The observations were attributed to a suspected lower crystallinity of the surface layers. Isogai and Atalla (1998) observed different propensities of different cellulose samples for dissolution in cold NaOH solution. These differences were attributed so such factors as the types of crystals (cellulose I vs. cellulose II, for instance) and differences in cellulose molecular weight in some specimens. Those results were consistent with those of Kontturi et al. (2011), who found that certain highly amorphous thin cellulose films swelled dramatically in water and were soluble in dilute NaOH solution.

The insolubility of crystalline cellulose domains in water

Though many of the hydrogen bonds that hold cellulosic structures together can be readily replaced by similar bonds with water molecules, there is one area of exception. Crystalline domains within cellulose remain stubbornly resistant to the effects of water. There appear to be two contributing explanations for such behavior. One involves cooperative effects of multiple hydrogen bonds acting in an organized fashion. The other involves van der Waals forces and hydrophobic interaction. Cooperative effects due to multiple hydrogen bonds acting together can be understood based on statistics. A single hydrogen bond, in an aqueous system, will have an average lifetime of less than a picosecond (Rapaport 1983). In such a pattern of bonding, though the existence of any one bond may be transient, the chance that all of the bonds would release simultaneously is essentially zero (Hubbe et al. 2023). Such situations, especially when macromolecules are involved, easily can lead to trapped non-equilibrium states rather than perfect crystals or uniform gels (Claesson et al. 2005; Hubbe 2021). During biosynthesis of cellulose, as shown earlier in Fig. 6, it appears that the details are set up in such a way as to achieve a high degree of perfection of the crystalline domains, such that the result may be regarded as being close to a theoretical goal represented by defect-free crystals.

Another useful way to view the situation is through the lens of self-assembly (Hubbe et al. 2023). In the case of cellulosic microfibrils and other cellulosic surfaces, one can envision the possibility of zipper-like processes of hydrogen bond formation. Thus, Budtova and Navard (2016) suggested that such as zipping process can lead to insolubility of cellulose. Newman (2004) proposed that such a mechanism could account for irreversible loss of swelling ability when some cellulosic materials, such as kraft fibers, are dried. The concept was that the facing surfaces undergo a form of co-crystallization as the boundary between them dries out. This topic has been explored further by others (Pönni et al. 2012; Sjöstrand et al. 2023). Zhang et al. (2018) who proposed a mechanism by which cellulose dissolution can be promoted by relocation of cellulose chains so that they are “out of place” with respect to an otherwise possible zipper-like repair process. It was proposed that some cellulose solvent systems depend on such disorganizing effects to favor a net unraveling and dissolution.

As noted by Yamane et al. (2006), the cellulose crystal presents contrasting hydrophilic or hydrophobic character, depending on which face is being presented. This makes sense due to the fact that the hydrophilic –OH groups on cellulose face in the equatorial direction from the anhydroglucose repeating units of the polymer. Such an organization among adjacent chains means that in one direction the interactions are dominated by hydrogen bonding, whereas in the orthogonal direction, they are not. Rather, van der Waals forces and hydrophobic interactions are dominant in that direction. Such a dual-nature of how cellulose is held together has been discussed as a way to explain why it has been relatively difficult to find effective solvent systems for cellulose (Lindman et al. 2010; Glasser et al. 2012; Medronho et al. 2012; Budtova and Navard 2016). Reid (2017) proposed that van der Waals forces played a dominant role in the formation of certain CNC thin films; such a concept was helpful for explaining why observed swelling in that case has not been affected by salt ions. Hydrogen bonding has received dominant attention that hydrogen bonding has received as an explanation for many of the properties of cellulosic materials, including paper (Norgren et al. 2023). In that situation it is easy to overlook that the importance of van der Waals forces has been recognized for a long time (Warwicker and Wright 1967).

Sheet-like cellulose structures

A sheet-like nanostructure of water-swollen cellulose had been proposed by Stone and Scallan (1966, 1968) based on their observations of pore dimensions and the effects of drying. Such effects have been discussed by Nazhad (1994) in the context of accounting for loss of swelling ability when chemical pulp fibers are dried. In addition, electron micrographs of cross-sections of highly swollen cotton fibers under intermediate solvent conditions (ethylene diamine and ZnCl2) appear to show separated sheet-like lamellas at the nano scale (Aravindanath et al. 1992a,b). Such an outcome is consistent with the concept described above, in which the bonding within cellulose chains is dominated by water-resistant van der Waals forces in one of the orthogonal directions (Khazraji and Robert 2013). The cited authors backed up such a characterization with molecular modeling. Based on the models and findings mentioned above, one can envision swollen cellulosic materials as potentially having slit-like, readily expandable pores that each can fill with quite a lot of water, thus accounting for the ability of such fibers to take up much more water compared to their dry mass (Parham and Hebert 1980).

Accounting for High Swelling Systems

When technologists have the aim of dissolving cellulose, for instance with the use of specialized solvent systems, there are some factors that can become increasingly relevant. These include the molecular weight of the cellulose and its arrangements at the nano scale. Thus, treatments that break down the cellulose molecule can be important in cases where the goal of swelling the cellulose may include dissolution.

Depolymerization by chemical means

Thermodynamics dictates that dissolution of otherwise uniform polymeric materials will become more difficult with increasing molecular weight. Such effects can be expected to have parallel effects on swelling. The reason is that polymeric materials with increasing chain length gain less and less translational entropy upon their dissolution (Flory 1953; Medronho et al. 2012). The degrees of freedom within segments of a dissolved polymer are low due to the fact that each repeating unit is not free to move away from its immediate neighbors.

It has been shown that low molecular weight cellulose tends to be more readily soluble in specialized solvent systems, such as aqueous NaOH, which will be discussed in more detail later. Isogai (1996) found this in NMR studies. The cellulose employed was microcrystalline cellulose (MCC), in which the acid-destruction of non-crystalline zones of cellulose leads to decreased molecular weight. Since each cellulose chain needs to fit completely within a crystal domain of the remaining material after the acid hydrolysis, the molecular weight tends to have a limiting minimum value, i.e., a leveling-off value. Parallel tests carried out with cotton linters, using the same NaOH solvent system, showed low levels of solubility.

Certain treatment conditions may induce a combination of cellulose swelling and simultaneous breakdown in mass, such that the two effects can lead to confounding of effects. For instance, placement of kraft fibers in –butyl-3-methylimidazolium hydrogen sulfate solutions gave rise to simultaneous swelling and breakdown (Mao et al. 2016). The breakdown of molecular mass was especially noticed at the highest temperature considered for treatment in that work. Likewise, treatment of bagasse fibers with phosphoric acid and hydrogen peroxide led to intergranular swelling, at the same time as hydrolytic breakdown (Bai et al. 2022). The combination of swelling and breakdown appeared to have favorable effects on subsequent enzymolysis, leading to the recovery of glucose in that study.

Cellulase treatment

Enzymatic hydrolysis of cellulose also has been shown to promote swelling in some cases. Thus, cellulase was used as a pretreatment or post-treatment in the preparation of lignin-containing NFC (Bian et al. 2019). The pretreatment led to higher values of WRV, consistent with higher swelling. Those results suggest that the cellulase pretreatment had the effect of weakening the material, making it more susceptible to the mechanical fibrillation process. By contrast, post-treatment with the cellulase led to lower WRV than the control sample (no cellulase). It is known that post-treatment of fibrillated cellulose with cellulase may essentially dissolve and remove some of the slenderest fibrils, leading to faster drainage rates during paper formation (Gruber and Gelbrich 1997; Kim et al. 2006). Biodegradation of single crystals of cellulose has been shown to lead to swelling (Wang et al. 2012). This process was followed under aqueous conditions by means of atomic force microcopy. It was found that significant swelling took place only after substantial hydrolysis had been completed. Josefsson et al. (2008) showed that cellulase treatment could increase the swelling within cellulose ultrathin films. The effects were detected using the dissipation mode of the quartz crystal microbalance method (QCM-D). Thus, endoglucanases are able to create new end groups within the cellulose structure, making the material softer, with less restriction on swelling, whereas cellulases of cellobiohydrolase type lowered the swelling probably by degrading exposed structures (Josefsson et al. 2008).

Sublayer wrapping within cellulosic fibers

Some findings related to the layered structure of plant-derived cellulosic fibers already have been considered, but here the focus is on such findings specific to study of high levels of swelling. Some key findings of this type are included in Table 9. As shown, the restraining effects due to various cellulose layers can take many forms. In particular, the manner in which outer layers wrap around inner layers of cellulose fibrils in plant-derived cellulosic fibers tends to constrain fiber swelling in directions perpendicular to the axis. Ballooning, in fact, represents a partial failure in such a restraining action, whereby swelling within the thick S2 sublayer causes it to burst through the P and S1 layers, in a process that might be called herniation.

Table 9. Findings Related to How the Layered Structure of Cellulosic Fibers May Act to Restrain Swelling in the Presence of Cellulose Solvent Systems

Some revealing features of the ballooning phenomenon were shown in micrographs of NaOH-water swollen bleached fir kraft fibers (Cuissinat and Navard (2006). Representative images from that work, which had been obtained by optical microscopy, were shown earlier in simplified form, in Fig. 11.

Crosslinking effects

In addition to the layered structures that are either present within cellulosic fibers or that develop in the course of their dissolution, another effect that can restrict fiber swelling is crosslinking. Such effects are well known in the field of hydrogels (Chang and Zhang 2011; Salam et al. 2011; Hubbe et al. 2013; Ma et al. 2015). But some studies have shown that some of the same reagents can be utilized to constrain the extents of swelling of cellulosic fibers (Gamstedt 2016). For example, Almgren et al. (2010) treated bleached birch kraft fibers with butanetetracarboxylic acid. The treated fibers were heated for 15 minutes at 150 °C, which would be expected to lead to formation of di-ester crosslinks, probably via initial dehydration to form the five-carbon-ring-type anhydride groups. The crosslinked fibers were less susceptible to swelling. Similar findings, based on treatments with butanetetracarboxylic acid and citric acid, were found to decrease the swelling of paper (Caulfield 1994). Steps in the likely mechanism, for the case of citric acid, are shown in Fig. 15. Note that a higher level of heating is required for both of the reactions shown at the left of the figure, involving the formation of successive dicarboxylic anhydride rings. Once such rings are formed, the activation energy for reaction with an –OH group (for instance on cellulose or other polysaccharide) will be lower (Nguyen and Pham 2020), and somewhat less vigorous conditions can be employed to complete the ester formation. It is worth noting, however, that the final structure in the depicted reaction contains a free carboxylic acid group, which can be expected to contribute to a swelling effect when the pH is raised to near or above its pKa value.

Fig. 15. Steps in a reaction of citric acid that can lead to crosslinking of a poly-alcohol, such as cellulose, which is here represented by the groups ROH and R’OH

Solhi et al. (2023), in their review article, suggested that crosslinking agents can be used to preserve the integrity of chiral nematic films prepared from CNC. In such applications, the crosslinks would limit swelling and keep the film from excessive swelling, which could be a step towards the individual nanoparticles from going back into aqueous suspension and losing the striking optical affects often associated with chiral nematic character.

Solvent Systems

While the main focus of this review is on swelling, rather than dissolution, the effects of cellulose solvent systems can be regarded as an extreme manifestation of swelling. Many findings related to solvent systems already have been described in connection to other factors affecting swelling. Here the effects of some different solvent systems will be considered.

Swelling agents and solubility principles

Certain so-called swelling agents have been shown to induce swelling but not dissolution of cellulose. Highlights of such treatments and their results are given in Table 10.

Table 10. Swelling Agents for Cellulosic Fibers and their Reported Effects

As is evident from the work of Chen et al. (2015), it makes sense to consider solubility principles as a way to predict, or at least to qualitatively judge, the potential for different media to swell cellulose. A practical approach to such analysis, which has become widely used in science and industry, is that of Hansen (2007). Such analysis is grounded on three types of interaction between the solvent and various candidate substances to be added. As illustrated in Fig. 16, the first is the Hildebrand parameter, δD, which is the square-root of the cohesive energy density of the material (either the solvent or the candidate for dissolution). The second parameter, δP, has to do with the extent of polar nature. The third parameter, δH, has to do with the capability for hydrogen bonding. In principle, if there is a good match between the solubility parameters of a certain solid and a certain solvent, then one would anticipate favorable swelling and a higher chance of solubility. The hypothetical example shown in Fig. 16 represents a case where the solubility spheres of cellulose and prospective solvent do not overlap, meaning that solubility is not expected. Though, based on the analysis of Chen et al. (2015), it makes sense to regard the Hansen method as a good starting point, the approach has limitations. In particular, the Hansen method does not have a way to deal with differing affinity characteristics of different parts, ends, or faces of a given molecule or material. Such details tend to become more important when dealing with large molecules, e.g. polymers such as cellulose.

Fig. 16. Concept of solubility spheres, obtained by tests with probe liquids to determine the likely Hansen solubility parameters corresponding to solubility

In principle, mutual solubility can be promoted when one of the substances has Lewis acidic character and the other has Lewis basic character (van Oss et al. 2001; Jia et al. 2008). Van Oss et al. (2001) proposed that such interactions could make a contribution even in some aqueous systems. Such a term is not included in the Hansen system. Boluk (2005) came to the conclusion that Drago acid-base theory could help to account for hydrogen bonding interactions related to spruce kraft pulp and mechanical pulp. Pure cellulose, i.e., alpha cellulose, appeared to have a balanced acid-base character. El Seoud et al. (2008) used Gutmann’s donor and acceptor numbers to consider the solubility of a variety of different cellulose materials. The inclusion of these values, along with other parameters, allowed them to make more accurate predictions of cellulose solubility.

Fig. 17. Phase diagram for NaOH hydrates as a function of temperature and mass composition, also showing several eutectic points (blue circles)

NaOH as a cellulose swelling agent and solvent

The ideal conditions for dissolving cellulose in NaOH solution involve a concentration of about 7.6% and a temperature of about –5 °C (Cuissinat and Navard 2006). These rather specific conditions appear to result in the most favorable hydrated species of ions in a eutectic-like liquid (Budtova and Navard 2016). Figure 17 shows a NaOH-water diagram, which has been redrawn based on the work of Cohen-Adad et al. (1960). Note that a liquid phase is present for conditions of temperature and mass composition above the red line, and different NaOH hydrates will be represented at equilibrium in several solid phase mixtures, which are represented below the red line.

As mentioned earlier, it is likely that such hydrates tend to shift the position of recently unraveled cellulose chains. In principle, such a mechanism could inhibit the continual reassembly of cellulose crystallites, thereby allowing dissolution to proceed (Isogai and Atalla 1998; Zhang et al. 2018). Table 11 gives highlights from studies focusing on NaOH solution as a cellulose solvent or swelling agent.

Table 11. NaOH Solutions and the Swelling and Dissolution of Cellulose

NaOH with thiourea and or urea

It has been discovered that the solubility of cellulose in cold NaOH solution becomes even greater in the presence of optimized amounts of urea and thiourea (Yan et al. 2007; Zhang et al. 2013; Budtova and Navard 2016). The effect is supposedly attributable to weakening of hydrophobic interactions between the cellulose chains by the additives (Medronho et al. 2012). Zhang et al. (2013) found that adding thiourea to a NaOH system produced similar but more expanded balloons of swollen cellulose compared to NaOH alone.

Viscose

The viscose process, based on carbon disulfide treatment of the sodium form of cellulose, apparently does not involve true dissolution of cellulose, but rather a transient derivatization to cellulose xanthate (Lenz et al. 1993; Budtova and Navard 2016). The derivatization becomes reversed in the next step of the process wherein the viscous solution is converted back into cellulose. Steps in this chemical transformation are illustrated in Fig. 18. Lenz et al. (1993) found that the ability of viscose fibers to become fibrillated was correlated with the degree of orientation of its crystallites. In general, the drawing process employed while regenerating viscose fibers favors increased tensile strength, but also the ability to swell or delaminate perpendicular to the axis. Okugawa et al. (2020) reported that regenerated cellulose fibers are much more sensitive to water, in comparison to native cellulose fibers. This is consistent with the lack of, for instance, an S1 sublayer wrapping the structure. Okugawa et al. (2021) proposed that swelling takes place in the amorphous zones between axially oriented microfibrils of cellulose in a regenerated cellulose filament.

Fig. 18. Steps in the viscose process for preparation of regenerated cellulose (e.g. rayon) from cellulose (e.g. high-purity bleached kraft pulp or cotton)

NMMO

Partly due to environmental concerns about the viscose process, there has been a motivation to implement alternative ways to prepare regenerated cellulose products. The N-methyl-morpholine-N-oxide (NMMO) solvent process, yielding a regenerated cellulose product called lyocell, is regarded as a more eco-friendly alternative (Jiang et al. 2020). A likely mechanism to explain the ability of NMMO to dissolve cellulose is suggested in Fig. 19. As shown, it has been proposed that the NMMO replaces some of the key hydrogen bonding connections that connect the cellulose chains (Jiang et al. 2020). Though it is understood that a complete solubilization mechanism also likely will involve van der Waals interactions, that aspect is not considered in the depicted scheme.

Fig. 19. A mechanism that has been proposed for NMMO’s effectiveness for swelling and dissolving cellulose (redrawn based on scheme shown by Jiang et al. (2020)

Chaudemanche and Navard (2011) reimmersed lyocell fibers back into NMMO and studied the patterns and rates of swelling. Their results led them to the conclusion that the lyocell fibers were more porous in their outer layers, leading to high swelling in the NMMO solvent. However, Cuissinat and Navard (2006) described NMMO as a dissolution inhibitor. Cuissinat and Navard (2008) and Cuissinat et al. (2008) showed that NMMO treatment of a wide range of plant fibers consistently led to the ballooning effect described earlier. Lokhande (1978) attributed the effectiveness of NMMO as a swelling agent for cotton fibers to its combination of a relatively high Hildebrand parameter and a high polarity parameter, according to the Hansen (2007) approach. Sayyed et al. (2018) showed that dissolution of hardwood acid sulfite pulp in NMMO could be promoted by ultrasonication.

Phosphoric acid

Though phosphoric acid has been known as a cellulose solvent option since the 1930s (Budtova and Navard 2016), its usage tends to reduce the cellulose molecular weight due to acid hydrolysis. Possibly it is due to that attribute that the system has seldom been described in published articles. Y. H. P. Zhang et al. (2006) found that swelling of cellulose in phosphoric acid, followed by rinsing, rendered the material more susceptible to enzymatic hydrolysis. X. M. Zhang et al. (2018) showed that phosphoric acid was able to convert crystalline zones of cellulose to amorphous character and to impede the ongoing repair mechanism for the crystal zones. As was noted earlier, phosphoric acid has been used in combination with hydrogen peroxide as a swelling system to enhance enzymatic hydrolysis (Wang et al. 2020; Bai et al. 2022).

Ionic liquids

An ionic liquid can be defined as a salt that has a melting point below the boiling point of water.

Table 12. Ionic Liquids and the Swelling and Dissolution of Cellulose

In principle, ionic liquids represent combinations of positive and negative ions that have a mismatch in key properties such that they do not fit well together in a crystal form. Some of them have been shown to be able to swell or dissolve cellulose. Such findings are summarized in Table 12. The mechanism by which ionic liquids swell and dissolve cellulose has been reviewed (Li et al. 2018a; Khoo et al. 2021). Li et al. (2018a) proposed that the key to effectiveness of such solvent systems may lie the optimization of separate interactions with cellulose by the two types of ions. Khoo et al. (2021) emphasized the advantage of using an ionic liquid having a tendency to catalyze the breakdown of cellulose molecular weight; those authors were focused on an end goal of further converting the cellulose to biofuels. Molecular dynamic simulation work suggests that ionic liquids induce a progressive decrease in hydrogen bonding (Ishida 2020), which seems compatible with the zipper-like mechanism suggested by others for the action of cellulose solvents (Budtova and Navard 2016).

Kinetic Aspects of Swelling

In many current and potential applications, the rate at which the swelling of cellulosic materials occurs will affect success of a process. Time is required for such processes as diffusion, changes in conformation of soluble polymers, zipping and unzipping processes, and capillary flow.

Hysteresis in dry cellulosic materials

To begin with relatively dry cellulosic materials, researchers have found strong hysteresis effects when cellulosic paper takes up or loses water from or to the atmosphere (Ramarao and Chatterjee 1997; Nilsson et al. 1998; Hill 2009; Shrestha et al. 2017). Such effects are evidence that the material is not at a true equilibrium; rather, rate-determined processes govern the properties. For example, Shrestha et al. (2017) found that the amount of water present in a thin film of cellulose nanocrystals (CNC) at a specified relative humidity was often about 0.02 g water per g cellulose lower during the adsorption tests than during the subsequent desorption steps in the procedure. These findings are shown replotted in Fig. 20.

Fig. 20. Example of a hysteresis during adsorption of water vapor into a cellulose nanocrystal (CNC) film during increasing relative humidity, followed by desorption during decreasing relative humidity (redrawn from Shrestha et al. (2017)

Such findings support the view that water-cellulose interactions can depend on cooperative effects involving multiple hydrogen bonds. Whereas a single hydrogen bonding site on cellulose might be expected to be in pure equilibrium with changes in relative humidity, a sufficiently large group of such sites may require larger changes in relative humidity to prompt a change from local a “bonded area” condition to a condition in which cellulose-to-cellulose hydrogen bonds have been replaced in that region by cellulose-to-water hydrogen bonds, etc.

More recently, an approximate match to adsorption-desorption equilibria, as shown in Fig. 10, has been achieved by molecular dynamics simulation (Chen et al. 2019b). However, better fits were achieved when taking into account the coupled deformation of the soft material in the course of adsorption and desorption of water molecules (Chen et al. 2019a, 2020). This approach takes into account the adsorption-induced stress that contributes to swelling. Calculations based on either finite elements or analytical analysis based on a uniform strain assumption both were able to account for hysteresis loops in plots of water content of cellulose vs. relative humidity (Chen et al. 2020). Notably, the analyses just described did not consider capillarity, which was justified by an assumed presence of only nano-sized pores too small for development of a meniscus.

Capillary suction

Fibers and fibrous networks constitute a capillary system. When such networks are exposed to water or other wetting liquids, the extent and the rate at which the liquid enters will influence the extent and rate of swelling. A previous section discussed the change in Gibbs free energy for swelling, in which contributions stem from capillarity, i.e. both surface tension-area wetting work, and capillary action pressure-volume work, . The two quantities can be treated separately, but many studies dealing with fiber networks, such as paper, discuss such liquid-surface interactions together based on the Lucas-Washburn (L-W) relation, named after Lucas (1918) and Washburn (1921). An example of the type of situation considered by Lucas and Washburn is illustrated in Fig. 21.

Fig. 21. Simplified geometry to estimate rates of wetting into a porous solid under the assumptions of a known radius of smooth cylindrical pores having constant contact angle

 

Simplifying assumptions include the presence of cylindrical smooth pores of one radial size and one angle of contact with solution having a known Newtonian viscosity. Based on these assumptions, it is possible to calculate a contribution of capillary pressure to enter into a pore,

(5)

where γ is the interfacial tension, θ is the contact angle within the pore, as defined in Fig. 21, and R is the assumed radius representing the capillaries. This pressure is balanced by a Hagen-Poiseuille flow resistance term related to the viscosity of the solution,

(6)

where η is the dynamic viscosity of the solution, v is the average velocity of the fluid passing into the pore, l is the wetted length at the present time, and R is the same value already assumed for the pore radius. If one sets these two pressure terms equal to each other, then one can solve for the wetting length l at an elapsed time equal to t after immersing the dry porous material into the solution,

(7)

which gives a square-root time dependence. The L-W model does not consider effects of inertia and gravity, which can be of large importance. For pores in fibers and fiber networks, R and θ are not easily determined.

It is well known that the L-W model is too simplified to describe cellulosic fibers and network, but the model is still used with, in many cases, a good fit to determine the rate of capillary flow. Examples are by Kvick et al. (2017), in which also swelling of the pores during imbibition was included and by Salminen (1988), in which also other contributions are considered, such as the morphology and the effect of an advancing fluid front in pores due to fluid vapor diffusion, in front of the liquid.

Deviations from the Lucas-Washburn model are expected due to such factors as changes of contact angle as a function of time of contact, changes in morphology of the material due to swelling, adsorption or desorption of chemical contaminants, and various effects related to the roughness of real surfaces (Hubbe et al. 2015). Already Bosanquet (1923) introduced the inertial capillarity which gave a solution of two time-scales of the liquid flow, one of linear time dependence at short times and square-root dependence at longer times and which can explain the experimental observation that larger pores fill more easily than smaller pores (Quéré 1997; Gane 2005). The effect of gravity has been added, also presenting an analytical solution (Fries and Dreyer 2008).

The influence of the complex geometries in fibers and fiber networks (e.g. expanding or contracting pores, dislocations and obstructions) was pointed out by Kent and Lyne (1989). Variations in pore morphology have been experimentally verified (Senden et al. 2000). Studies have included various relevant well-defined geometries of the imbibed material (thin, thick, or deformable, such as swelling) of the flow situation (infinite or finite fluid reservoir, or drops) as well as if the fluid expands unidirectional, spherically or radially. To summarize their findings, the capillary rise proceeds as a function of for unidirectional capillary flow into a porous material (Marmur 2003). For an expanding pore, the rate is a function of or ) (Reyssat et al. 2008). The rate becomes a function of and goes as for flow imbibing spherically from a small orifice (Xiao et al. 2012) or from a fluid-saturated porous material into a capillary (Danino and Marmur 1994). The latter study also discusses flow from originally larger pores into smaller pores as an explanation for a lower flow rate than would be expected. For the special case of fluid drops into porous materials (Oko et al. 2014, 2016) was for an isotropic material but for paper. The slower rate in a fiber network is a result of a rough liquid front during capillary flow, which is a similar effect as the presence of obstructions for flow inside a porous network.

When fibers and fiber networks imbibe a wetting fluid by capillary suction, there are, as seen, a number of different effects that can both increase and decrease the capillary flow rate. This can in fact give a final result resembling a -dependence. This means that the Lucas-Washburn model may still apply. Knowing variability of the imbibing and swelling material is of large importance to further understanding capillary and swelling processes.

The relation between sorption in fibers, with accompanied fiber swelling, to the fluid uptake into the porous network between fibers, giving a swelling of the porous network, can be understood (Bristow 1986) based on experiments using oil as a non-swelling fluid and water as a swelling fluid. Zhmud et al. (2000) compared different theoretical approaches, corroborated with experiments, and discussed the important effect of liquid surface tension and surfactancy. Roberts et al. (2003) visualized the liquid advancing front using cryo-microscopy in papers. They described the liquid flow as being more along fiber surfaces and via fiber-fiber cross-points, rather than in the pores.

Several studies have started from Darcy’s law of fluid motion in porous media (Scheidegger 1974; Lyne 2002) in generalized form and in specific patterns of flow, e.g., flow in a pipe.

(8)

In Eq. 8, Q is the liquid flow rate, k is the permeability, A is the cross-sectional area, µ is the dynamic viscosity, L is the length penetrated by liquid, and ΔP is the pressure difference. This makes it possible to avoid the experimental issue of determining R and θ is avoided because the material properties are instead permeability and porosity. Permeability and capillary pressure can be estimated for an experimental system (Carman 1937; Marmur 2003). Approaches may also start from Navier-Stokes equations to include inertia and gravity effects. This may lead to equations that are too complex to be analytically solved, but instead numerically, or dimensional analyses may be used to conclude if some factors can be neglected. This was used in the case of small imbibing drops (Oko et al. 2014), in which both inertia and gravity terms were much smaller than the pressure term, and the complete equation was simplified to Darcy’s law.

Masoodi and Pillai (2010, 2011) used Darcy’s law to model the rates of wicking of liquid into paper-like materials. A new contribution of that work was to account for simultaneous swelling of the porous material in the course of wetting. Mark (2012) used Darcy’s law to model edge wicking into paperboard to also include the effect of an external (hydrostatic) pressure. Pejic et al. (2008) found that wicking of water into hemp fibers was increased by removal of lignin and some of the hemicellulose; those results are consistent with changes in wettability and pore sizes. Welo et al. (1952) used such principles to understand the WRV test, in which a centrifugal dewatering is used to estimate the amounts of water remaining in mesopores in the cell wall. Wohlert et al. (2022) emphasized the importance of capillary forces, acting during the evaporation of water from paper, in drawing cellulosic fiber surfaces into molecular contact so that hydrogen bonds could then form in the contact zones. As noted earlier, that aspect of capillary forces, in combination with drying, has been found to close up mesopores in a semi-irreversible manner, leading to lower levels of water sorption if and when the material is wetted again.

Diffusion-controlled processes

Diffusion processes often can be at the root of time-dependent phenomena related to the swelling of cellulosic materials (Ganji 2010). As noted by Alexandersson and Ristinmaa (2019, 2021), processes of adsorption and diffusion often control the rate at which water will invade and change the properties of cellulosic materials. Reid et al. (2017) and Shrestha et al. (2017) concluded that diffusion was the controlling process in the uptake of water into thin films of CNC. Conversely, water that is very closely associated with cellulose in tiny pores and gel structures can exhibit a slower rate of diffusion (Torstensen et al. 2022).

Table 13. Molecular Dynamics Simulation Studies of the Swelling of Cellulosic Materials

Mathematical modeling of swelling

Kinetic effects related to the swelling of cellulosic material often can be expressed in terms of mathematical models, but such models can be complex. For example, Masoodi et al. (2012) showed that a finite element model was able to account for rates of wicking into paper-like media. Their model allowed for concurrent swelling. Geffert et al. (2017) showed that swelling of cellulosic materials could be well described by a combination of two models, namely a generalized hydroscopicity model and a simple bounded growth model. Sayyed et al. (2021) noted that swelling of cellulose pulp in NMMO solution often follows a second-order rate expression. In other words, the rate slows with time, tending toward a final saturated condition. As noted by the authors, a deceleration in a process of swelling might be attributed to a plugging up of diffusion pathways, keeping the water or other liquid from being able to swell some isolated zones.

Molecular Dynamics Simulations

Further perspectives on swelling mechanisms can be gained by carrying out molecular dynamics simulation (MDS) studies. Such studies related to the swelling of cellulosic materials are highlighted in Table 13. As shown, in most of these studies, the MDS work was done to either confirm or shed additional light on the findings from other methods. Unlike those studies listed in Table 13, most publications dealing with MDS and cellulose do not provide information related to swelling, since such evidence is often absent or indirect. Studies listed in Table 13 represent some exceptions in which authors discussed their findings in relation to swelling.

APPLICATIONS OF SWELLING OF CELLULOSIC MATERIALS

This section highlights selected examples in which the swelling behavior of cellulosic materials, as well as its control, can have important effects in various applications. Rather than attempt to cover all potential applications, the objective here is to illustrate different ways in which the swelling of cellulosic materials can be important for industry and society.

Pulping of Cellulosic Fibers

In the course of preparing cellulosic fibers for papermaking, the dominant process involves breakdown and removal of the lignin component of wood, mainly using the kraft process (Fardim and Tikka 2011). The kraft process involves placing wood chips in a pressured vessel called a digester. The pulping liquor, which is a solution of sodium hydroxide and sodium hydrogen sulfide, needs to be able to diffuse into the wood chips, and this can be affected by swelling – either as a result of pretreatments or due to the alkaline nature of certain pulping liquors. The pretreatment of wood chips with alkaline aqueous solution has been shown to swell the wood and to promote permeability of pulping reagents (Minor and Springer 1993). The cited authors attributed part of the effect to saponification of ester groups in the wood. It has been proposed that the swelling of wood, brought about by a combination of temperature and immersion in water, is also a key to effective preparation of mechanical wood pulps (Fjellström et al. 2012). Such swelling helps the fibers to come apart with less damage on loss of fiber length. As discussed earlier, Eucalyptus wood with high content of calcium is pulped more slowly, which partly could be due to a decreased swelling (Vegunta et al. 2022),

Pulp Refining Optimization

Swelling is also important when using mechanical refining to get kraft fibers ready for formation into paper (Gharehkhani et al. 2015). As was noted earlier in a different context, a strong correlation has been found between the swollen nature of papermaking fibers, as represented by WRV, and their dry-strength properties (Jayme and Büttel 1968). The effect has been attributed to an expectation that a swollen fiber will be more conformable and better able to develop molecular contact between the cellulosic surfaces, leading to high levels of hydrogen bonding (Page 1969).

Absorbent Products

Swelling is a necessary process in cellulose-based absorbent products. Not only does the material need to swell in order to contain the absorbed fluid, but the material needs to be strong enough and have sufficiently strong capillary effects to hold onto that fluid. Ordinary cellulose can absorb 3.5 to 10 times its weight of water, depending on various details of composition and preparation (Aberson 1969; Parham and Hebert 1980; Ang 1991). Fluff pulp, which typically is comprised of bleached softwood kraft fibers, is widely used in disposable absorbent products, such as diapers (Parham and Hebert 1980). Such products also contain super-absorbent hydrogels, which optionally can be prepared with carboxymethylcellulose (CMC) as a component (Ganji et al. 2010; Chang et al. 2011; Hubbe et al. 2013; Ma et al. 2015). Moriwaki and Hanasaki (2023) noted that nanopaper sheets, composed of NFC, became immediately transformed to hydrogels upon contact with water. Sun et al. (2015) achieved related effects by first swelling cellulose at low temperature in NaOH and then very rapidly regenerating it by exposure to HCl solution; the resulting material acted like a hydrogel. As an alternative, Karlsson et al. (1998) achieved high levels of swelling by graft polymerization of regenerated cellulose fibers with acrylic acid.

Dissolving Pulp Manufacture

The ability of fibers to swell upon immersion in a cellulose solvent is a key criterion for high-quality dissolving grades of cellulose pulps, since this makes it easier for the cellulose to be chemically derivatized. This ability can be judged, for instance, by swelling in a dilute cupriethylenediamine (CUEN) aqueous solution (Arnoul-Jarriault et al. 2016). Such a solution is a diluted version of the medium that is often used for viscometric assessment of cellulose molecular mass. Pulps having a rapid ability to swell and dissolve in various solvent systems are known as “reactive”. It has been shown that reactivity can be increased by such treatments as TEMPO-mediated oxidation of the fibers (Gehmayr et al. 2012). Part of the concept of reactivity can be addressed as accessibility and refers to the ease of which reactive sites of the molecule can be reached by reaction chemicals and/or solvents and depends on several physical and chemical properties of the pulp (Li et al. 2018b).

Bioenergy

Swelling ability also can be important when using enzyme-based strategies to prepare biofuels. For example, it has been shown that swelling of hemp fibers with 5% NaOH solution promoted subsequent enzymatic saccharification (George et al. 2015). Related studies using different swelling agents have been reported (Bendzalova et al. 1996). Once sucrose has been prepared from cellulose in such a way, it can then be converted to biofuels such as ethanol and butanol by fermentation. Contrary results have been reported, in which increased swelling of pulp xylan led to less enzymatic hydrolysis (Buchert et al. 1993). The latter results suggest that certain enzymes have evolved to favor their action on dense cellulosic material, including crystalline cellulose rather than artificially swollen versions. However, as noted earlier, it has been shown that the loss of swelling, due to drying of cellulosic material, generally has a negative effect on rates of enzymatic saccharification (Luo and Zhu 2011; Duan et al. 2015).

CLOSING COMMENTS

This review article has focused on factors affecting the swelling of cellulosic fibers, published findings related to such effects, mechanistic explanations, and some published examples of current or potential applications. The word “swelling,” as used in the related literature, was found to cover a wide range of phenomena. As illustrated in Fig. 22, these phenomena can range from the slight dimensional changes associated with changes in relative humidity. Successively greater swelling occurs when the cellulosic material is immersed in water, and even more so when immersed in a select group of agents able to strongly swell and even dissolve cellulose. The mechanistic explanations for forces encouraging swelling range from adsorption, to osmotic pressure, and to disruption of the mechanical structures that ordinarily would limit swelling, such as the outer layers of plant-derived cellulose fibers.

Fig. 22. Envisioning cellulose swelling as a continuum of overlapping processes, using some earlier figures to provide examples

Some principles that were shown, in this review, to help explain some of the cited findings, can be summarized as follows:

  • Crosswise swelling of cellulose fibers and other cellulose-based structures is typically several times greater than axial swelling; such behavior is consistent with a model in which slit-like spaces between sheets of cellulosic material can open up and accommodate water, depending on such factors as osmotic pressure and the application of mechanical shearing.
  • Swelling of cellulosic fibers in water is positively influenced by the polar nature of cellulose, as well as hemicellulose, and by osmotic effects that arise from the presence of ionizable groups such as -COOH.
  • Cellulose’s non-solubility in water, which clearly acts to restrain swelling in many cases, is related to its typically high level of crystallinity, which can further by attributed to a dense and regular pattern of hydrogen bonding, in combination with hydrophobic (van der Waals) association acting in a perpendicular direction.
  • The microfibril angles within different layers of the cell walls of plant fibers, as well as the presence of lignin and lignin-polysaccharide complexes all play important roles relative to the restraint of swelling.
  • Very high levels of swelling, even leading to dissolution of cellulose, can be achieved by a variety of cellulose solvent systems. At least some of these systems simultaneously attack the hydrogen bonding within the cellulose domains and also weaken the van der Waals associations between the non-polar faces of adjacent cellulose chains or sheets.

The study of aspects of the swelling of cellulosic materials continues to provide opportunities for new researchers. Because the field is diverse in many ways, it is hard to place limits on the kinds of research that will be useful in the years ahead. On the one hand, there will be a need for fundamental research. Although such theories as osmotic pressure and its controlling factors are by now well established, the general theories often do not take into account the structural details of real cellulosic fibers. The need for developmental scientists and engineers, working to implement aspects of research in industry, can be expected to grow. An expected growing focus on photosynthetically renewable materials is likely to promote more emphasis on the practical usage of cellulose-based materials in a wide range of products, ranging from absorbents to enzymatically produced biofuels, to packages, and to medical devices. In many applications, greater swelling may be the goal, whereas in others it may be an advantage to limit swelling. As has been shown by the articles cited in this review, by varying the manner in which cellulosic fibers are prepared and by controlling the environmental to which they are exposed, it is possible to achieve a wide range of outcomes relative to swelling and its restraint.

ACKNOWLEDGEMENTS

The authors are thankful for three experts who inspected an earlier version of this article and offered useful suggestions: Thaddeus Maloney (Aalto University, School of Chemical Engineering, Department of Bioproducts and Biosystems, Aalto 00076, Finland), Jacob Wohlert (KTH Royal Inst Technol, Dept Fibre & Polymer Technol, School of Engineering Science Chemistry Biotechnology and Health, S-10044, Stockholm, Sweden), and Chi Zhang (School of Flexible Electronics, Sun Yat-sen University, 518107 Shenzhen, China). The work of Martin A. Hubbe related to papermaking science is supported by the Buckman Foundation.

REFERENCES CITED

Aarne, N., Kontturi, E., and Laine, J. (2012). “Influence of adsorbed polyelectrolytes on pore size distribution of a water-swollen biomaterial,” Soft Matter 8(17), 4740-4749. DOI: 10.1039/c2sm07268h

Aberson, G. (1969). “The water absorbency of pads of dry, unbonded fibre,” TAPPI STAP 8, 282-305.

Ahola, S., Salmi, J., Johansson, L. S., Laine, J., and Österberg, M. (2008). “Model films from native cellulose nanofibrils. Preparation, swelling, and surface interactions,” Biomacromol. 9(4), 1273-1282. DOI: 10.1021/bm701317k

Alanko, K., Paulapuro, H., and Stenius, P. (1995). “Recyclability of thermomechanical pulp fibers,” Paperi ja Puu 77(5), 315.

Alexandersson, M., and Ristinmaa, M. (2018). “Modelling multiphase transport in deformable cellulose based materials exhibiting internal mass exchange and swelling,” Int. J. Eng. Sci. 128, 101-126. DOI: 10.1016/j.ijengsci.2018.03.013

Alexandersson, M., and Ristinmaa, M. (2019). “Multiphase transport model of swelling cellulose based materials with variable hydrophobicity,” Int. J. Eng. Sci. 141, 112-140. DOI: 10.1016/j.ijengsci.2019.05.010

Alexandersson, M., and Ristinmaa, M. (2021). “Coupled heat, mass and momentum transport in swelling cellulose based materials with application to retorting of paperboard packages,” Appl. Math. Mod. 92, 848-883. DOI: 10.1016/j.apm.2020.11.041

Alince, B. (1991). “Comments on porosity of swollen pulp fibers analyzed by solute exclusion,” TAPPI J. 74(11), 200-202.

Alince, B. (2002). “Porosity of swollen pulp fibers revisited,” Nordic Pulp Paper Res. J. 17(1), 71-73. DOI: 10.3183/npprj-2002-17-01-p071-073

Alince, B., and van de Ven, T. G. M. (1997). “Porosity of swollen pulp fibers evaluated by polymer adsorption,” in: The Fundamentals of Papermaking Materials, Transactions of the 11th Fundamental Research Symposium, Cambridge, Sept. 1997, C. F. Baker (ed.), FRC, Manchester, Vol. 2, pp. 771-788. DOI: 10.15376/frc.1997.2.771

Alkorta, I., Garbisu, C., Llama, M. J., and Serra, J. L. (1998). “Industrial applications of pectic enzymes: A review,” Process Biochemistry 33(1), 21-28. DOI: 10.1016/S0032-9592(97)00046-0

Almgren, K. M., Gamstedt, E. K., and Varna, J. (2010). “Contribution of wood fiber hydroexpansion to moisture induced thickness swelling of composite plates,” Polym. Compos. 31(5), 762-771. DOI: 10.1002/pc.20858

Andersson, K., and Lindgren, E. (1996). “Important properties of colloidal silica in microparticulate system,” Nordic Pulp Paper Res. J. 11(1), 15-21, 57. DOI: 10.3183/npprj-1996-11-01-p015-021

Andersson, K., Sandström, A., Ström, K., and Barla, P. (1986). “The use of cationic starch and colloidal silica to improve the drainage characteristics of kraft pulps,” Nordic Pulp Paper Res. J. 1(2), 26-30. DOI: 10.3183/npprj-1986-01-02-p026-030

Andersson, S., Serimaa, R., Torkkeli, M., Paakkari, T., Saranpaa, P., and Pesonen, E. (2000). “Microfibril angle of Norway spruce [Picea abies (L.) Karst.] compression wood: Comparison of measuring techniques,” J. Wood Sci. 46(5), 343-349. DOI: 10.1007/BF00776394

Ang, J. F. (1991). “Water-retention capacity and viscosity effect of powdered cellulose,” J. Food Sci. 56(6), 682-1684. DOI: 10.1111/j.1365-2621.1991.tb08670.x

Araujo, C., Freire, C. S. R., Nolasco, M. M., Ribeiro-Claro, P. J. A., Rudic, S., Silvestre, A. J. D., and Vaz, P. D. (2018). “Hydrogen bond dynamics of cellulose through inelastic neutron scattering spectroscopy,” Biomacromol. 19, 1305-1313. DOI: 10.1021/acs.biomac.8b00110

Aravindanath, S., Iyer, P. B., and Sreenivasan, S. (1992a). “Layer morphology and its relation to swelling and structure. 1. Cotton fibers treated in alkali-metal hydroxides,” J. Appl. Polym. Sci. 46(12), 2239-2244. DOI: 10.1002/app.1992.070461222

Aravindanath, S., Iyer, P. B., and Sreenivasan, S. (1992b). “Layer morphology and its relation to swelling and structure. 2. Cotton fibers treated with ethylenediamine and zinc-chloride,” J. Appl. Polym. Sci. 46(12), 245-2250. DOI: 10.1002/app.1992.070461223

Arnoul-Jarriault, B., Passas, R., Lachenal, D., and Chirat, C. (2016). “Characterization of dissolving pulp by fibre swelling in dilute cupriethylenediamine (CUEN) solution in a MorFi analyser,” Holzforschung 70(7), 611-617. DOI: 10.1515/hf-2015-0167

Arola, S., and Linder, M. B. (2016). “Binding of cellulose binding modules reveal differences between cellulose substrates,” Scientific Reports 6(1), 1-9. DOI: 10.1038/srep35358

Assaf, A. G., Haas, R. H., and Purves, C. B. (1944). “A new interpretation of the cellulose-water adsorption isotherm and data concerning the effect of swelling and drying on the colloidal surface of cellulose,” J. Amer. Chem. Soc. 66, 66-73. DOI: 10.1021/ja01229a020

Aulin, C., Ahola, S., Josefsson, P., Nishino, T., Hirose, Y., Österberg, M., and Wågberg, L. (2009). “Nanoscale cellulose films with different crystallinities and mesostructures – Their surface properties and interaction with water,” Langmuir 25(13), 7675-7685. DOI: 10.1021/la900323n

Aulin, C., Gällstedt, M., and Lindström, T. (2010). “Oxygen and oil barrier properties of microfibrillated cellulose films and coatings,” Cellulose 17, 559-574. DOI 10.1007/s10570-009-9393-y

Ayoub, A., Venditti, R. A., Pawlak, J. J., Salam, A., and Hubbe, M. A. (2013). “Novel hemicellulose-chitosan biosorbent for water desalination and heavy metal removal,” ACS Sustainable Chem. Eng. 1(9), 1102-1109. DOI: 10.1021/sc300166m

Back, E. L., and Salmén, N. L. (1982). “Glass transitions of wood components hold implications for molding and pulping processes,” Tappi 65(7), 107-110.

Bai, F. T., Dong, T. T., Zhou, Z., Chen, W., Cai, C. C., and Li, X. S. (2022). “Enhancing for bagasse enzymolysis via intercrystalline swelling of cellulose combined with hydrolysis and oxidation,” Polymers 14(17), article 3587. DOI: 10.3390/polym14173587

Bardestani, R., Patience, G. S., and Kaliaguine, S. (2019). “Experimental methods in chemical engineering: Specific surface area and pore size distribution measurements-BET, BJH, and DFT,” Can. J. Chem. Eng. 97(11), 2781-2791. DOI: 10.1002/cjce.23632

Barnett, J. R., and Bonham, V. A. (2004). “Cellulose microfibril angle in the cell wall of wood fibres,” Biol. Rev. 79(2), 461-472. DOI: 10.1017/S1464793103006377

Barrett, E. P., Joyner, L. G., and Halenda, P. P. (1951). “The determination of pore volume and area distributions in porous substances. 1. Computations from nitrogen isotherms,” J. Amer. Chem. Soc. 73, 373-380. DOI: 10.1021/ja01145a126

Bendzalova, M., Pekarovicova, A., Kokta, B. V., and Chen, R. (1996). “Accessibility of swollen cellulosic fibers,” Cellulose Chem. Technol. 30, 19-32.

Benselfelt, T., Ciftci, G. C., Wågberg, L., Wohlert, J., and Hamedi, M. M. (2024). “Entropy drives interpolymer association in water: Insights into molecular mechanisms,” Langmuir 40(13), 6718-6729. DOI: 10.1021/acs.langmuir.3c02978

Benselfelt, T., Kummer, N., Nordenström, M., Fall, A. B., Nyström, G., and Wågberg, L. (2023). “The colloidal properties of nanocellulose,” ChemSusChem. DOI: 10.1002/cssc.202201955

Benselfelt, T., Nordenström, M., Hamedi, M. M., and Wågberg, L. (2019). “Ion-induced assemblies of highly anisotropic nanoparticles are governed by ion-ion correlation and specific ion effects,” Nanoscale 11(8), 3514-3520. DOI: 10.1039/c8nr10175b

Bergenstråhle, M., Wohlert, J., Himmel, M. E., and Brady, J. W. (2010). “Simulation studies of the insolubility of cellulose,” Carbohydr. Res. 345(14), 2060-2066. DOI: 10.1016/j.carres.2010.06.017

Berglund, J., Mikkelsen, D., Flanagan, B. M., Dhital, S., Gaunitz, S., Henriksson, G., Lindström, M. E., Yakubov, G. E., Gidley, M. J., and Vilaplana, F. (2020). “Wood hemicelluloses exert distinct biomechanical contributions to cellulose fibrillar networks,” Nature Communications 11(1), article 4692. DOI: 10.1038/s41467-020-18390-z

Berthold, J., Rinaudo, M., and Salmen, L. (1996). “Association of water to polar groups; Estimations by an adsorption model for ligno-cellulosic materials,” Colloids Surf. A – Physicochem. Eng. Aspects 112(2-3), 117-129. DOI: 10.1016/0927-7757(95)03419-6

Berthold, J., and Salmén, L. (1997a). “Inverse size exclusion chromatography (ISEC) for determining the relative pore size distribution of wood pulps,” Holzforschung 51(4), 361-368. DOI: 10.1515/hfsg.1997.51.4.361

Berthold, J., and Salmén, L. (1997b). “Effects of mechanical and chemical treatments on the pore-size distribution in wood pulps examined by inverse size-exclusion chromatography,” J. Pulp Paper Sci. 23(6), J245-J253.

Bian, H. Y., Dong, M. L., Chen, L. D., Zhou, X. L., Ni, S. Z., Fang, G. G., and Dai, H. Q. (2019). “Comparison of mixed enzymatic pretreatment and post-treatment for enhancing the cellulose nanofibrillation efficiency,” Biores. Technol. 293, article 122171. DOI: 10.1016/j.biortech.2019.122171

Boluk, Y. (2005). “Acid-base interactions and swelling of cellulose fibers in organic liquids,” Cellulose 12(6), 577-593. DOI: 10.1007/s10570-005-9004-5

Borrega, M., Paarnila, S., Greca, L. G., Jaaskelainen, A. S., Ohra-aho, T., Rojas, O. J., and Tamminen, T. (2020). “Morphological and wettability properties of thin coating films produced from technical lignins,” Langmuir 36(33), 9675-9684. DOI: 10.1021/acs.langmuir.0c00826

Borrero-Lopez, A. M., Greca, L. G., Rojas, O. J., and Tardy, B. L. (2023). “Controlling superstructure formation and macro-scale adhesion via confined evaporation of cellulose nanocrystals,” Cellulose 30, 741-751. DOI: 10.1007/s10570-022-04937-4

Bosanquet, C. H. (1923). “On the flow of liquids into capillary tubes,” Phil. Mag. 1923, S6 45(267), 525-531. DOI: 10.1080/14786442308634144

Botkova, M., Suty, S., Jablonsky, M., Kucerkova, L., and Vrska, M. (2013). “Monitoring of kraft pulps swelling in water,” Cellulose Chem. Technol. 47(1-2), 95-102.

Brännvall, E., Larsson, P. T., and Stevanic, J. S. (2021). “Changes in the cellulose fiber wall supramolecular structure during the initial stages of chemical treatments of wood evaluated by NMR and X-ray scattering,” Cellulose 28, 3951-3965. DOI: 10.1007/s10570-021-03790-1

Brestkin, I. V., and Frenkel, S. Y. (1971). “On cellulose degradation in heterogeneous reactions,” European Polymer Journal 7(5), 543-547. DOI: 10.1016/0014-3057(71)90085-1

Brett, C. T., Baydoun, E. H., and Abdel-Massih, R. M. (2005). “Pectin–xyloglucan linkages in type I primary cell walls of plants,” Plant Biosystems – An International Journal Dealing with all Aspects of Plant Biology 139(1), 54-59. DOI: 10.1080/11263500500056732

Bristow, J. A. (1986). “The pore structure and the sorption of liquids,” in: Paper – Structure and Properties, J. A. Bristow and P. Kolseth (eds.), Marcel Dekker Inc.

Brodin, F. W., and Theliander, H. (2012). “Absorbent materials based on kraft pulp: Preparation and material characterization,” BioResources 7(2), 1666-1683. DOI: 10.15376/biores.7.2.1666-1683

Brown, R. M., and Saxena, I. M. (2000). “Cellulose biosynthesis: A model for understanding the assembly of biopolymers,” Plant Physiol. Biochem. 38, 57-67. DOI: 10.1016/S0981-9428(00)00168-6

Brunauer, S., Emmett, P. H., and Teller, E. (1938). “Adsorption of gases in multimolecular layers,” J. Amer. Chem. Soc. 60, 309-319. DOI: 10.1021/ja01269a023

Buchert, J., Tenkanen, M., Pitkanen, M., and Viikari, L. (1993). “Role of surface-charge and swelling on the action of xylanases on birch kraft pulp,” TAPPI J. 76(11), 131-135.

Budtova, T., and Navard, P. (2016). “Cellulose in NaOH-water based solvents: A review,” Cellulose 23(1), 5-55. DOI: 10.1007/s10570-015-0779-8

Bui, H. M., Lenninger, M., Manian, V. P., Abu-Rous, M., Schimper, C. B., Schuster, K. C., and Bechtold, T. (2007). “Treatment in swelling solutions modifying cellulose fiber reactivity – Part 2: Accessibility and reactivity,” Macromol. Symp. 262, 50-64. DOI: 10.1002/masy.200850206

Burchard, W. (2003). “Solubility and solution structure of cellulose derivatives,” Cellulose 10(3), 213-225. DOI: 10.1023/A:1025160620576

Burgess, R. R. (2018). “A brief practical review of size exclusion chromatography: Rules of thumb, limitations, and troubleshooting,” Protein Expression Purif. 150, 81-85. DOI: 10.1016/j.pep.2018.05.007

Caffall, K. H., and Mohnen, D. (2009). “The structure, function, and biosynthesis of plant cell wall pectic polysaccharides,” Carbohydrate Research 344(14), 1879-1900. DOI: 10.1016/j.carres.2009.05.021

Cao, L., Lu, W., Mata, A., Nishinari, K., and Fang, Y. (2020). “Egg-box model-based gelation of alginate and pectin: A review,” Carbohydrate Polymers 242, article 116389. DOI: 10.1016/j.carbpol.2020.116389

Cao, L. M., Zhu, J. T., Deng, B. J., Zeng, F. Y., Wang, S. S., Ma, Y., Qin, C. R., and Yao, S. Q. (2022). “Efficient swelling and mercerization of bagasse fiber by freeze-thaw-assisted alkali treatment,” Frontiers Energy Res. 10, article 851543. DOI: 10.3389/fenrg.2022.851543

Capitani, D., Emanuele, M. C., Bella, J., Segre, A. L., Attanasio, D., Focher, B., and Capretti, G. (1999). “1H NMR relaxation study of cellulose and water interaction in paper,” TAPPI J. 82(9), 117-124.

Carlsson, G., Kolseth, P., and Lindström, T. (1983). “Poly-electrolyte swelling behavior of chlorite delignified spruce wood fibers,” Wood Sci. Technol. 17(1), 69-73. DOI: 10.1007/BF00351833

Carman, P. C. (1937). “Fluid flow through granular beds,” Trans. Int. Chem. Eng. 15, 150-166.

Caulfield, D. F. (1994). “Ester cross-linking to improve wet performance of paper using multifunctional carboxylic-acids, butanetetracarboxylic and citric-acid,” TAPPI J. 77(3), 205-212.

Chang, C. Y., He, M., Zhou, J. P., and Zhang, L. N. (2011). “Swelling behaviors of pH- and salt-responsive cellulose-based hydrogels,” Macromol. 44(6), 1642-1648. DOI: 10.1021/ma102801f

Chang, C. Y., and Zhang, L. N. (2011). “Cellulose-based hydrogels: Present status and application prospects,” Carbohyd. Polym. 84(1), 40-53. DOI: 10.1016/j.carbpol.2010.12.023

Chaudemanche, C., and Navard, P. (2011). “Swelling and dissolution mechanisms of regenerated Lyocell cellulose fibers,” Cellulose 18(1), 1-15. DOI: 10.1007/s10570-010-9460-4

Chen, F., Sawada, D., Hummel, M., Sixta, H., and Budtova, T. (2020). “Swelling and dissolution kinetics of natural and man-made cellulose fibers in solvent power tuned ionic liquid,” Cellulose 27(13), 7399-7415. DOI: 10.1007/s10570-020-03312-5

Chen, G. Y., Yu, H. Y., Zhang, C. H., Zhou, Y., and Yao, J. M. (2016). “A universal route for the simultaneous extraction and functionalization of cellulose nanocrystals from industrial and agricultural celluloses,” J. Nanoparticle Res. 18(2), article no. 48. DOI: 10.1007/s11051-016-3355-8

Chen, J., Heitmann, J. A., and Hubbe, M. A. (2003). “Dependency of polyelectrolyte complex stoichiometry on the order of addition. 1. Effect of salt concentration during streaming current titrations with strong poly-acid and poly-base,” Colloids Surf. A 223(1-3), 215-230. DOI: 10.1016/S0927-7757(03)00222-X

Chen, L. Y., Wang, B. J., Chen, J. G., Ruan, X. H., and Yang, Y. Q. (2015). “Comprehensive study on cellulose swelling for completely recyclable nonaqueous reactive dyeing,” Indust. Eng. Chem. Res. 54(9), 2439-2446. DOI: 10.1021/ie504677z

Chen, M.-Y., Coasne, B., Derome, D., and Carmeliet, J. (2019a). “Coupling of sorption and deformation in soft nanoporous polymers: Molecular simulation and poromechanics,” J. Mechan. Phys. Solids 137, article 103830. DOI: 10.1016/j.jmps.2019.103830

Chen, M.-Y., Zhang, C., Shomali, A., Coasne, B., Carbeliet, J., and Derome, D. (2019b). “Wood-moisture relationships studied with molecular simulations: Methodological guidelines,” Forests 10, article 628. DOI: 10.3390/f10080628

Chen, M.-Y., Coasne, B., Guyer, R., Derome, D., and Carmeliet, J. (2020). “A poromechanical model for sorption hysteresis in nanoporous polymers,” J. Phys. Chem. B 124(39), 8690-8703. DOI: 10.1021/acs.jpcb.0c04477

Chen, P., Wohlert, J., Berglund, L., and Furo, I. (2022). “Water as an intrinsic structural element in cellulose fibril aggregates,” J. Phys. Chem. Lett. 13, 5424-5430. DOI: 10.1021/acs.jpclett.2c00781

Cheng, Q. Z., Wang, J. X., McNeel, J. F., and Jacobson, P. M. (2010). “Water retention value measurements of cellulosic materials using a centrifuge technique,” BioResources 5(3), 1945-1954. DOI: 10.15376/biores.5.3.1945-1954

Choi, K. H., Kim, A. R., and Cho, B. U. (2016). “Effects of alkali swelling and beating treatments on properties of kraft pulp fibers,” BioResources 11(2), 3769-3782. DOI: 10.15376/biores.11.2.3769-3782

Choi, K. H., Kim, A. R., and Cho, B. U. (2018). “Manufacture of high bulk paper using alkali swollen kraft pulp,” Nordic Pulp Paper Res. J. 33(3), 503-511. DOI: 10.1515/npprj-2018-3059

Claesson, P. M., Poptoshev, E., Blomberg, E., and Dedinaite, A. (2005). “Polyelectrolyte-mediated surface interactions,” Advan. Colloid Interface Sci. 114, 173-187. DOI: 10.1016/j.cis.2004.09.008

Cohen-Adad, R., Tranquard, A., Peronne, R., Negri, P., and Rollet, A. P. (1960). “Le système eau-hydroxyde de sodium,” Comptes Rendus de l’Académie des Sciences 251, 2035-2037, as reported by Budtova and Navard (2016).

Cosgrove, D. J. (2014). “Re-constructing our models of cellulose and primary cell wall assembly,” Current Opinion in Plant Biology 22, 122-131. DOI: 10.1016/j.pbi.2014.11.001

Cousins, W. J. (1976). “Elastic modulus of lignin as related to moisture content,” Wood Sci. Technol. 10(1), 9-17. DOI: 10.1007/BF00376380

Crawshaw, J., Bras, W., Mant, G. R., and Cameron, R. E. (2002). “Simultaneous SAXS and WAXS investigations of changes in native cellulose fiber microstructure on swelling in aqueous sodium hydroxide,” J. Appl. Polymer Sci. 83(6), 1209-1218. DOI: 10.1002/app.2287

Cresswell, R., Dupree, R., Brown, S. P., Pereira, C. S., Skaf, M. S., Sorieul, M., Dupree, P., and Hill, S. (2021). “Importance of water in maintaining softwood secondary cell wall nanostructure,” Biomacromol. 22, 4669-4680. DOI: 10.1021/acs.biomac.1c00937

Cuissinat, C., and Navard, P. (2006). “Swelling and dissolution of cellulose Part II: Free floating cotton and wood fibres in NaOH-water-additives systems,” Macromol. Symp. 244, 19-30. DOI: 10.1002/masy.200651202

Cuissinat, C., and Navard, P. (2008). “Swelling and dissolution of cellulose, Part III: Plant fibres in aqueous systems,” Cellulose 15(1), 67-74. DOI: 10.1007/s10570-007-9158-4

Cuissinat, C., Navard, P., and Heinze, T. (2008). “Swelling and dissolution of cellulose, Part V: Cellulose derivatives fibres in aqueous systems and ionic liquids,” Cellulose 15(1), 75-80. DOI: 10.1007/s10570-007-9159-3

Daly, D. C., Cameron, K. M., and Stevenson, D. W. (2001). “Plant systematics in the age of genomics,” Plant Physiology 127(4), 1328-1333. DOI: 10.1104/pp.010788

Danino, D., and Marmur, A. (1994). “Radial capillary penetration into paper – Limited and unlimited liquid reservoirs,” Journal Colloid Interface Science 166(1), 245-250. DOI: 10.1006/jcis.1994.1290

Dargaville, B. L., and Hutmacher, D. W. (2022). “Water as the often neglected medium at the interface between materials and biology,” Nature Commun. 13, article 4222. DOI: 10.1038/s41467-022-31889-x

Dautzenberg, H., and Jaeger, W. (2002). “Effect of charge density on the formation and salt stability of polyelectrolyte complexes,” Macromol. Chem. Phys. 203(14), 2095-2102. DOI: 10.1002/1521-3935(200210)203:14<2095::AID-MACP2095>3.0.CO;2-9

Debnath, M., Sarder, R., Pal, L., and Hubbe, M. A. (2022). “Molded pulp products for sustainable packaging: Production rate challenges and product opportunities,” BioResources 17(2), 3810-3870. DOI: 10.15376/biores.17.2.Debnath

Derjaguin, B. V., and Landau, L. D. (1941). “Theory of the stability of strongly charged lyophobic sols and the adhesion of strongly charged particles in solution of electrolytes,” Acta Pysicochim. URSS 14, 633-662.

Dimic-Misic, K., Rantanen, J., Maloney, T. C., and Gane, P. A. C. (2016). “Gel structure phase behavior in micro nanofibrillated cellulose containing in situ precipitated calcium carbonate,” J. Appl. Polymer Sci. 133(22), article no. 43486. DOI: 10.1002/app.43486

Dinand, E., Vignon, M., Chanzy, H., and Heux, L. (2002). “Mercerization of primary wall cellulose and its implication for the conversion of cellulose I→ cellulose II,” Cellulose 9, 7-18. DOI: 10.1023/A:1015877021688

Djahedi, C., Bergenstråhle-Wohlert, M., Berglund, L. A., and Wohlert, J. (2016). “Role of hydrogen bonding in cellulose deformation: The leverage effect analyzed by molecular modeling,” Cellulose 23(4), 2315-2323. DOI: 10.1007/s10570-016-0968-0

Doblin, M. S., Kurek, I., Jacob-Wilk, D., and Delmer, D. P. (2002). “Cellulose biosynthesis in plants: From genes to rosettes,” Plant and cell physiology 43(12), 1407-1420. DOI: 10.1093/pcp/pcf164

Donaldson, L. (2008). “Microfibril angle: Measurement, variation and relationships – A review,” IAWA J. 29(4), 345-386. DOI: 10.1163/22941932-90000192

Duan, C., Long, Y. D., Li, J. G., Ma, X. J., and Ni, Y. H. (2015). “Changes of cellulose accessibility to cellulase due to fiber hornification and its impact on enzymatic viscosity control of dissolving pulp,” Cellulose 22(4), 2729-2736. DOI: 10.1007/s10570-015-0636-9

Dufresne, A. (2012). “Swelling and barrier properties,” in: Nanocellulose: From Nature to High Performance Tailored Materials, pp. 373-409. DOI: 10.1515/9783110254600.373

Ebringerová, A., Hromádková, Z., and Heinze, T. (2005). “Hemicellulose,” in: Polysaccharides I: Structure, Characterization and Use, pp. 1-67. DOI: 10.1007/b136816

Ehrnrooth, E., Htun, M., and de Ruvo, A. (1977). “Esterification as a means of improving the properties of once-dried fibers,” Trans. BPBIF Symp. Fiber-Water Interactions in Papermaking, Oxford, 8899-915. DOI: 10.15376/frc.1977.1.899

Eichhorn, S. J., Dufresne, A., Aranguren, M., Marcovich, N. E., Capadona, J. R., Rowan, S. J., Weder, C., Thielemans, W., Roman, M., Renneckar, S., Gindl, W., Veigel, S., Keckes, J., Yano, H., Abe, K., Nogi, M., Nakagaito, A. N. , Mangalam, A., Simonsen, J., Benight, A. S., Bismarck, A., Berglund, L. A., and Peijs, T. (2010). “Review: Current international research into cellulose nanofibres and nanocomposites,” J. Mater. Sci. 45(1), 1-33. DOI: 10.1007/s10853-009-3874-0

Ekeberg, D., Gretland, K. S., Gustafsson, J., Braten, S. M., and Fredheim, G. E. (2006). “Characterisation of lignosulphonates and kraft lignin by hydrophobic interaction chromatography,” Anal. Chimica Acta 565(1), 121-128. DOI: 10.1016/j.aca.2006.02.008

Eklund, D., and Lindström, T. (1991). Paper Chemistry – An introduction,” DT Paper Science Publications, Grankulla, Finland.

Elder, T. (2007). “Quantum chemical determination of Young’s modulus of lignin. Calculations on a beta-O-4’ model compound,” Biomacromol. 8(11), 3619-3627. DOI: 10.1021/bm700663y

El Seoud, O. A., Fidale, L. C., Ruiz, N., D’Almeida, M. L. O., and Frollini, E. (2008). “Cellulose swelling by protic solvents: Which properties of the biopolymer and the solvent matter?” Cellulose 15(3), 371-392. DOI: 10.1007/s10570-007-9189-x

Enomae, T., and Lepoutre, P. (1998). “Observation of the swelling behavior of kraft fibers and sheets in the environmental scanning electron microscope,” Nordic Pulp Paper Res. J. 13(4), 280-284. DOI: 10.3183/npprj-1998-13-04-p280-284

Engquist, I., Lestelius, M., and Liedberg, B. (1997). “Microscopic wettability of ester- and acetate-terminated self-assembled monolayers,” Langmuir 13, 4003-4012. DOI: 10.1021/la9608526

Eriksson, I., Haglind, I., Lidbrandt, O., and Salmén, L. (1991). “Fiber swelling favoured by lignin softening,” Wood Sci. Technol. 25, 135-144. DOI: 10.1007/BF00226813

Etale, A., Onyianta, A. J., Turner, S. R., and Eichhorn, S. (2023). “Cellulose: A review of water interactions, applications in composites, and water treatment,” Chem. Rev. 123, 2016-2048. DOI: 10.1021/acs.chemrev.2c00477

Fält, S., and Wågberg, L. (2003). “Influence of electrolytes on the swelling and strength of kraft-liner pulps,” Nordic Pulp Paper Res. J. 18(1), 69-73. DOI: 10.3183/npprj-2003-18-01-p069-073

Fält, S., Wågberg, L., and Vesterlind, E. L. (2003). “Swelling of model films of cellulose having different charge densities and comparison to the swelling behavior of corresponding fibers,” Langmuir 19(19), 7895-7903. DOI: 10.1021/la026984i

Fardim, P., and Tikka, P. (2011). Chemical Pulping, 2nd Ed., Paper Engineer’s Assoc, Paperi ja Puu Oy.

Fengel, D., and Wegener, G. (1989). Wood – Chemistry, Ultrastructure, Reactions, 2nd Ed., Walter de Gruyter, Berlin.

Fernandes, A. N., Thomas, L. H., Altaner, C. M., Callow, P., Forsyth, V. T., Apperley, D. C., Kennedy, C. J., and Jarvis, M. C. (2011). “Nanostructure of cellulose microfibrils in spruce wood,” Proceedings of the National Academy of Sciences 108(47), E1195-E1203. DOI: 10.1073/pnas.1108942108

Fernandes Diniz, J. M. B., Gil, M. H., and Castro, J. A. A. M. (2004). “Hornification – Its origin and interpretation in wood pulps,” Wood Science and Technology 37, 489-494. DOI: 10.1007/s00226-003-0216-2

Ferreira, J. C., Evtuguin, D. V., and Prates, A. (2020). “Effect of cellulose structure on reactivity of eucalyptus acid sulphite dissolving pulp,” Cellulose 27, 4763-4772. DOI: 10.1007/s10570-020-03092-y

Ferreira, S. R., Silva, F. de A., Lima, P. R. L., and Toledo Filho, R. D. (2017). “Effect of hornification on the structure, tensile behavior and fiber matrix bond of sisal, jute and curauá fiber cement based composite systems,” Construction and Building Materials 139, 551-561. DOI: 10.1016/j.conbuildmat.2016.10.004

Fidale, L. C., Ruiz, N., Heinze, T., and El Seoud, O. A. (2008). “Cellulose swelling by aprotic and protic solvents: What are the similarities and differences?,” Macromol. Chem. Phys. 209(12), 1240-1254. DOI: 10.1002/macp.200800021

Figueiredo, A. B., Magina, S., Evtuguin, D. V., Cardoso, E. F., Ferra, J. M., and Cruz, P. (2016). “Factors affecting the dimensional stability of decorative papers under moistening,” BioResources 11(1), 2020-2029. DOI: 10.15376/biores.11.1.2020-2029

Fisher, L. R., Gamble, R. A., and Middlehurst, J. (1981). “The Kelvin equation and the capillary condensation of water,” Nature 290(5807), 575-576. DOI: 10.1038/290575a0

Fjellström, H., Engstrand, P., and Htun, M. (2012). “Aspects of fibre wall swelling in high-yield pulp,” in: Proceedings of the 4th International Conference on Pulping, Papermaking, and Biotechnology (ICPPB ’12), pp. 1183-1186.

Flory, P. J. (1953). Principles of Polymer Chemistry, Cornell University Press: Ithaca, NY, USA.

Forsström, J., Andreasson, B., and Wågberg, L. (2005). “Influence of pore structure and water retaining ability fibres on the strength of papers from unbleached kraft fibres,” Nordic Pulp Paper Res. J. 20(2), 176-185. DOI: 10.3183/npprj-2005-20-02-p176-185

Frank, H. S., and Wen, W. Y. (1957). “Structural aspects of ion-solvent interaction in aqueous solutions – A suggested picture of water structure,” Disc. Faraday Soc. 24, 133-140. DOI: 10.1039/df9572400133

Fredriksson, M. (2019). “On wood-water interactions in the over-hygroscopic moisture range-mechanisms, methods, and influence of wood modification,” Forests 10(9), article 779. DOI: 10.3390/f10090779

Freire, C. S. R., Silvestre, A. J. D., Neto, C. P., Belgacem, M. N., and Gandini, A. (2006). “Controlled heterogeneous modification of cellulose fibers with fatty acids: Effect of reaction conditions on the extent of esterification and fiber properties,” M. Appl. Polm. Sci. 100(2), 1093-1102. DOI: 10.1002/app.23454

Fries, N., and Dreyer, M. (2008). “An analytic solution of capillary rise restrained by gravity,” Journal of Colloid and Interface Science 320(1), 259-263. DOI: 10.1016/j.jcis.2008.01.009

Froix, M. F., and Nelson, R. (1975). “The interaction of water with cellulose from nuclear magnetic resonance relaxation times,” Macromol. 8(6), 726-730. DOI: 10.1021/ma60048a011

Furman, G., Goren, S., Meerovich, V., Panich, A., Sokolovsky, V., and Xia, Y. (2021). “Anisotropy of transverse and longitudinal relaxations in liquids entrapped in nano- and micro-cavities of a plant stem,” J. Magnetic Resonance 331, article 107051. DOI: 10.1016/j.jmr.2021.107051

Gallay, W. (1950). “The measurement and significance of swelling of cellulose fibers,” TAPPI 33(9), 425-429.

Galvin, K. P. (2005). “A conceptually simple derivation of the Kelvin equation,” Chem. Eng. Sci. 60(16), 4659-4660. DOI: 10.1016/j.ces.2005.03.030

Gamstedt, E. K. (2016). “Moisture induced softening and swelling of natural cellulose fibres in composite applications,” in: 37th Riso International Symposium on Materials Science, book series: IOP Conference Series – Materials Science and Engineering, Vol. 139, article 012003. DOI: 10.1088/1757-899X/139/1/012003

Gamstedt, E. K., Joffre, T., Isaksson, P., Sticko, S., Dumont, P. J. J., du Roscoat, S. R., and Orgeas, L. (2015). “Moisture-induced swelling properties of natural cellulose fibres characterized by synchrotron X-ray computed tomography,” 20th Int. Conf. Composite Materials, O. T. Thomsen, C. Berggreen, and B. F. Sorensen (eds.), Copenhagen, Denmark.

Gane, P. A. C. (2005). “Absorption properties of coatings: A selected overview of absorption criteria derived from recent pore network modelling,” J. Dispersion Sci. Technol. 25(4), 389-408. DOI: 10.1081/DIS-200025737

Ganji, F., Vasheghani-Farahani, S., and Vasheghani-Farahani, E. (2010). “Theoretical description of hydrogel swelling: A review,” Iranian Polymer J. 19(5), 375-398.

Gao, L. X., Shi, S., Hou, W. S., Wang, S. H., Yan, Z. F., and Ge, C. (2020). “NaOH/urea swelling treatment and hydrothermal degradation of waste cotton fiber,” J. Renew. Mater. 8(6), 703-713. DOI: 10.32604/jrm.2020.09055

Gao, X., Zhuang, S. Z., Jin, J. W., and Cao, P. X. (2015). “Bound water content and pore size distribution in swollen cell walls determined by NMR technology,” BioResources 10(4), 8208-8224. DOI: 10.15376/biores.10.4.8208-8224

Geffert, A., Vacek, O., Jankech, A., Geffertova, J., and Milichovsky, M. (2017). “Swelling of cellulosic porous materials – Mathematical description and verification,” BioResources 12(3), 5017-5030. DOI: 10.15376/biores.12.3.5017-5030

Geffertova, J., Geffert, A., and Seman, B. (2013). “Air permeability of sheets and swelling of pulp fibers during recycling and ageing,” Acta Facultatis Xylologiae Zvolen 55(2), 97-104.

Gehmayr, V., Potthast, A., and Sixta, H. (2012). “Reactivity of dissolving pulps modified by TEMPO-mediated oxidation,” Cellulose 19(4), 1125-1134. DOI: 10.1007/s10570-012-9729-x

Gellerstedt, F., Wågberg, L., and Gatenholm, P. (2000). “Swelling behaviour of succinylated fibers,” Cellulose 7(1), 67-86. DOI: 10.1023/A:1009212610018

George, M., Mussone, P. G., and Bressler, D. C. (2015). “Improving the accessibility of hemp fibres using caustic to swell the macrostructure for enzymatic enhancement,” Indust. Crops Prod. 67, 74-80. DOI: 10.1016/j.indcrop.2014.10.043

Gharehkhani, S., Sadeghinezhad, E., Kazi, S. N., Yarmand, H., Badarudin, A., Safaei, M. R., and Zubir, M. N. M. (2015). “Basic effects of pulp refining on fiber properties – A review,” Carbohyd. Polym. 115, 785-803. DOI: 10.1016/j.carbpol.2014.08.047

Ghosh, T., García, A. E., and Garde, S. (2002). “Enthalpy and entropy contributions to the pressure dependence of hydrophobic interactions,” The Journal of Chemical Physics 116(6), 2480-2486. DOI: 10.1063/1.1431582

Glasser, W. G., Atalla, R. H., Blackwell, J., Brown, R., Jr., Burchard, W., French, A. D., Klemm, D. O., and Nishiyama, Y. (2012). “About the structure of cellulose: Debating the Lindman hypothesis,” Cellulose 19, 589-598; and erratum: Cellulose 19: 599. DOI: 10.1007/s10570-012-9691-7

Greca, L. G., Klockars, K. W., Rojas, O. J., and Tardy, B. L. (2023). “Drying stresses to tune strength and long-range order in nanocellulosic materials,” Cellulose 30, 8275-8286. DOI: 10.1007/s10570-023-05353-y

Grignon, J., and Scallan, A. M. (1980). “Effect of pH and neutral salts upon the swelling of cellulose gels,” J. Appl. Polym. Sci. 25(12), 2829-2843. DOI: 10.1002/app.1980.070251215

Groen, J. C., Peffer, L. A. A., and Perez-Ramirez, J. (2003). “Pore size determination in modified micro- and mesoporous materials. Pitfalls and limitations in gas adsorption data analysis,” Micropor. Mesopor. Mater. 60, 1-17. DOI: 10.1016/S1387-1811(03)00339-1

Gruber, E., and Gelbrich, M. (1997). “Effects of enzymatic and chemical drainage aids. Part 1. Freeness and water retention,” Papier 51(4), 166-174.

Grunin, Y. B., Grunin, L. Y., Schiraya, V. Y., Ivanova, M. S., and Masas, D. S. (2020). “Cellulose-water system’s state analysis by proton nuclear magnetic resonance and sorption measurements,” Biores. Bioproc. 7, article 41. DOI: 10.1186/s40643-020-00332-8

Gu, F., Wang, W. X., Cai, Z. S., Xue, F., Jin, Y. C., and Zhu, J. Y. (2018). “Water retention value for characterizing fibrillation degree of cellulosic fibers at micro and nanometer scales,” Cellulose 25(5), 2861-2871. DOI: 10.1007/s10570-018-1765-8

Gupta, P. R., and Goring, D. A. I. (1960). “Physicochemical studies of alkali lignins. 3. Size and shape of the macromolecule,” Can. J. Chem. 38(2), 270-279. DOI: 10.1139/v60-036

Gurnagul, N. (1995). “Sodium-hydroxide addition during recycling – Effects on fiber swelling and sheet strength,” TAPPI J. 78(12), 119-124.

Habibi, Y., Lucia, L. A., and Rojas, O. J. (2010). “Cellulose nanocrystals: Chemistry, self-assembly, and applications,” Chem. Res. 110, 3479-3500. DOI: 10.1021/cr900339w

Han, X.-S., Ye, Y.-H., Lam, F., Pu, J.-W., and Jiang, F. (2019). “Hydrogen-bonding-induced assembly of aligned cellulose nanofibers into ultrastrong and tough bulk materials,” J. Mater. Chem A 7, article 27023. DOI: 10.1039/C9TA11118B

Hanninen, T., Tukiainen, P., Svedstrom, K., Serimaa, R., Saranpaa, P., Kontturi, E., Hughes, M., and Vuorinen, T. (2012). “Ultrastructural evaluation of compression wood-like properties of common juniper (Juniperus communis L.),” Holzforschung 66(3), 389-395. DOI: 10.1515/HF.2011.166

Hansen, C. M. (2007). Hansen Solubility Parameters: A User’s Handbook, 2nd Ed., CRC Press, Boca Raton, FL, USA.

Hantel, M. M., Armstrong, M. J., DaRosa, F., and l’Abee, R. (2017). “Characterization of tortuosity in polyetherimide membranes based on Gurley and electrochemical impedance spectroscopy,” J. Electrochem. Soc. 164(2), A334-A339. DOI: 10.1149/2.1071702jes

Harris, P. J. (2006). “Primary and secondary plant cell walls: A comparative overview,” New Zealand Journal of Forestry Science 36(1), 36.

Hashemzehi, M., and Sjöstrand, B. (2022). “Synthesis of cationized cellulose by deep eutectic solvent for application as a papermaking wet-end additive,” Cellulose Workshop, 14-17.

Hashemzehi, M., and Sjöstrand, B. (2023). “Cationized dialdehyde cellulose synthesized with deep eutectic-like solvents: Effects on sheet dewatering and mechanical properties when used as wet-end additive,” BioResources 18(1), 1347-1367. DOI: 10.15376/biores.18.1.1347-1367

Hashemzehi, M., Sjöstrand, B., Håkansson, H., and Henriksson, G. (2024). “Degrees of hornification in softwood and hardwood kraft pulp during drying from different solvents,” Cellulose Early access. DOI: 10.1007/s10570-023-05657-z

Hassan, M. L., Hassan, E. A., and Oksman, K. N. (2011). “Effect of pretreatment of bagasse fibers on the properties of chitosan/microfibrillated cellulose nanocomposites,” Journal of Materials Science 46(6), 1732-1740. DOI: 10.1007/s10853-010-4992-4

Hayashi, T. (1989). “Xyloglucans in the primary cell wall,” Annual Review of Plant Biology 40(1), 139-168. DOI: 10.1146/annurev.pp.40.060189.001035

Heinonen, E., Henriksson, G., Lindström, M. E., Vilaplana, F. and Wohlert, J. (2022). “Xylan adsorption on cellulose: Preferred alignment and local surface immobilizing effect,” Carbohydrate Polymers 285, article 119221. DOI: 10.1016/j.carbpol.2022.119221

Heinze, T. (2015). “Cellulose: Structure and properties,” in: Rojas, O. (ed.) Cellulose Chemistry and Properties: Fibers, Nanocelluloses and Advanced Materials, Series: Advances in Polymer Science, Vol. 271, Springer, Cham. DOI: 10.1007/12_2015_319

Heinze, T. (2016). “Cellulose: Structure and properties,” in: Cellulose Chemistry and Properties: Fibers, Nanocelluloses and Advanced Materials, pp. 1-52. DOI: 10.1007/12_2015_319

Henriksson, G. (2017), “What are the biological functions of lignin and its complexation with carbohydrates,” Nordic Pulp Paper Res. J. I32(4), 527-541.

Henriksson, G., Akin, D. E., Rigsby, L. L., Patel, N., and Eriksson, K.-E. L. (1997). “Influence of chelating agents and mechanical pretreatment on enzymatic retting of flax,” Textile Research Journal 67(11), 829-836. DOI: 10.1177/004051759706701107

Herrington, T. M., and Petzold, J. C. (1992). “An investigation into the nature of charge on the surface of papermaking woodpulps. I. Charge/pH isotherms,” Colloids Surf. 64(2), 97-108. DOI: 10.1016/0166-6622(92)80088-J

Higgins, H. G., and McKenzie, A. W. (1963). “The structure and properties of paper. XIV. Effects of drying on cellulose fibers and the problem of maintaining pulp strength,” Appita 16(6), 145-164.

Hill, C. A. S., Norton, A., and Newman, G. (2009). “The water vapor sorption behavior of natural fibers,” J. Appl. Polym. Sci. 112(3), 1524-1537. DOI: 10.1002/app.29725

Horn, D., and Melzer, J. (1975). “Influence of macromolecular cationic drainage aids on electrokinetic properties of pulp,” Papier 29(12), 534-541.

Huang, S., Sun, Y. Z., Xu, Y., and Meng, Z. (2014). “Studies on influence of ammonia on properties of cellulose Iβ based on molecular dynamics simulation,” Acta Polymerica Sinica 2014(2), 188-193. DOI: 10.3724/SP.J.1105.2014.13185

Huang, S., Wu, X., and Li, P. X. (2019). “Diffusion behaviors of liquid ammonia in the cellulose based on +molecular dynamics simulation,” Int. J. Clothing Sci. Technol. 31(5), 705-714. DOI: 10.1108/IJCST-12-2018-0163

Hubbe, M. A. (2005). “Microparticle programs for drainage and retention,” in: Rodriguez, J. M. (ed.), Micro and Nanoparticles in Papermaking, TAPPI Press, Atlanta, Chapter 1, 1-36.

Hubbe, M. A. (2019). “Nanocellulose, cationic starch, and paper strength,” APPITA J. 72(2), 82-93.

Hubbe, M. A. (2021). “Contributions of polyelectrolyte complexes and ionic bonding to performance of barrier films for packaging: A review,” BioResources 16(2), 4544-4605. DOI: 10.15376/biores.16.2.Hubbe

Hubbe, M. A., Ayoub, A., Daystar, J. S., Venditti, R. A, and Pawlak, J. J. (2013). “Enhanced absorbent products incorporating cellulose and its derivatives: A review,” BioResources 8(4), 6556-6629. DOI: 10.15376/biores.8.4.6556-6629

Hubbe, M. A., Gardner, D. J., and Shen, W. (2015). “Contact angles and wettability of cellulosic surfaces: A review of proposed mechanisms and test strategies,” BioResources 10(4), 8657-8749. DOI: 10.15376/biores.10.4.Hubbe_Gardner_Shen

Hubbe, M. A., and Heitmann, J. A. (2007). “Review of factors affecting the release of water from cellulosic fibers during paper manufacture,” BioResources 2(3), 500-533. DOI: 10.15376/biores.2.4.500-533

Hubbe, M. A., Heitmann, J. A., and Cole, C. A. (2008). “Water release from fractionated stock suspensions. 2. Effects of consistency, flocculants, shear, and order of mixing,” TAPPI J. 7(7), 28-32. DOI: 10.32964/TJ7.7.28

Hubbe, M. A., and Panczyk, M. (2007a). “Dewatering of refined, bleached hardwood kraft pulp by gravity, vacuum, and centrifugation with applied pressure. Part 1. Physical and ionic effects,” O Papel 68(10), 74-87.

Hubbe, M. A., and Panczyk, M. (2007b). “Dewatering of refined, bleached hardwood kraft pulp by gravity, vacuum, and centrifugation with applied pressure. Part 2. Effects of wet-end additives,” O Papel 68(10), 88-100.

Hubbe, M. A., and Rojas, O. J. (2008). “Colloidal stability and aggregation of lignocellulosic materials in aqueous suspension: A review,” BioResources 3(4), 1419-1491. DOI: 10.15376/biores.3.4.1419-1491

Hubbe, M. A., Rojas, O. J., Lucia, L. A., and Jung, T. M. (2007a). “Consequences of the nanoporosity of cellulosic fibers on their streaming potential and their interactions with cationic polyelectrolytes,” Cellulose 14(6), 655-671. DOI: 10.1007/s10570-006-9098-4

Hubbe, M. A., Sjöstrand, B., Nilsson, L., Kopponen, A., and McDonald, J. D. (2020). “Rate-limiting mechanisms of water removal during the formation, vacuum dewatering, and wet-pressing of paper webs: A review,” BioResources 15(4), 9672-9755. DOI: 10.15376/biores.15.4.Hubbe

Hubbe, M. A., Sundberg, A., Mocchiutti, P., Ni, Y., and Pelton, R. (2012). “Dissolved and colloidal substances (DCS) and the charge demand of papermaking process waters and suspensions: A review,” BioResources 7(4), 6109-6193. DOI: 10.15376/biores.7.4.6109-6193

Hubbe, M. A., Tayeb, P., Joyce, M., Tyagi, P., Kehoe, M., Dimic-Misic, K., and Pal, L. (2017). “Rheology of nanocellulose-rich aqueous suspensions: A review,” BioResources 12(4), 9556-9661. DOI: 10.15376/biores.12.1.2143-2233

Hubbe, M. A., Trovagunta, R., Zambrano, F., Tiller, P., and Jardim, J. (2023). “Self-assembly fundamentals in the reconstruction of lignocellulosic materials: A review,” BioResources 18(2), 4262-4331. DOI: 10.15376/biores.18.2.Hubbe

Hubbe, M. A., Venditti, R. A., and Rojas, O. J. (2007b). “What happens to cellulosic fibers during papermaking and recycling? A review,” BioResources 2(4), 739-788. DOI: 10.15376/biores.2.4.739-788

Hult, E.-L., Iversen, T., and Sugiyama, J. (2003). “Characterization of the supermolecular structure of cellulose in wood pulp fibers,” Cellulose 10, 103-110. DOI: 10.1023/A:1024080700873

Hult, E.-L., Larsson, P. T., and Iversen, T. (2001). “Cellulose fibril aggregation – An inherent property of kraft pulps,” Polymer 42(8), 3309-3314. DOI: 10.1016/S0032-3861(00)00774-6

Ibbett, R. N., and Hsieh, Y. L. (2001). “Effect of fiber swelling on the structure of lyocell fabrics,” Textile Res. J. 71(2), 164-173. DOI: 10.1177/004051750107100212

Inglesby, M. K. and Zeronian, S. H. (1996). “The accessibility of cellulose as determined by dye adsorption,” Cellulose 3(1), 165-181. DOI: 10.1007/BF02228799

Ioelovich, M. Y. (2016). “Models of supramolecular structure and properties of cellulose,” Polymer Science Series A 58(6), 925-943. DOI: 10.1134/S0965545X16060109

Ioelovich, M. (2021). “Preparation, characterization and application of amorphized cellulose—A review,” Polymers 13(24), article 4313. DOI: 10.1134/S0965545X16060109

Ishida, T. (2020). “Theoretical investigation of dissolution and decomposition mechanisms of a cellulose fiber in ionic liquids,” J. Phys. Chem. B 124(15), 3090-3102. DOI: 10.1021/acs.jpcb.9b11527

Isogai, A. (1996). “NMR analysis of cellulose dissolved in aqueous NaOH solutions,” Cellulose 4(2), 99-107. DOI: 10.1023/A:1018471419692

Isogai, A., and Atalla, R. H. (1998). “Dissolution of cellulose in aqueous NaOH solutions,” Cellulose 5(4), 309-319. DOI: 10.1023/A:1009272632367

Iwamoto, S., Kai, W. H., Isogai, A., and Iwata, T. (2009). “Elastic modulus of single cellulose microfibrils from tunicate measured by atomic force microscopy,” Biomacromol. 10(9), 2571-2576. DOI: 10.1021/bm900520n

Iyer, P. B., Sreenivasan, S., Patel, G. S., Chidambareswaran, P. K., Patil, N. B. (1989). “Studies on swelling of cotton fibers in alkali-metal hydroxides. 2. Influence of morphology and fine-structure on tensile behavior,” J. Appl. Polym. Sci. 37(7), 1739-1752. DOI: 10.1002/app.1989.070370701

Jablonsky, M., Botkova, M., Suty, S., Smatko, L., and Sima, J. (2014). “Accelerated ageing of newsprint paper: Changes in swelling ability, WRV and electrokinetic properties of fibres,” Fib. Textiles E. Europe 22(2), 108-113.

Jarvis, M. C. (2018). “Structure of native cellulose microfibrils, the starting point for nanocellulose manufacture,” Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences 376(2112). DOI: 10.1098/rsta.2017.0045

Jarvis, M. C. (2023). “Hydrogen bonding and other non-covalent interactions at the surfaces of cellulose microfibrils,” Cellulose 30, 667-687. DOI: 10.1007/s10570-022-04954-3

Jaturapiree, A., Ehrhardt, A., Groner, S., Oeztuerk, H. B., Siroka, B., and Bechtold, T. (2008). “Treatment in swelling solutions modifying cellulose fiber reactivity – Part 1: Accessibility and sorption,” Macromol. Symp. 262, 39-49. DOI: 10.1002/masy.200850205

Jayme, G., and Büttel, H. (1964). “Die Abhängigkeit des Wasserrückhaltevermögens (WRV-Wert) verschiedener gebleichter und ungebleichter Zellstoffe,” Wochenbl. Papierfabr. 92(23-24), 718-727.

Jayme, G., and Büttel, H. (1968). “Über die Bestimmung und Bedeutung des Wasserrückhaltevermögens (des WRV-Wert) verschiedenen gebleichten und ungebleichten Zellstoffen. II. Beziehung zwischen dem WRV-West und anderen Zellstoffeigenschaften,” Wochenbl. Papierfabr. 92(23-24), 718-727.

Jia, L., Wang, Y. P., Wang, B., He, M. J., Zhang, Q. R., and Shi, B. L. (2008). “Relationship between Hansen solubility parameters and Lewis acid-base parameters of polymers,” J. Macromol. Sci. Part B – Phys. 47(2), 378-383. DOI: 10.1080/00222340701849293

Jiang, X.-Y., Bai, Y.-Y., Chen, X.-F., and Liu, W. (2020). “A review on raw materials, commercial production and properties of lyocell fiber,” J. Bioresour. Bioprod. 5(1), 16-25. DOI: 10.1016/j.jobab.2020.03.002

Joffre, T., Isaksson, P., Dumont, P. J. J., Roscoat, S. R., Sticko, S., Orgeas, L., and Gamstedt, E. K. (2016). “A method to measure moisture induced swelling properties of a single wood cell,” Exper. Mechanics 56(5), 723-733. DOI: 10.1007/s11340-015-0119-9

Josefsson, P., Henriksson, G., and Wågberg, L. (2008). “The physical action of cellulases revealed by a quartz crystal microbalance study using ultrathin cellulose films and pure cellulases,” Biomacromol. 9(1), 249-254. DOI: 10.1021/bm700980b

Joubert, J. M., Kirge, G. J. R., and Borgin, K. (1959). “Evidence for a hydrate of cellulose from studies of its surface properties,” Nature 184(4698), 1561-1562. DOI: 10.1038/1841561b0

Kahar, P., Taku, K., and Tanaka, S. (2013). “Multiple effects of swelling by sodium bicarbonate after delignification on enzymatic saccharification of rice straw,” J. Biosci. Bioeng. 116(6), 725-733. DOI: 10.1016/j.jbiosc.2013.05.036

Kang, K. Y., Hwang, K. R., Park, J. Y., Lee, J. P., Kim, J. S., and Lee, J. S. (2018). “Critical point drying: An effective drying method for direct measurement of the surface area of a pretreated cellulosic biomass,” Polymers 10(6), article 676. DOI: 10.3390/polym10060676

Karlsson, J. O., Andersson, N., Berntsson, P., Chihani, T., and Gatenholm, P. (1998). “Swelling behavior of stimuli-responsive cellulose fibers,” Polymer 39(16), 3589-3595. DOI: 10.1016/S0032-3861(97)10353-6

Karlsson, R. M. P., Larsson, P. T., Yu, S., Pendergraph, S. A., Pettersson, T., Hellwig, J., and Wågberg, L. (2018). “Carbohydrate gel beads as model probes for quantifying non-ionic and ionic contributions behind the swelling of delignified plant fibers,” J. Colloid Interface Sci. 519, 119-129. DOI: 10.1016/j.jcis.2018.02.052

Kasahara, K., Donkai, N., Sasaki, H., and Takagishi, T. (2002a). “Novel observation technique for swollen structure of cellulosic fibers in water,” Sen-I Gakkaishi 58(7), 273-276. DOI: 10.2115/fiber.58.273

Kasahara, K., Sasaki, H., Donkai, N., Ito, H., and Takagishi, T. (2002b). “Characterization of swollen cellulosic fibers by liquid chromatography,” Sen-I Gakkaishi 58(9), 332-336. DOI: 10.2115/fiber.58.332

Kato, K. L., and Cameron, R. E. (1999). “A review of the relationship between thermally-accelerated ageing of paper and hornification,” Cellulose 6, 23-40. DOI: 10.1023/A:1009292120151

Katz, S., Liebergott, N., and Scallan, A. M. (1981). “A mechanism for the alkali strengthening of mechanical pulps,” TAPPI 64(7), 97-100.

Kent, H. J., and Lyne, M. B. (1989). “Influence of paper morphology on short term wetting and sorption phenomena,” in: Fundamentals of Papermaking, Trans. of the IXth Fund. Res. Symp. Cambridge, C. F. Baker and V. Punton (eds.), pp. 895-920, FRC, Manchester. DOI: 10.15376/frc.1989.2.895.

Kerekes, R. J. (2015). “Perspectives on high and low consistency refining in mechanical pulping,” BioResources 10(4), 8795-8811. DOI: 10.15376/biores.10.4.8795-8811

Khazraji, A. C., and Robert, S. (2013). “Interaction effects between cellulose and water in nanocrystalline and amorphous regions: A novel approach using molecular modeling,” J. Nanomater. 2013, article 409676. DOI: 10.1155/2013/409676

Khoo, K. S., Tan, X. F., Ooi, C. W., Chew, K. W., Leong, W. H., Chai, Y. H., Ho, S. H., and Show, P. L. (2021). “How does ionic liquid play a role in sustainability of biomass processing?,” J. Cleaner Prod. 284, article 124772. DOI: 10.1016/j.jclepro.2020.124772

Kihlman, M., Medronho, B. F., Romano, A. L., Germgård, U., and Lindman, B. (2013). “Cellulose dissolution in an alkali based solvent: Influence of additives and pretreatments,” Journal of the Brazilian Chemical Society 24(2), 295-303. DOI: 10.5935/0103-5053.20130038

Kim, H. J., Jo, B. M., and Lee, S. H. (2006). “Potential for energy saving in refining of cellulase-treated kraft pulp,” J. Indust. Eng. Chem. 12(4), 578-583.

Kitaoka, T., Isogai, A., and Onabe, F. (1999). “Chemical modification of pulp fibers by TEMPO-mediated oxidation,” Nordic Pulp Paper Res. J. 14(4), 279-284. DOI: 10.3183/npprj-1999-14-04-p279-284

Köhnke, T., Lund, K., Brelid, H., and Westman, G. (2010). “Kraft pulp hornification: A closer look at the preventive effect gained by glucuronoxylan adsorption,” Carbohydrate Polymers 81(2), 226-233. DOI: 10.1016/j.carbpol.2010.02.023

Kong, Y., Li, L., and Fu, S. (2022). “Insights from molecular dynamics simulations for interaction between cellulose microfibrils and hemicellulose,” Journal of Materials Chemistry A, 10(27). DOI: 10.1039/D2TA03164G

Kontturi, E., Suchy, M., Penttila, P., Jean, B., Pirkkalainen, K., Torkkeli, M., and Serimaa, R. (2011). “Amorphous characteristics of an ultrathin cellulose film,” Biomacromol. 12(3), 770-777. DOI: 10.1021/bm101382q

Kontturi, E., Tammelin, T., and Österberg, M. (2006). “Cellulose—Model films and the fundamental approach,” Chemical Society Reviews 35(12), 1287-1304. DOI: 10.1039/B601872F

Kroon-Batenburg, L. M. J., Kroon, J., and Northolt, M. G. (1986). “Chain modulus and intramolecular hydrogen-bonding in native and regenerated cellulose fibers,” Polym. Commun. 27(10), 290-292.

Kovalenko, V. I. (2010). “Crystalline cellulose: Structure and hydrogen bonds,” Russian Chemical Reviews 79(3), 231- 241. DOI: 10.1070/RC2010v079n03ABEH004065

Kulasinski, K., Keten, S., Churakov, S. V., Guyer, R., Carmeliet, J., and Derome, D. (2014). “Molecular mechanism of moisture-induced transition in amorphous cellulose,” ACS Macro Letters 3(10), 1037-1040. DOI: 10.1021/mz500528m

Kummala, R., Xu, W. Y., Xu, C. L., and Toivakka, M. (2018). “Stiffness and swelling characteristics of nanocellulose films in cell culture media,” Cellulose 25(9), 4969-4978. DOI: 10.1007/s10570-018-1940-y

Kushner, D. J. (1969). “Self-assembly of biological structures,” Bacterolog. Rev. 33(2), 302-345. DOI: 10.1128/br.33.2.302-345.1969

Kvick, M., Martinez, D. M., Hewitt, D. R., and Balmforth, N. J. (2017). “Imbibition with swelling: Capillary rise in thin deformable porous media,” Phys. Rev. Fluids 2(7), article 074001. DOI: 10.1103/PhysRevFluids.2.074001

Lahtinen, P., Liukkonen, S., Pere, J., Sneck, A., and Kangas, H. (2014). “A comparative study of fibrillated fibers from different mechanical and chemical pulps,” BioResources 9(2), 2115-2127. DOI: 10.15376/biores.9.2.2115-2127

Laine, J., Buchert, J., Viikari, L., and Stenius, P. (1996). “Characterization of unbleached kraft pulps by enzymatic treatment, potentiometric titration, and polyelectrolyte adsorption,” Holzforschung 50(3), 208-214. DOI: 10.1515/hfsg.1996.50.3.208

Laine, J., Lindström, T., Bremberg, C., and Glad-Nordmark, G. (2003). “Studies on topochemical modification of cellulosic fibres – Part 5. Comparison of the effects of surface and bulk chemical modification and beating of pulp on paper properties,” Nordic Pulp Paper Res. J. 18(3), 325-332. DOI: 10.3183/npprj-2003-18-03-p325-332

Laine, J., Lindström, T., Nordmark, G. G., and Risinger, G. (2002). “Studies on topochemical modification of cellulosic fibres – Part 2. The effect of carboxymethyl cellulose attachment on fibre swelling and paper strength,” Nordic Pulp Paper Res. J. 17(1), 550-56. DOI: 10.3183/npprj-2002-17-01-p050-056

Laine, J., and Stenius, P. (1997). “Effect of charge on the fibre and paper properties of bleached industrial kraft pulps,” Paperi Puu- Paper Timber 79(4), 257-266.

Laivins, G. V., and Scallan, A. M. (1993). “The mechanism of hornification of wood pulps,” Transactions of the Xth Fundamental Research Symposium, Oxford, 1993, C.F. Baker, (ed.), FRC, Manchester, pp. 1235-1260. DOI: 10.15376/frc.1993.2.1235

Laivins, G. V., and Scallan, A. M. (1996). “The influence of drying and beating on the swelling of fines,” J. Pulp Paper Sci. 22(5), J178-J184.

Langley, J. G., and Litchfield, E. (1986). “Dewatering aids for paper applications,” Proc. TAPPI 1986 Papermakers Conf., TAPPI Press, Atlanta, pp. 89-92.

Larsson, P. A., and Wågberg, L. (2008). “Influence of fibre-fibre joint properties on the dimensional stability of paper,” Cellulose 15(4), 515-525. DOI: 10.1007/s10570-008-9203-y

Lavoine, N., Desloges, I., Dufresne, A., and Bras, J. (2012). “Microfibrillated cellulose: Its barrier properties and applications in cellulosic materials: A review,” Carbohyd. Polym. 90, 735-764. DOI: 10.1016/j.carbpol.2012.05.026

Lawoko, M., Henriksson, G., and Gellerstedt, G. (2003). “New method for quantitative preparation of lignin-carbohydrate complex from unbleached softwood kraft pulp: Lignin-polysaccharide networks. I,” Holzforschung 57(1), 69-74. DOI: 10.1515/HF.2003.011

Lawoko, M., Henriksson, G., and Gellerstedt, G. (2005). “Structural differences between the lignin− carbohydrate complexes present in wood and in chemical pulps,” Biomacromolecules 6(6), 3467-3473. DOI: 10.1021/bm058014q

Le Moigne, N., Bikard, J., and Navard, P. (2010). “Rotation and contraction of native and regenerated cellulose fibers upon swelling and dissolution: The role of morphological and stress unbalances,” Cellulose 17(3), 507-519. DOI: 10.1007/s10570-009-9395-9

Lenz, J., Schurz, J., and Wrentschur, E. (1993). “Properties and structure of solvent-spun and viscose-type fibers in the swollen state,” Colloid Polym. Sci. 271(5), 460-468. DOI: 10.1007/BF00657390

Letkova, E., Letko, M., and Vrska, M. (2011). “Influence of recycling and temperature on the swelling ability of paper,” Chem. Papers 65(6), 822-828. DOI: 10.2478/s11696-011-0089-z

Li, B., Li, H. M., Zha, Q. Q., Bandekar, R., Alsaggaf, A., and Ni, Y. H. (2011). “Review: Effects of wood quality and refining process on TMP pulp and paper quality,” BioResources 6(3), 3569-3584. DOI: 10.15376/biores.6.3.3569-3584

Li, H., Legere, S., He, Z., Zhang, H., Li, J., Yang, B., Zhang, S., Zhang, L., Zheng, L., and Ni, Y. (2018a). “Methods to increase the reactivity of dissolving pulp in the viscose rayon production process: A review,” Cellulose 25, 3733-3753. DOI: 10.1007/s10570-018-1840-1

Li, T.-Q., Hendriksson, U., and Ödberg, L. (1993). “Determination of pore sizes in wood cellulose fibers by 2H and 1H NMR,” Nordic Pulp Paper Res. J. 8(3), 326-330. DOI: 10.3183/npprj-1993-08-03-p326-330

Li, Y., Wang, J. J., Liu, X. M., and Zhang, S. J. (2018). “Towards a molecular understanding of cellulose dissolution in ionic liquids: Anion/cation effect, synergistic mechanism and physicochemical aspects,” Chem. Sci. 9(17), 4027-4043. DOI: 10.1039/c7sc05392d

Lindh, E. L., Terezi, C., Salmén, L., and Furó, I. (2017). “Water in cellulose: Evidence and identification of immobile and mobile adsorbed phases by 2H MAS NMR,” Phys. Chem. Chem. Phys. 19, 4360-4369. DOI: 10.1039/C6CP08219J

Lindman, B., Karlström, G., and Stigsson, L. (2010). “On the mechanism of dissolution of cellulose,” J. Mol. Liq. 156, 76-81. DOI: 10.1016/j.molliq.2010.04.016

Lindman, B., Medronho, B., Alves, L., Costa, C., Edlund, H., and Norgren, M. (2017), “The relevance of structural features of cellulose and its interactions to dissolution, regeneration, gelation and plasticization phenomena,” Phys. Chem. Chem. Phys. 19(35), 23704-23718. DOI: 10.1039/c7cp02409f

Lindström, T. (1986). “The concept and measurement of fiber swelling,” in: Paper Structure and Properties, J. A. Bristow and P. Kolseth (eds.), Marcel Dekker, New York.

Lindström, T. (1992). “Chemical factors affecting the behavior of fibres during papermaking,” Nordic Pulp Paper Res. J. 7(4), 181-192. DOI: 10.3183/npprj-1992-07-04-p181-192

Lindström, T., and Carlsson, G. (1978). “The effect of chemical environment on fiber swelling,” Proc. EUCEPA Symp., pp. 32-52.

Lindström, T., and Carlsson, G. (1982). “The effect of carboxyl groups and their ionic form during drying on the hornification of cellulose fibers,” Svensk Papperstidn. 85(15), R146-R151.

Lindström, T., and Kolman, M. (1982). “The effect of pH and electrolyte concentration during beaching and sheet forming on paper strength,” Svensk Papperstidning 85(15), R140-R145.

Lindgren, J., Öhman, L.-O., Gunnars, S., and Wågberg, L. (2002). “Charge determinations of cellulose fibers of different origin – Comparison between different methods,” Nordic Pulp Paper Res. J. 17(1), 89-96. DOI: 10.3183/npprj-2002-17-01-p089-096

Lisy, A., Haz, A., Nadanyi, R., Jablonsky, M., and Surina, I. (2022). “About hydrophobicity of lignin: A review of selected chemical methods for lignin valorisation in biopolymer production,” Energies 15(17), article 6213. DOI: 10.3390/en15176213

Lloyd, J. A., and Horne, C. W. (1993). “The determination of fiber charge and acidic groups of radiate pine pulps,” Nordic Pulp Paper Res. J. 8(1), 48-52, 67. DOI: 10.3183/npprj-1993-08-01-p048-052

Lokhande, H. T. (1978). “Swelling behavior of cotton fibers in morpholine and piperidine,” J. Appl. Polym. Sci. 22(5), 1243-1253. DOI: 10.1002/app.1978.070220507

Lucas, R. (1918). “Ueber das Zeitgesetz des kapillaren Aufstiegs von Flüssigkeiten,” Kolloid Zeitschrift 23(1), 15-22. DOI: 10.1007/BF01461107

Luo, X. L., Zhu, J. Y., Gleisner, R., and Zhan, H. Y. (2011). “Effects of wet-pressing-induced fiber hornification on enzymatic saccharification of lignocelluloses,” Cellulose 18(4), 1055-1062. DOI: 10.1007/s10570-011-9541-z

Luo, X., and Zhu, J. Y. (2011). “Effects of drying-induced fiber hornification on enzymatic saccharification of lignocelluloses,” Enzyme and Microbial Technology 48, 92-99. DOI: 10.1016/j.enzmictec.2010.09.014

Lyne, M. B. (2002). “Wetting and the penetration of liquids into paper,” in: Handbook of Physical Testing of Paper, Volume 2, 2nd Edition, J. Borch, M. B. Lyne, R. E. Mark, and C. Habeger (eds.), Marcel Dekker Inc., pp. 303-332.

Lyne, L. M., and Gallay, W. (1950). “The effect of drying and heating on the swelling of cellulose fibers and paper strength,” TAPPI 33(9), 429-435.

Ma, J. Z., Li, X. L., and Bao, Y. (2015). “Advances in cellulose-based superabsorbent hydrogels,” RSC Advan. 5(73), 59745-59757. DOI: 10.1039/c5ra08522e

Maloney, T. C. (2015). “Network swelling of TEMPO-oxidized nanocellulose,” Holzforschung 69(2), 207-213. DOI: 10.1515/hf-2014-0013

Maloney, T. C., Johansson, T., and Paulapuro, H. (1998a). “Removal of water from the cell wall during drying,” PITA Water Removal Conference 1997, pp. 2-8.

Maloney, T. C., Paulapuro, H., and Stenius, P. (1998b). “Hydration and swelling of pulp fibers measured with differential scanning calorimetry,” Nordic Pulp Paper Res. J. 13(1), 31-36. DOI: 10.3183/npprj-1998-13-01-p031-036

Manan, S., Ullah, M. W., Ul-Islam, M., Shi, Z. J., Gauthier, M., and Yang, G. (2022). “Bacterial cellulose: Molecular regulation of biosynthesis, supramolecular assembly, and tailored structural and functional properties,” Prog. Mater. Sci. 129, article 100972. DOI: 10.1016/j.pmatsci.2022.100972

Mao, J., Abushammala, H., Pereira, L. B., and Laborie, M. P. (2016). “Swelling and hydrolysis kinetics of kraft pulp fibers in aqueous 1-butyl-3-methylimidazolium hydrogen sulfate solutions,” Carbohyd. Polym. 153, 284-291. DOI: 10.1016/j.carbpol.2016.07.092

Mark, A., Tryding, J., Amini, J., Edelvik, F., Fredlund, M., Glatt, E., Lai, R., Martinsson, L., Nyman, U., Rentzhog, M., Rief, S., and Wiegmann, A. (2012). “Modeling and simulation of paperboard edge wicking,” Nord. Pulp Pap. Res. J. 27(2), 397-402. DOI: 0.3183/npprj-2012-27-02-p397-402

Marmur, A. (2003). “Kinetics of penetration into uniform porous media: Testing the equivalent-capillary concept,” Langmuir 19(14), 5956-5959. DOI: 10.1021/la034490v

Marsh, P. B., Merola, G. V., and Simpson, M. E. (1953). “Experiments with an alkali swelling-centrifuge test applied to cotton fiber,” Textile Res. J. 23(11), 831-841. DOI: 10.1177/004051755302301111

Masoodi, R., and Pillai, K. M. (2010). “Darcy’s law-based model for wicking in paper-like swelling porous media,” AIChE J. 56(9), 2257-2267. DOI: 10.1002/aic.12163

Masoodi, R., Tan, H., and Pillai, K. M. (2012). “Numerical simulation of liquid absorption in paper-like swelling porous media,” AIChE J. 58(8), 2536-2544. DOI: 10.1002/aic.12759

Mathiasson, A., and Kubat, D. G. (1994). “Lignin as binder in particle boards using high-frequency heating – Properties and modulus calculations,” Holz als Roh- un Werkstoff 52(1), 9-18. DOI: 10.1007/BF02615010

Mazeau, K. (2011). “On the external morphology of native cellulose microfibrils,” Carbohydrate Polymers 84(1), 524-532. DOI: 10.1016/j.carbpol.2010.12.016

Mboowa, D. (2021). “A review of the traditional pulping methods and the recent improvements in the pulping processes,” Biomass Conver. Bioref. 14, 1-12. DOI: 10.1007/s13399-020-01243-6

McNamara, J. T., Morgan, J. L. W., and Zimmer, J. (2015). “A molecular description of cellulose biosynthesis,” Annual Rev. Biochem. 84, 895-921. DOI: 10.1146/annurev-biochem-060614-033930

Medronho, B., Romano, A., Miguel, M. G., Stigsson, L., and Lindman, B. (2012). “Rationalizing cellulose (in)solubility: Reviewing basic physicochemical aspects and role of hydrophobic interactions,” Cellulose 19(3), 581-587. DOI: 10.1007/s10570-011-9644-6

Moon, R. J., Martini, A., John Nairn, J., Simonsen, J., and Youngblood, J. (2011). “Cellulose nanomaterials review: Structure, properties and nanocomposites,” Chemical Society Reviews 40(7), 3941-3994. DOI: 10.1039/c0cs00108b

Moreira, L. R. S., and Filho, E. X. F. (2008). “An overview of mannan structure and mannan-degrading enzyme systems,” Applied Microbiology and Biotechnology 79, 165-178. DOI: 10.1007/s00253-008-1423-4

Moriwaki, S., and Hanasaki, I. (2023). “Swelling-based gelation of wet cellulose nanopaper evaluated by single particle tracking,” Sci. Technol. Adv. Mater. 24(1), article 2153622. DOI: 10.1080/14686996.2022.2153622

Moser, C., Backlund, H., Lindström, M., and Henriksson, G. (2018a). “Xyloglucan for estimating the surface area of cellulose fibers,” Nordic Pulp & Paper Research Journal 33(2), 194-199. DOI: 10.1515/npprj-2018-3035

Moser, C., Henriksson, G., and Lindström, M. (2018b). “Improved dispersibility of once-dried cellulose nanofibers in the presence of glycerol,” Nordic Pulp & Paper Research Journal 33(4), 647-650. DOI: 10.1515/npprj-2018-0054

Moss, P. A., and Pere, J. (2006). “Microscopical study on the effects of partial removal of xylan on the swelling properties of birch kraft pulp fibres,” Nordic Pulp Paper Res. J. 21(1), 8-12. DOI10.3183/npprj-2006-21-01-p008-012

Nakano, T. (2017). “Modeling of the morphological change of cellulose microfibrils caused with aqueous NaOH solution: The longitudinal contraction and laterally swelling during decrystallization,” J. Molec. Modeling 23(4), article 129. DOI: 10.1007/s00894-017-3307-y

Nanko, H., and Ohsawa, J. (1989). “Mechanisms of fiber bond formation,” in: Fundamentals of Papermaking, Trans. 9th Fundamental Research Symp., Cambridge, C. F. Baker and V. Punton (eds), pp. 783-830. DOI: 10.15376/frc.1989.2.783

Nazhad, M. M. (1994). Fundamentals of Strength Loss in Recycled Paper, University of British Columbia.

Newman, R. H. (2004). “Carbon-13 NMR evidence for cocrystallization of cellulose as a mechanism for hornification of bleached kraft pulp,” Cellulose 11, 45-52. DOI: 10.1023/B:CELL.0000014768.28924.0c

Nguyen, D. T., Pham, Q. T. (2020). “A theoretical and experimental study on esterification of citric acid with the primary alcohols and the hydroxyl groups of cellulose chain (n=1-2) in parched condition,” J. Chem. 2020, article 8825456. DOI: 10.1155/2020/8825456

Nilsson, N., Singleton, M., and Parker, I. H. (1998). “The preconditioning of paper,” 52nd APPITA Annual General Conference, 1998 Proc., pp. 549-555.

Nishiyama, Y., Asaadi, S., Ahvenainen, P., and Sixta, H. (2019). “Water-induced crystallization and nano-scale spinodal decomposition of cellulose in NMMO and ionic liquid dope,” Cellulose 26(1), 281-289. DOI: 10.1007/s10570-018-2148-x

Nishiyama, Y., Kim, U.-J., Kim, D.-Y., Katsumata, K. S., May, R. P., and Langan, P. (2003). “Periodic disorder along ramie cellulose microfibrils,” Biomacromolecules 4(4), 1013-1017. DOI: 10.1021/bm025772x

Norgren, M., Costa, C., Alves, L., Eivazi, A., Dahlstrom, C., Svanedal, I., Edlund, H., and Medronho, B. (2023). “Perspectives on the Lindman hypothesis and cellulose interactions,” Molecules 28(10), article 4216. DOI: 10.3390/molecules28104216

Notley, S. M., and Norgren, M. (2010). “Surface energy and wettability of spin-coated thin films of lignin isolated from wood,” Langmuir 26(8), 5484-5490. DOI: 10.1021/la1003337

Ogawa, Y., Nishiyama, Y., and Mazeau, K. (2020). “Drying-induced bending deformation of cellulose nanocrystals studied by molecular dynamics simulations,” Cellulose 27(17), 9779-9786. DOI10.1007/s10570-020-03451-9

Oko, A., Claesson, P. M., Niga, P., and Swerin, A. (2016). “Measurements and dimensional scaling of spontaneous imbibition of inkjet droplets on paper,” Nord. Pulp Pap. Res. J. 31(1), 156-169. DOI: 10.3183/NPPRJ-2016-31-01-p156-169

Oko, A., Martinez, D. M., and Swerin, A. (2014). “Infiltration and dimensional scaling of inkjet droplets on thick isotropic porous materials,” Microfluid. Nanofluid. 17(2), 413-422. DOI: 10.1007/s10404-013-1313-7

Oksanen, T., Buchert, J., and Viikari, L. (1997). “The role of hemicelluloses in the hornification of bleached kraft pulps,” Holzforschung 51(4), 355-360. DOI: 10.1515/hfsg.1997.51.4.355

Okugawa, A., Sakaino, M., Yuguchi, Y., and Yamane, C. (2020). “Relaxation phenomenon and swelling behavior of regenerated cellulose fibers affected by water,” Carbohyd. Polym. 231, article 115663. DOI: 10.1016/j.carbpol.2019.115663

Okugawa, A., Yuguchi, Y., and Yamane, C. (2021). “Relaxation phenomenon and swelling behavior of regenerated cellulose fibers affected by organic solvents,” Carbohyd. Polym. 259, article 117656. DOI: 10.1016/j.carbpol.2021.117656

Olejnik, K., Skalski, B., Stanislawska, A., and Wysocka-Robak, A. (2017). “Swelling properties and generation of cellulose fines originating from bleached kraft pulp refined under different operating conditions,” Cellulose 24(9), 3955-3967. DOI: 10.1007/s10570-017-1404-9

Östblom, M., Ekeroth, J., Konradsson, P., and Liedberg, B. (2006). “Structure and desorption energetics of ultrathin D2O ice overlayers on serine- and serinephosphate-terminated self-assembled monolayers,” J. Phys. Chem. B. 8, 1695-1700. DOI: 10.1021/jp055169j

Ottesen, V., and Syverud, K. (2020). “Swelling of individual cellulose nanofibrils in water, role of crystallinity: An AFM study,” Cellulose 28(1), 19-29. DOI: 10.1007/s10570-020-03517-8

Ozturk, H. B., and Bechtold, T. (2007). “Effect of NaOH treatment on the interfibrillar swelling and dyeing properties of lyocell (Tencel (R)) fibres,” Fibres Textiles E. Europe 15(5-6), 114-117.

Paajanen, A., Zitting, A., Rautkari, L., Ketoja, J. A., and Penttilä, P. A. (2022). “Nanoscale mechanism of moisture-induced swelling in wood microfibril bundles,” Nano Letters 22(13), 5143-5150. DOI: 10.1021/acs.nanolett.2c00822

Pääkkö, M., Ankerfors, M., Kosonen, H., Nykanen, A., Ahola, S., Österberg, M., Ruokolainen, J., Laine, J., Larsson, P. T., Ikkala, O., et al. (2007). “Enzymatic hydrolysis combined with mechanical shearing and high-pressure homogenization for nanoscale cellulose fibrils and strong gels,” Biomacromolecules 8(6), 1934-1941. DOI: 10.1021/bm061215p

Page, D. H. (1969). “A theory for the tensile strength of paper,” TAPPI 52(4), 674-681.

Page, D. (1993). “A quantitative theory of the strength of wet webs,” J. Pulp Paper Sci. 19(4), J175-J176.

Page, D. H., and Tydeman, P. A. (1963). “Transverse swelling and shrinkage of softwood tracheids,” Nature 199(489), 471-472. DOI: 10.1038/199471a0

Palasingh, C., Ström, A., Amer, H., and Nypelö, T. (2021). “Oxidized xylan additive for nanocellulose films – A swelling modifier,” Int. J. Biol. Macromol. 180, 753-759. DOI: 10.1016/j.ijbiomac.2021.03.062

Parham, R., and Hergert, H. (1980). “Fluff pulp: A review of its development and current technology,” Pulp Paper 54(3), 110-115, 121.

Park, S., Venditti, R. A., Jameel, H., and Pawlak, J. J. (2007). “Studies of the heat of vaporization of water associated with cellulose fibers characterized by thermal analysis,” Cellulose 14, 195-204. DOI: 10.1007/s10570-007-9108-1

Pejic, B. M., Kostic, M. M., Skundric, P. D., and Praskalo, J. Z. (2008). “The effects of hemicelluloses and lignin removal on water uptake behavior of hemp fibers,” Bioresour. Technol. 99(15), 7152-7159. DOI: 10.1016/j.biortech.2007.12.073

Peng, P., and She, D. (2014). “Isolation, structural characterization, and potential applications of hemicelluloses from bamboo: A review,” Carbohydrate Polymers 112, 701-720. DOI: 10.1016/j.carbpol.2014.06.068

Pönni, R., Galvis, L., and Vuorinen, T. (2014). “Changes in accessibility of cellulose during kraft pulping of wood in deuterium oxide,” Carbohyd. Polym. 101, 792-797. DOI: 10.1016/j.carbpol.2013.10.001

Pönni, R., Vuorinen, T., and Kontturi, E. (2012). “Proposed nano-scale coalescence of cellulose pulp fibers during technical treatments,” BioResources 7(4), 6077-6108. DOI: 10.15376/biores.7.4.6077-6108

Proctor, H. R. (1914). “The equilibrium of dilute hydrochloric acid and gelatin,” J. Chem. Soc. (London) 105, 313-327. DOI: 10.1039/CT9140500313

Przybysz, P., Kucner, M., Dubowik, M., and Przybysz, K. (2017). “Laboratory refining of bleached softwood kraft pulp in water and a series of alcohols of different molecular weights and polarities: Effects on swelling and fiber length,” BioResources 12(1), 1737-1748. DOI: 10.15376/biores.12.1.1737-1748

Qi, H. S., Yang, Q. L., Zhang, L. N., Liebert, T., and Heinze, T. (2011). “The dissolution of cellulose in NaOH-based aqueous system by two-step process,” Cellulose 18(2), 237-245. DOI: 10.1007/s10570-010-9477-8

Qing, Y., Wu, Y. Q., Cai, Z. Y., and Li, X. J. (2013). “Water-triggered dimensional swelling of cellulose nanofibril films: Instant observation using optical microscope,” J. Nanomater. 2013, article 594734. DOI: 10.1155/2013/594734

Quéré, D. (1997). “Inertial capillarity,” Europhys. Lett. 39(5), 533-538. DOI: 10.1209/epl/i1997-00389-2

Racz, I., and Borsa, J. (1997). “Swelling of carboxymethylated cellulose fibres,” Cellulose 4(4), 293-303. DOI: 10.1023/A:1018400226052

Rahman, M. B. A., Ishak, Z. I., Abdullah, D. K., Aziz, A. A., Basri, M., and Salleh, A. B. (2012). “Swelling and dissolution of oil palm biomass in ionic liquids,” J. Oil Palm Res. 24, 1267-1276.

Ramarao, B. V., and Chatterjee, S. G. (1997). “Moisture sorption by paper materials under varying humidity conditions,” in: Fundamentals of Papermaking Materials, C. F. Baker (ed.), Trans. 11th Fundamental Research Symposium in the Fundamentals of Papermaking Materials, Cambridge, UK, pp. 703-749. DOI: 10.15376/frc.1997.2.703

Rapaport, D. C. (1983). “Hydrogen-bonds in water network organization and lifetimes,” Molec. Phys. 50(5), 1151-1162. DOI: 10.1080/00268978300102931

Rehfeldt, F., and Tanaka, M. (2003). “Hydration forces in ultrathin films of cellulose,” Langmuir 19(5), 1467-1473. DOI: 10.1021/la0261702

Reid, M. S., Kedzior, S. A., Villalobos, M., and Cranston, E. D. (2017). “Effect of ionic strength and surface charge density on the kinetics of cellulose nanocrystal thin film swelling,” Langmuir 33(30), 7403-7411. DOI: 10.1021/acs.langmuir.7b01740

Reid, M. S., Villalobos, M., and Cranston, E. D. (2016). “Cellulose nanocrystal interactions probed by thin film swelling to predict dispersibility,” Nanoscale 8(24), 12247-12257. DOI: 10.1039/c6nr01737a

Remadevi, R., Gordon, S., Wang, X. G., and Rajkhowa, R. (2017). “Investigation of the swelling of cotton fibers using aqueous glycine solutions,” Textile Res. J. 87(18), 2204-2213. DOI: 10.1177/0040517516665267

Reyssat, M., Courbin, L., Reyssat, E., and Stone, H. A. (2008). “Imbibition in geometries with axial variations,” Journal of Fluid Mechanics 615, 335-344. DOI: 10.1017/S0022112008003996

Roberts, G. A. F. (1996). “Accessibility of cellulose,” in: Paper Chemistry, Second Ed., J. C. Roberts (ed.), Blackie Acad. Prof., London. DOI: 10.1007/978-94-011-0605-4_2

Roberts, R. J., Senden, T. J., Knackstedt, M. A., and Lyne, M. B. (2003). “Spreading of aqueous liquids in unsized papers is by film flow,” Journal of Pulp and Paper Science 29(4), 123-131.

Rowland, B., Kadagathur, N. S., Devlin, J. P., Buch, V., Feldman, T., and Wojcik, M. J. (1995). “Infrared spectra of ice surfaces and assignment of surface-localized modes from simulated spectra of cubic ice,” J. Chem. Phys. 102, 8328-8341. DOI: 10.1063/1.468825

Salam, A., Pawlak, J. J., Venditti, R. A., and El-Tahlawy, K. (2011). “Incorporation of carboxyl groups into xylan for improved absorbency,” Cellulose 18(4), 1033-1041. DOI: 10.1007/s10570-011-9542-y

Salem, K. S., Naithani, V., Jameel, H., Lucia, L., and Pal, L. (2022). “A systematic examination of the dynamics of water-cellulose interactions on capillary force-induced fiber collapse,” Carbohyd. Polym. 295, article 119856. DOI: 10.1016/j.carbpol.2022.119856

Salmén, L. (2022). “On the organization of hemicelluloses in the wood cell wall,” Cellulose 29(3), 1349-1355. DOI: 10.1007/s10570-022-04425-9

Salmén, L., and Stevanic, J. S. (2018). “Effect of drying conditions on cellulose microfibril aggregation and ‘hornification’,” Cellulose 25(11), 6333-6344. DOI: 10.1007/s10570-018-2039-1

Salminen, P. (1988). Studies of Water Transport in Paper during Short Contact Times, Ph.D. thesis, Åbo Academy University, Finland.

Saxena, I. M., and Brown, R. M. Jr. (2005). “Cellulose biosynthesis: Current views and evolving concepts,” Annals of Botany 96(1), 9-21. DOI: 10.1093/aob/mci155

Sayyed, A. J., Mohite, L. V., Deshmukh, N. A., and Pinjari, D. V. (2018). “Effect of ultrasound treatment on swelling behavior of cellulose in aqueous N-methyl-morpholine-N-oxide solution,” Ultrasonics Sonochem. 49, 161-168. DOI: 10.1016/j.ultsonch.2018.07.042

Sayyed, A. J., Mohite, L. V., Deshmukh, N. A., and Pinjari, D. V. (2021). “Swelling kinetic study with mathematical modeling of cellulose pulp in aqueous N-methyl-morpholine-N-oxide solution,” Reaction Kinetics Mechan. Catal. 133(1), 101-115. DOI: 10.1007/s11144-021-02000-0

Scallan, A. M. (1983). “The effect of acid groups on the swelling of pulps: A review,” TAPPI J. 66(11), 73-75.

Scallan, A. M., and Carles, J. E. (1972). “The correlation of the water retention value with the fibre saturation point,” Svensk Papperstidn. 75, 699-703.

Scallan, A. M., and Grignon, J. (1979). “The effect of cations on pulp and paper properties,” Svensk Papperstidning 82(2), 40-47.

Scallan, A. M., and Tigerström, A. (1992). “Swelling and elasticity of the cell walls of pulp fibers,” J. Pulp Paper Sci. 18(5), J188-J193.

SCAN (2000). “Water retention value,” SCAN-C 62:00, Scandinavian Pulp, Paper and Board Testing Committee.

Scheidegger, A. H. (1974). “The physics of flow through porous media,” 3rd Ed., University of Toronto Press, ISBN 0-8020-1849-1

Scheller, H. V., and Ulvskov, P. (2010). “Hemicelluloses,” Annual Review of Plant Biology 61, 263-289. DOI: 10.1146/annurev-arplant-042809-112315

Schenzel, K., Almlöf, H., and Germgård, U. (2009). “Quantitative analysis of the transformation process of cellulose I→ cellulose II using NIR FT Raman spectroscopy and chemometric methods,” Cellulose 16, 407-415. DOI: 10.1007/s10570-009-9286-0

Seeberger, P. H. (2015). “Monosaccharide diversity,” in: Essentials of Glycobiology, 3rd Ed., Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY, USA. PMID: 28876863.

Sellman, F. A., Benselfelt, R., Larsson, P. T., and Wågberg, L. (2023). “Hornification of cellulose-rich materials – A kinetically trapped state,” Carbohyd. Polym. 318, article 121132. DOI: 10.1016/j.carbpol.2023.121132

Senden, T. J., Knackstedt, M. A., and Lyne, M. B. (2000). “Droplet penetration into porous networks: Role of pore morphology,” Nord. Pulp Pap. Res. J. 15(5), 554-563, DOI: 10.3183/NPPRJ-2000-15-05-p554-563

Shafiei-Sabet, S., Martinez, M., and Olson, J. (2016). “Shear rheology of micro-fibrillar cellulose aqueous suspensions,” Cellulose 23(5), 2943-2953. DOI: 10.1007/s10570-016-1040-9

Shen, T.-Y., and Gnanakaran, S. (2009). “The stability of cellulose: A statistical perspective from a coarse-grained model of hydrogen bond networks,” Biophys. J. 96, 3032-3040. DOI: 10.1016/j.bpj.2008.12.3953

Shrestha, S., Diaz, J. A., Ghanbari, S., and Youngblood, J. P. (2017). “Hygroscopic swelling determination of cellulose nanocrystal (CNC) films by polarized light microscopy digital image correlation,” Biomacromol. 18(5), 1482-1490. DOI: 10.1021/acs.biomac.7b00026

Sim, G., Alam, M. N., Godbout, L., and van de Ven, T. (2014). “Structure of swollen carboxylated cellulose fibers,” Cellulose 21(6), 4595-4606. DOI: 10.1007/s10570-014-0425-x

Simões, R. M. S., Ferreira, C. I. A., Pires, F., Martins, M., Ramos, A. N. A., Pinto, P. C. R., and Jorge, R. (2023). “Recycling performance of softwood and hardwood unbleached kraft pulps for packaging papers,” TAPPI Journal 22(2), 73-86. DOI: 10.32964/TJ22.2.73

Sing, K. (2001). “The use of nitrogen adsorption for the characterisation of porous materials,” Colloids Surf. A – Physicochem. Eng. Aspects 187, 3-9. DOI: 10.1016/S0927-7757(01)00612-4

Sinko, R., Qin, X., and Keten, S. (2015). “Interfacial mechanics of cellulose nanocrystals,” MRS Bull. 41, 340-348. DOI: 10.1557/mrs.2015.67

Sjöstedt, A., Wohlert, J., Larsson, P. T., and Wågberg, L. (2015). “Structural changes during swelling of highly charged cellulose fibres,” Cellulose 22(5), 2943-2953. DOI: 10.1007/s10570-015-0701-4

Sjöstrand, B., Karlsson, C.-A., Barbier, C., and Henriksson, G. (2023). “Hornification in commercial chemical pulps, dependence on water removal and bonding mechanisms,” BioResources 18(2), 3856-3869. DOI: 10.15376/biores.18.2.3856-3869

Solhi, L., Guccini, V., Heise, K., Solala, I., Niinivaara, E., Xu, W., Mihhels, K., Kr, M., Meng, Z., Wohlert, J., Tao, H., Cranston, E. D., and Kontturi, E. (2023). “Understanding nanocellulose − water interactions: Turning a detriment into an asset,” Chemical Reviews 123(5), 1925-2015. DOI: 10.1021/acs.chemrev.2c00611

Spinu, M., Dos Santos, N., Le Moigne, N., and Navard, P. (2011). “How does the never-dried state influence the swelling and dissolution of cellulose fibres in aqueous solvent?,” Cellulose 18(2), 247-256. DOI: 10.1007/s10570-010-9485-8

Sreenivasan, S., Iyer, P. B., and Patel, G. S. (1993). “Studies on the swelling of cotton fibers in alkali-metal hydroxides. 3. Structure property relations in fibers swollen at 0-degrees-C,” J. Appl. Polym. Sci. 48(3), 393-404. DOI: 10.1002/app.1993.070480303

Sreenivasan, S., Iyer, P. B., Patel, G. S., and Chidambareswaran, P. K. (1989). “Studies on swelling of cotton fibers in alkali-metal hydroxides. 1. Influence of variations in fine-structure on tensile behavior,” J. Appl. Polym. Sci. 37(8), 2191-2201. DOI: 10.1002/app.1989.070370811

Sreenivasan, S., Iyer, P. B., Patel, G. S., and Iyer, K. R. K. (1995). “Studies on swelling of cotton fibers in alkali-metal hydroxides. 4. Influence of initial fiber properties and variations in fine-structure on tensile behavior,” J. Appl. Polymer Sci. 58(13), 2405-2413. DOI: 10.1002/app.1995.070581307

Stone, J., and Scallan, A. (1966). “Influence of drying on the pore structures of the cell wall,” in: Consolidation of the Paper Web. Transactions of the IIIrd Symposium, Cambridge, 1965, F. Bolam (ed.), Tech. Sec. British Paper & Board Makers’ Assoc., Longon, Vol. 1, 145-166. DOI: 10.15376/frc.1965.1.145

Stone, J. E., and Scallan, A. M. (1967). “Effect of component removal upon porous structure of cell wall of wood. 2. Swelling in water and fiber saturation point,” TAPPI 50(10), 496-501.

Stone, J., and Scallan, A. (1968). “A structural model for the cell wall of water-swollen wood pulp fibers based on their accessibility to macromolecules,” Cellulose Chem. Technol. 2, 343-358.

Ström, G., and Kunnas, A. (1991). “The effect of cationic polymers on the water retention value of various pulp,” Nordic Pulp Paper Res. J. 6(1), 12-19. DOI: 10.3183/npprj-1991-06-01-p012-019

Suchy, M., Virtanen, J., Kontturi, E., and Vuorinen, T. (2010). “Impact of drying on wood ultrastructure observed by deuterium exchange and photoacoustic FT-IR spectroscopy,” Biomacromol. 11, 515-520. DOI: 10.1021/bm901268j

Sun, B. Z., Peng, G. G., Duan, L., Xu, A. H., and Li, X. X. (2015). “Pretreatment by NaOH swelling and then HCl regeneration to enhance the acid hydrolysis of cellulose to glucose,” Bioresour. Technol. 196, 454-458. DOI: 10.1016/j.biortech.2015.08.009

Swerin, A., Ödberg, L., and Lindström, T. (1990). “Deswelling of hardwood kraft pulp fibers by cationic polymers,” Nordic Pulp Paper Res. J. 5(4), 188-196. DOI: 10.3183/npprj-1990-05-04-p188-196

Tait, M. J., Ablett, S., and Wood, F. W. (1972). “Binding of water on starch, an NMR investigation,” J. Colloid Interface Sci. 41(3), 594-603. DOI: 10.1016/0021-9797(72)90381-5

TAPPI (1981). “Water retention value (WRV),” Useful Test Method, UM 256, TAPPI Press, Atlanta.

Tatarova, I., Manian, A. P., Siroka, B., and Bechtold, T. (2010). “Nonalkali swelling solutions for regenerated cellulose,” Cellulose 17(5), 913-922. DOI: 10.1007/s10570-010-9429-3

Thakur, B. R., Singh, R. K., Handa, A. K., and Rao, M. A. (1997). “Chemistry and uses of pectin—A review,” Critical Reviews in Food Science & Nutrition 37(1), 47-73. DOI: 10.1080/10408399709527767

Thomson, W. (Lord Kelvin) (1871). “On the equilibrium of vapour at a curved surface of liquid,” Philosophical Mag., Ser. 4, 42(282), 448-452. DOI: 10.1080/14786447108640606

Thybring, E. E., Boardman, C. R., Zelinka, S. L., and Glass, S. V. (2021). “Common sorption isotherm models are not physically valid for water in wood,” Colloids Surf. A: Physicochem. Eng. Aspects 627, article 127214. DOI: 10.1016/j.colsurfa.2021.127214

Thybring, E. E., Fredriksson, M., Zelinka, S. L., and Glass, S. V. (2022). “Water in wood: A review of current understanding and knowledge gaps,” Forests 13, article 2051. DOI: 10.3390/f13122051

Tian, S. B., Jiang, J. A., Zhu, P. G., Yu, Z. Y., Oguzlu, H., Balldelli, A., Wu, J., Zhu, J. Y., Sun, X., and Saddler, J. (2022). “Fabrication of a transparent and biodegradable cellulose film from kraft pulp via cold alkaline swelling and mechanical blending,” ACS Sustain. Chem. Eng. 10(32), 10560-10569. DOI: 10.1021/acssuschemeng.2c01937

Tobias, L. M., Spokevicius, A. V., McFarlane, H. E., and Bossinger, G. (2020). “The cytoskeleton and its role in determining cellulose microfibril angle in secondary cell walls of woody tree species,” Plasts – Basel 9(1), article 90. DOI: 10.3390/plants9010090

Torstensen, J. O., Liu, M., Jin, S. A., Deng, L. Y., Hawari, A. I., Syverud, K., Spontak, R. J., and Gregersen, O. W. (2018). “Swelling and free-volume characteristics of TEMPO-oxidized cellulose nanofibril films,” Biomacromol. 19(3), 1016-1025. DOI: 10.1021/acs.biomac.7b01814

Torstensen, J., Ottesen, V., Rodriguez-Fabia, S., Syverud, K., Johansson, L., and Lervik, A. (2022). “The influence of temperature on cellulose swelling at constant water density,” Sci. Rep. 12(1), article 20736. DOI: 10.1038/s41598-022-22092-5

Toven, K. (2003). “Paper properties and swelling properties of ozone-based ECF bleached softwood kraft pulps,” TAPPI J. 2(2), 3-7.

Tydeman, P. A., Wembridge, D. R., and Page, D. H. (1966). “Transverse shrinkage of individual fibres by micro-radiography,” in: Consolidation of the Paper Web, Trans. of the IIIrd Fund. Res. Symp., Cambridge, 1965, F. Bolam (ed.), Techn. Sect. BPBMF, pp.119-144. DOI: 10.15376/frc.1965.1.119

Tze, W., and Gardner, D. (2001). “Swelling of recycled wood pulp fibers: Effect on hydroxyl availability and surface chemistry,” Wood and Fiber Science 33(3), 364-376.

Valencia, J., and Pierola, I. F. (2007). “Interpretation of the polyelectrolyte and antipolyelectrolyte effects of poly(N-vinylimidazoleco-sodium styrenesulfonate) hydrogels,” J. Polym. Sci. Part B – Polym. Phys. 45(13), 1683-1693. DOI: 10.1002/polb.21194

van Oss, C. J., Wu, W., Docoslis, A., and Giese, R. F. (2001). “The interfacial tensions with water and the Lewis acid-base surface tension parameters of polar organic liquids derived from their aqueous solubilities,” Colloids Surf. B – Biointerfaces 20(1), 87-91. DOI: 10.1016/S0927-7765(00)00169-7

Vänskä, E., Vihelä, T., Peresin, M. S., Vartiainen, J., Hummel, M., and Vuorinen, T. (2016). “Residual lignin inhibits thermal degradation of cellulosic fiber sheets,” Cellulose 23(1), 199-212. DOI: 10.1007/s10570-015-0791-z

Vegunta, V., Senthilkumar, E. R., Lindén, P., Sevastyanova, O., Vilaplana, F., Garcia, A., Björk, M., Jansson, U., Henriksson, G., and Lindström, M. E. (2022). “High calcium content of Eucalyptus dunnii wood affects delignification and polysaccharide degradation in kraft pulping,” Nordic Pulp & Paper Research Journal 37(2), 338-348. DOI: 10.1515/npprj-2021-0069

Vermaas, D., and Hermans, P. H. (1947). “Course of acetylation and deacetylation reactions of cellulose fibers. 1. Optical and swelling properties,” J. Polym. Sci. 2(4), 397-405. DOI: 10.1002/pol.1947.120020405

Verwey, E. J. W., and Overbeek, J. Th. G. (1948). Theory of the Stability of Lyophobic Colloids, Elsevier, New York.

Vidal, P. F., Basora, N., Overend, R. P., and Chornet, E. (1991). “A pseudouniversal calibration procedure for the molecular-weight determination of cellulose,” J. Appl. Polym. Sci. 42(6), 1659-1664. DOI: 10.1002/app.1991.070420620

von Schreeb, A., Sjöstrand, B., Ek, M., and Henriksson, G. (202X). “Drying and hornification of very swollen cellulose,” SUBMITTED

Vu-Manh, H., Ozturk, H. B., and Bechtold, T. (2010a). “Swelling and dissolution mechanism of regenerated cellulosic fibers in aqueous alkaline solution containing ferric tartaric acid complex: Part I. Viscose fibers,” Carbohyd. Polym. 82(3), 761-767. DOI: 10.1016/j.carbpol.2010.05.048

Vu-Manh, H., Ozturk, H. B., and Bechtold, T. (2010b). “Swelling and dissolution mechanism of regenerated cellulosic fibers in aqueous alkaline solution containing ferric-tartaric acid complex – Part II: Modal fibers,” Carbohyd. Polym. 82(4), 1068-1073. DOI: 10.1016/j.carbpol.2010.06.043

Wang, J. L., Chen, W., Dong, T. T., Wang, H. Q., Si, S. R., and Li, X. S. (2021). “Enabled cellulose nanopaper with outstanding water stability and wet strength via activated residual lignin as a reinforcement,” Green Chem. 23(24), 10062-10070. DOI: 10.1039/d1gc03906g

Wang, J. L., Wang, Q., Wu, Y. T., Bai, F. T., Wang, H. Q., Si, S. R., Lu, Y. F., Li, X. S., and Wang, S. F., (2020). “Preparation of cellulose nanofibers from bagasse by phosphoric acid and hydrogen peroxide enables fibrillation via a swelling, hydrolysis, and oxidation cooperative mechanism,” Nanomater. 10(11), article 2227. DOI: 10.3390/nano10112227

Wang, J. P., Quirk, A., Lipkowski, J., Dutcher, J. R., Hill, C., Mark, A., and Clarke, A. J. (2012). “Real-time observation of the swelling and hydrolysis of a single crystalline cellulose fiber catalyzed by cellulase 7B from Trichoderma reesei,” Langmuir 28(25), 9664-9672. DOI: 10.1021/la301030f

Wang, Y., Lindström, M. E., and Henriksson, G. (2014a). “Increased degradability of cellulose by dissolution in cold alkali,” BioResources 9(4), 7566-7578. DOI: 10.15376/biores.9.4.7566-7578

Wang, Y., Yao, S., Jia, C. M., Chen, P. F., and Song, H. (2014b). “Swelling behaviors of natural cellulose in ionic liquid aqueous solutions,” J. Appl. Polym. Sci. 131(9). DOI: 10.1002/app.40199

Warwicker, J. O., and Wright, A. C. (1967). “Function of sheets of cellulose chains in swelling reactions on cellulose,” J. Appl. Polym. Sci. 11(5), 659-671. DOI: 10.1002/app.1967.070110504

Washburn, E. W. (1921). “The dynamics of capillary flow,” Physics Review 17(3), 273-283. DOI: 10.1103/PhysRev.17.273

Weise, U., Maloney, T., and Paulapuro, H. (1996). “Quantification of water in different states of interaction with wood pulp fibres,” Cellulose 3(4), 189-202. DOI: 10.1007/BF02228801

Weise, U. and Paulapuro, H. (1999). “Effect of drying and rewetting cycles on fiber swelling,” J. Pulp Paper Sci. 25(5), 163-166.

Welf, E. S., Venditti, R. A., Hubbe, M. A., and Pawlak, J. (2005). “The effects of heating without water removal and drying on the swelling as measured by water retention value and degradation as measured by intrinsic viscosity of cellulose papermaking fibers,” Prog. Paper Recycling 14(3), 1-9.

Welo, L. A., Ziifle, H. M., and McDonald, A. W. (1952). “Swelling capacities of fibers in water. 2. Centrifuge studies,” Textile Res. J. 22(4), 261-273. DOI: 10.1177/004051755202200404

Wenzel, R. N. (1949). “Surface roughness and contact angle,” J. Phys. Colloid Chem. 53(9), 1466-1467. DOI: 10.1021/j150474a015

Wohlert, M., Benselfelt, T., Wågberg, L., Furó, I., Berglund, L. A., and Wohlert, J. (2022). “Cellulose and the role of hydrogen bonds: Not in charge of everything,” Cellulose 29, 1-23. DOI: 10.1007/s10570-021-04325-4

Wong, C. L., Wang, S., Karimnejad, S., Wijburg, M. G., Mansouri, H., and Darhuber, A. A. (2023). “Transient deformation and swelling of paper by aqueous co-solvent solutions,” Soft Matter 19(6), 1202-1211. DOI: 10.1039/d2sm01388f

Wu, N., Hubbe, M. A., Rojas, O. J., and Park, S. (2009). “Permeation of polyelectrolytes and other solutes into the pore spaces of water-swollen cellulose: A review,” BioResources 4(3), 1222-1262. DOI: 10.15376/biores.4.3.1222-1262

Xiao, J., Stone, H. A., and Attinger, D. (2012). “Source-like solution for radial imbibition into a homogeneous semi-infinite porous medium,” Langmuir 28(9), 4208-4212. DOI: 10.1021/la204474f

Xu, J. L., Zhang, B. CT., Lu, X. M., Zhou, Y. H., Fang, J. Y., Li, Y., and Zhang, S. J. (2018). “Nanoscale observation of microfibril swelling and dissolution in ionic liquids,” ACS Sustain. Chem. Eng. 6(1), 909-917. DOI: 10.1021/acssuschemeng.7b03269

Yamane, C., Aoyagi, T., Ago, M., Sato, K., Okajima, K., and Takahashi, T. (2006). “Two different surface properties of regenerated cellulose due to structural anisotropy,” Polym. J. 38(8), 819-826. DOI: 10.1295/polymj.PJ2005187

Yan, L. F., Chen, J., and Bangal, P. R. (2007). “Dissolving cellulose in a NaOH/thiourea aqueous solution: A topochemical investigation,” Macromol. Biosci. 7(9–10), 1139-1148. DOI: 10.1002/mabi.200700072

Yao, Y., and Lenhoff, A. M. (2004). “Determination of pore size distributions of porous chromatographic adsorbents by inverse size-exclusion chromatography,” J. Chromatog. A 1037(1-2), 273-282. DOI: 10.1016/j.chroma.2004.02.054

You, X., Chen, F., Ma, Y. B., Roselli, A., Enqvist, E., and Hassi, H. (2021). “Single fiber swelling behavior for natural and man-made cellulose fibers under alkaline treatment,” Cellulose 28(18), 11287-11298. DOI: 10.1007/s10570-021-04280-0

Young, R. A. (1994). “Comparison of the properties of chemical cellulose pulps,” Cellulose 1(2), 107-130. DOI: 10.1007/BF00819662

Yui, T., Nishimura, S., Akiba, S., and Hayashi, S. (2006). “Swelling behavior of the cellulose I beta crystal models by molecular dynamics,” Carbohyd. Res. 341(15), 2521-2530. DOI: 10.1016/j.carres.2006.04.051

Zanuttini, M., and Marzocchi, V. (2003). “Alkaline chemi-mechanical pulp from poplar. Relationship between chemical state, swelling and papermaking properties,” Holzforschung 57(5), 489-495. DOI: 10.1515/HF.2003.073

Zauscher, S., and Klingenberg, D. J. (2000). “Normal forces between cellulose surfaces measured with colloidal probe microscopy,” J. Colloid Interface Sci. 229(2), 497-510. DOI: 10.1006/jcis.2000.7008

Zhang, J. H., Song, H. N., Lin, L., Zhuang, J. P., Pang, C. S., and Liu, S. J. (2012). “Microfibrillated cellulose from bamboo pulp and its properties,” Biomass Bioenergy 39, 78-83. DOI: 10.1016/j.biombioe.2010.06.013

Zhang, M., Hubbe, M. A., Venditti, R. A., and Heitmann, J. A. (2002). “Can recycled kraft fibers benefit from chemical addition before they are first dried?,” APPITA J. 55(2), 135-144.

Zhang, S., Wang, W. C., Li, F. X., and Yu, J. Y. (2013). “Swelling and dissolution of cellulose in NaOH aqueous solvent systems,” Cellulose Chem. Technol. 47(9-10), 671-679.

Zhang, W. S., Okubayashi, S., and Bechtold, T. (2005). “Fibrillation tendency of cellulosic fibers. Part 1: Effects of swelling,” Cellulose 12(3), 267-273. DOI: 10.1007/s10570-004-2786-z

Zhang, X. M., Qu, T. J., Mosier, N. S., Han, L. J., and Xiao, W. H. (2018). “Cellulose modification by recyclable swelling solvents,” Biotech. Biofuels 11, article 191. DOI: 10.1186/s13068-018-1191-z

Zhang, Y. H. P., Cui, J. B., Lynd, L. R., and Kuang, L. R. (2006). “A transition from cellulose swelling to cellulose dissolution by o-phosphoric acid: Evidence from enzymatic hydrolysis and supramolecular structure,” Biomacromol. 7(2), 644-648. DOI: 10.1021/bm050799c

Zhmud, B. V., Tiberg, F., and Hallstensson, K. (2000). “Dynamics of capillary rise,” Journal of Colloid and Interface Science 228(2), 263-269. DOI: 10.1006/jcis.2000.6951

Zimnitsky, D. S., Yurkshtovich, T. L., and Bychkovsky, P. M. (2004). “Synthesis and characterization of oxidized cellulose,” J. Polym. Sci. Part A – Polym. Chem. 42(19), 4785-4791. DOI: 10.1002/pola.20302