NC State
BioResources
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," BioRes. 8(4), 6556-6629.

Abstract

Cellulose and some cellulose derivatives can play vital roles in the enhancement of the performance of absorbent products. Cellulose itself, in the form of cellulosic fibers or nano-fibers, can provide structure, bulk, water-holding capacity, and channeling of fluids over a wide dimensional range. Likewise, cellulose derivatives such as carboxymethylcellulose (CMC) have been widely studied as components in superabsorbent polymer (SAP) formulations. The present review focuses on strategies and mechanisms in which inclusion of cellulose – in its various forms – can enhance either the capacity or the rate of aqueous fluid absorption in various potential applications.


Download PDF

Full Article

Enhanced Absorbent Products Incorporating Cellulose and Its Derivatives: A Review

Martin A. Hubbe, Ali Ayoub, Jesse S. Daystar, Richard A. Venditti, and Joel J. Pawlak

Cellulose and some cellulose derivatives can play vital roles in the enhancement of the performance of absorbent products. Cellulose itself, in the form of cellulosic fibers or nano-fibers, can provide structure, bulk, water-holding capacity, and channeling of fluids over a wide dimensional range. Likewise, cellulose derivatives such as carboxymethylcellulose (CMC) have been widely studied as components in superabsorbent polymer (SAP) formulations. The present review focuses on strategies and mechanisms in which inclusion of cellulose – in its various forms – can enhance either the capacity or the rate of aqueous fluid absorption in various potential applications.

Keywords: Absorbency; Cellulosic materials; Fluff pulp; Superabsorbents; Hydrogels; Carboxymethylcellulose (CMC); Disposable; Rate of absorption; Life cycle analysis

Contact information: Department of Forest Biomaterials, College of Natural Resources, North Carolina State University, Campus Box 8005, Raleigh, NC 27695-8005

INTRODUCTION

Cellulose-based products have been used for the absorption of water and other aqueous fluids throughout recent history (Buchholz and Peppas 1994; Chatterjee and Gupta 2002). Among such products, paper towels and tissue papers are some of the most widely used absorbents (Gigac and Fiserova 2008; Beuther et al. 2010). Fluff pulps, which are generally based on kraft pulping and optimized for high bulk and absorbency products, are in widespread use, especially for product niches such as disposable diapers (Lund et al. 2012). In recent decades the amount of cellulosic fiber employed in making an average disposable diaper, as well as in related products such as incontinence products and feminine hygiene products, has declined; this trend has been attributed to increased use of superabsorbent polymer (SAP) products (Nagorski 1994; Buchholz and Graham 2005; Lund et al. 2012). Thus, the goal of this review is to highlight published literature regarding enhanced absorbent materials and products that incorporate various forms of cellulose or its derivatives. Main sections are shown in the following table of contents:

Table of Contents

Although SAPs have a much greater ability to absorb fluids in comparison to cellulosic fibers alone, they have not displaced cellulose completely, nor is such a complete displacement expected. Cellulose, in its various forms, can contribute many positive attributes to absorbent products, including enhancement of SAP performance. Also, there is a widespread perspective that cellulose-based products are beneficial in terms of the environment and personal health.

Some Definitions

Certain terms will be used repeatedly in this article, e.g. hydrogel and superabsorbent. The term “absorbent” will be defined here by the phrase “capable of imbibing and holding onto aqueous fluids”. “Hydrogel” will be defined as a hydrophilic polymer structure capable of absorbing a lot of water or saline solution. According to Kabiri et al. (2011), an ordinary absorbent hydrogel is able to absorb as much as 10 g of pure water per gram of dry-based material. Hydrogels that can accommodate more water than 10 g/g are called superabsorbents. As will become clear from tabulated information in this review, absorbency levels in the thousands of grams of water per gram of SAP have been frequently reported. Laftah et al. (2011) suggested a possible further distinction between hydrogels and ordinary absorbents such as paper towels: only the hydrogel-based sorbents are able to provide significant resistance to the release of fluid due to squeezing. To paraphrase Lund et al. (2011), hydrogel-based absorbents “swell and restrain” liquids to a remarkable extent. The amount of aqueous fluid held within a gel at equilibrium will be governed by a balance between osmotic pressure, the restraining pressure exerted by the material itself, and any externally applied pressure. The term “osmotic pressure” can be defined as a chemically-induced pressure differential between the inside and the outside of a hydrogel in the presence of a liquid phase. An additional aspect of the definition of a superabsorbent hydrogel is that it needs to be sufficiently crosslinked so that it does not dissolve in the fluid that it is intended to absorb (Zohuriaan-Mehr and Kabiri 2008).

Motivations to Incorporate Cellulose-Based Materials

The most widely used SAP products, to be described in greater detail later in this document, are not readily biodegradable (Dutkiewics 2002; Sannino and Nicolais 2005). Additionally, they are primarily derived from non-renewable resources including petroleum-derived acrylamide (AAm) and acrylic acid (AA) monomers, together with crosslinking agents, most of which can be classified as non-renewable resources (Buchholz and Graham 2005). By contrast, the renewable and biodegradable nature of cellulose is well known (Cooke 1990). Certain derivatives of cellulose such as carboxymethylcellulose (CMC) and hydroxypropyl cellulose (HPC) not only are highly biodegradable, but also, as will be discussed in this article, can form the basis for production of high-performing superabsorbent hydrogels. Various cellulose-based SAPs have been intentionally designed or tested to achieve a suitable degree of biodegradability (Wach et al. 2001; Lionetto et al. 2003; Sahoo et al. 2005; Yoshimura et al. 2005, 2006, 2009; Wang et al. 2008; Sannino et al. 2009; Feng et al. 2010a; Kono and Fujita 2012). Kosemund et al. (2009) emphasized safety and health issues, which are additional factors supporting the use of cellulose-based polymers rather than the use of petroleum-based SAPs.

Costs of materials have been mentioned as a key factor governing the types and compositions of available absorbent products. For instance, Kabiri et al. (2011) reviewed various research studies showing that hybrid superabsorbent structures incorporating montmorillonite “nano-clays” or cellulose-based nanomaterials sometimes achieve superior mechanical properties and can be manufactured at lower costs relative to conventional acrylic-based SAPs. Kono and Fujita (2012) expressed a different view and proposed that a relatively high cost of superabsorbents based on CMC have constituted a key barrier to their widespread use. However, viewed from another perspective, Liang et al. (2009) suggest that the relatively high cost of even the acrylic-based SAPs, in addition to their lack of biodegradability, offers motivation for further development of high-performing cellulose-based SAPs, as well as hybrid products that combine favorable aspects of cellulose and acrylamide-based chemistry.

Since the perspective of the present review article is specialized – focusing on the incorporation of cellulose-based materials in absorbent products – readers are encouraged also to refer to related review articles emphasizing various different aspects of absorbency and absorbents. Information related to absorbent products in general has been amply discussed (Buchholz and Peppas 1994; Chatterjee and Gupta 2002; Buchholz and Graham 2005). Likewise, the somewhat narrower topic of superabsorbents also has been well reviewed (Matushek 2008; Zohuriaan-Mehr and Kabiri 2008; Kabiri et al. 2011; Laftah et al. 2011). The topic of porous hydrogels also has been reviewed (Omidian et al. 2005; Lu et al. 2007).

Regarding the absorbency of products made mainly from cellulosic fibers, Parham and Hergert (1980) provide a review of fluff pulp, including the history of its development. Gupta (1988) reviewed structural factors affecting the performance of nonwoven absorbents. Fink et al. (2001) reviewed the absorbent qualities of certain regenerated cellulose products.

Regarding gel materials, cellulose-based hydrogels have been reviewed by Sannino et al. (2009) and by Chang and Zhang (2011), who emphasized non-superabsorbent materials. Athawale and Lele (2001) reviewed starch-based SAPs. Chitosan, which has a molecular structure that is analogous to that of cellulose, has been reviewed in the context of superabsorbent products (Dutkiewicz 2002). Medical applications of superabsorbent materials have been described and reviewed (Wichterle and Lim 1960; Bourges et al. 2002). Agricultural applications of hydrogels have been reviewed (Kazanskii and Dubrovskii 1992). Theories to account for the capacities and rates of water absorption into hydrogels have also been reviewed (Scallan 1983; Brannon-Peppas and Peppas 1990; Schuchardt and Berg 1991; Wiryana and Berg 1991; Kazanskii and Dubrovskii 1992; Ganji et al. 2010; Zohariaan-Mehr et al. 2010).

Sustainability Aspects

The present review article is motivated in part by issues of sustainability, considering factors that may contribute to the minimization of adverse environmental impacts. In principle, resource sustainability is advanced when renewable, bio-based materials are employed in the manufacture of a product; thus the next section of this article will consider a variety of different absorbent products that have been reported to contain cellulose or its derivatives. Product sustainability is often improved when product goals are achieved with a reduction in material requirements. Thus, separate sections of the article will examine chemical and structural effects on the performance of cellulose-based materials that have been used in absorbent products. Aspects related to the composition of the liquid to be absorbed will also be considered. Theoretical aspects will be considered in sections related to absorption capacity and the corresponding rates of absorption. Finally, the life cycle of absorbent products will be considered, asking such questions as “Why can’t a highly absorbent product be designed so that it empties itself upon immersion in hot water, allowing it to be easily laundered and dried”? Alternatively, one might ask, “Will the next generation of renewable and biodegradable superabsorbents based on cellulose eventually compete successfully with and displace acrylamide-based SAPs?”

CLASSIFICATION OF ABSORBENTS INCORPORATING CELLULOSE

Tabulation

Since the same set of tabulated data will need to be referenced throughout the article, key information is given in Table A, which due to its length is placed in the appendix. The main categories are organized by absorbent class, fiber, hydrogel, “other,” fluid update, “key findings”, and a final column for the literature reference in abbreviated form. Within the “fiber” category, separate columns are provided to show the main fibrous components described in a given cited article and any chemical treatments of that fiber. Within the “Hydrogel” category, there are separate columns for the main chemical components of the hydrogel(s) described in a given article, the method of forming the hydrogels, and the type of crosslinking system employed. Under “Fluid uptake” there are two columns corresponding, respectively, to uptake of pure water and uptake of saline solution roughly corresponding to the ionic concentration of bodily fluids. Such concentrations, e.g. 0.9% NaCl, are often studied due to the widespread availability of physiological saline solutions.

Cellulose-based Materials and Aqueous Absorbency

Mechanically pulped wood

Among wood-based materials, one of the least expensive that could be considered for application as an absorbent would be mechanical pulp, a material derived by the compression and shearing of wood chips in the water-filled gap between the rotor and stator of a refiner. Eriksson et al. (1991) showed that such material can absorb approximately an equal mass of water, even without any specialized treatment. However, the same study revealed a key problem associated with such pulp: The rigid nature of the lignin in mechanically pulped wood resists the swelling of such fibers. One way to overcome such rigidity was demonstrated by removing up to 60% of the lignin. Following such treatment the material was able to take up two to four grams of water per dry gram of fiber. Water uptake also was increased if the temperature of the water was raised above about 70 oC, which is above the softening point of lignin under water-saturated conditions (Back and Salmén 1982).

Chemically pulped wood

According to Beuther et al. (2010), typical paper towels and tissue products that are composed primarily of bleached kraft fibers are capable of absorbing between five and ten times their mass of water without dripping. It was noted that the relative void volume fraction in such paper products can be greater than 0.9 in typical cases. Zohuriaan-Mehr and Kabiri (2008) reported the following values for the absorbency of water by lignin-free products: 1.8 g/g for filter paper; 4 g/g for facial tissue; 12 g/g for bleached kraft fluff pulp; and 19 g/g for cotton balls. The absorptive nature of kraft fibers can be attributed in part to the removal of the more hydrophobic lignin component of the fibers. In addition, the pores left behind by the breakdown and dissolution of lignin during pulping and bleaching gives rise to extensive porosity, though some of the pores close irreversibly when the fibers are dried (Stone and Scallan 1966, 1968; Berthold and Salmén 1997).

Despite their substantially greater absorbency in comparison to mechanical pulp fibers, kraft fibers tend to suffer from one key deficiency: Compared to the mechanically pulped fibers, the wet kraft fibers are not as capable of maintaining a bulky structure, especially when pressure is being applied. Lund et al. (2011) showed that such a deficiency can be overcome by crosslinking kraft fibers with a di-epoxide; the treatment makes the fibers remain stiffer when wet, such that greater bulk is achieved in the paper.

Another approach to enhancing the absorptive ability of kraft fibers has been through the use of rewetting agents (Shepherd and Xiao 1999). The surfactants employed significantly reduced the time required to take up a given amount of water. In addition to enhancing fiber wettability, the treatment often increased the size of inter-fiber pores formed by the structure of the paper. Ström and Carlsson (1992) took a different approach, making the kraft fibers more wettable by oxygen plasma treatment; it was found that such treatment could completely overcome the adverse effect of wood extractives, which otherwise tended to decrease the ability of the fibers to take up water.

Regenerated cellulose

Schuster et al. (2006) reviewed the water-absorbing properties of regenerated cellulose materials. Lyocell fibers, which are formed from a solution of -cellulose in N-methylmorpholine N-oxide (NMMO), were reported to absorb about 5 g/g water. The absorptive ability was attributed to fine-scale porosity, which was extensively docu-mented by means of scanning electron micrographs. The absorption behavior of regenerated cellulose products has been quantified in various studies (Fink et al. 2001; Cai and Zhang 2006; Schuster et al. 2006). Cellulose regeneration from solvent also can be used as a starting point for preparation of hydrogels (Nishio and Manley 1988; Lim et al. 2001, 2003; Saito et al. 2003; Chen et al. 2008; Chang et al.2010; Lin et al. 2010; Chang and Zhang 2011; Kono and Fujita 2012; Kono et al. 2013). This topic will be addressed in later sections

New insights regarding the wettability of regenerated cellulose fibers were reported by Yamane et al. (2006). These authors showed that the wettability of the resulting fibers was greatly affected by the liquid medium present during regeneration. Also, the wettability could be affected, to some extent, by post-equilibration of cellulose films with either a nonpolar solvent, or a polar liquid such as ammonia or hot glycerol. Such effects were attributed to preferential exposure of different crystal faces of cellulose at the surface. Based on molecular simulation, it was predicted that the equatorial vs. the axial sides of glucopyranose rings in cellulose crystallites should exhibit hydrophilic vs. hydrophobic character, respectively. In agreement with the work cited above, Quddus et al. (2012) carried out molecular simulations of the spreading of pure water onto three crystal planes of cellulose; significant differences in contact angle were predicted based on the analysis.

Aerogels and Foams Prepared from Cellulose

Porous superabsorbent materials, i.e. foams, have emerged as an important category of SAPs due to their faster sorption in comparison to conventional SAPs (Yoshinobu et al. 1992; Kabra et al. 1998; Omidian et al. 2005; Lu et al. 2007; Chang et al. 2009a; Zamani et al. 2010; Salam et al. 2011a). However, even without employing SAP technologies (see later), it has been shown that substantial absorbency can be achieved by suitable preparation of highly porous cellulose or its chemical derivatives. Brodin et al. (2012) showed that TEMPO oxidation of kraft pulp followed by freeze drying yielded porous foams that absorbed 21 to 65 g/g of water. Other authors (Pääkkö et al. 2008; Sehaqui et al. 2010, 2011; Theliander et al. 2011) showed that highly porous aerogels could be formed by freeze-drying of nanofibrillated cellulose.

Chemically derivatized fibers

Slight derivatization or grafting of the surface of cellulose is a known approach to changing the wettability and other attributes of cellulosic fibers. For instance, oxidation of fibers by (2,2,6,6-tetramethylpiperidin-1-yl)oxyl (TEMPO) has been found to increase the water uptake of kraft pulp (Brodin and Theliander 2012; Brodin et al. 2012). The effect was attributed to the generation of new carboxylate and aldehyde groups.

More profound effects have been achieved by grafting of polymeric chains by various means onto cellulosic fibers. For instance, improved absorption has been achieved by treating cellulose fiber surfaces with acrylic monomers (Lepoutre et al. 1973; Vitta et al. 1986; Deo and Gotmare 1999; Gürdağ et al. 2001; Sahoo et al. 2005; Aliouche et al. 2006; Jain et al.2007; Jain and Ni 2008). Warner and Rezai (1997) showed that such treatments can be expected to change the composition throughout the fiber structure, not just at its surface.

In cases where a similar grafting approach is carried out more extensively, rather than ending up mainly with surface-derivatized fibers, the cellulose material becomes extensively transformed (Fanta et al. 1987). As shown by Goetz et al. (2009, 2011), such an approach can lead to absorbencies in the range of 30 to 200 g/g, especially if a cross-linking agent is employed. Such results show that it is possible to blur the distinction between what is a surface-modified fiber and what is a true hydrogel. Indeed, a unique class of highly absorbent materials has been achieved by basically skipping the conversion of cellulose to CMC, etc., and just grafting cellulose directly, often in combination with crosslinking (Fanta et al. 1987; Yoshinobu et al. 1992; Gürdağ et al. 2001; Farag and Al-Afaleq 2002; Ibrahim and El-Zawawy 2004; Kim and Mun 2009; Yoshimura et al. 2006; Liu et al. 2009, 2011; Feng et al. 2010a; Zhang et al. 2010; Goetz et al. 2011; Shu et al. 2011; Kono and Fujita 2012; Pan and Ragauskas 2012; Wu et al. 2012; Xie et al. 2012; Kono et al. 2013). Thus, hydrogels will be discussed next.

Hydrogel Formation

As mentioned in the Introduction, an absorbent hydrogel can be defined by its ability to hold a relatively large amount of water (up to 10 g/g for ordinary absorbent hydrogels, or more in the case of superabsorbents), by its resistance to the release of aqueous fluids, depending on the applied pressure, and by its resistance to substantial dissolution, despite being chemically compatible with the fluid being absorbed. Absorbent hydrogels are most often prepared from hydrophilic polymers of relatively high molecular mass (Sannino et al. 2009; Chang and Zhang 2011). A key step in the formation of any hydrogel is “gelation,” which can be interpreted as the inter-connecting of the macromolecular chains in some manner, such that essentially the whole structure becomes linked together. Many different means of gelation have been demonstrated. For instance, one or more types of polymer chains may be precipitated from a suitable solvent by changing the conditions to reduce the solubility (Nishio and Manley 1988; Tang et al. 2010). In some cases progressive gelation can be induced by subjecting a polymer solution to freezing and thawing (El Salmawi 2007). Alternatively, the polymer can be precipitated from the solution by changing the temperature (Joshi et al. 2008). This gelation can be attributed to coalescence of polymeric chains due to thermal break-down of the organized cage-like structure of water surrounding each segment of a molecular chain as the temperature is raised through a critical range (Joshi et al. 2008). In other cases gelation may develop gradually by means of self-segregation of parts of adjacent chains that have differing hydrophobic or hydrophilic tendency (Sammon et al. 2006; Silva et al. 2008). Such a mechanism was supported by shifts in the infrared absorption spectra at the temperatures corresponding to gelation (Sammon et al. 2006). Further support for the coalescence of polysaccharide chain segments is provided by results of X-ray analysis, which revealed considerable crystalline character in certain hydrogels prepared from hemicellulose mixtures with chitosan (Gabrielli et al. 2000).

More aggressive gelation can be achieved by the addition of a suitable agent to form linkages among the chains. One of the simplest approaches, most suitable for the highly soluble polyacrylates and related copolymers, involves the addition of a well-chosen amount of multivalent cation (Kulicke and Nottelmann 1989; Yoshimura et al. 2009). Weerawarna and Su (2010) patented a system in which a mixture of kraft fiber and CMC was treated with aluminum sulfate to precipitate the CMC onto the fiber surface; the inventors characterized this approach as a form of crosslinking. In the present context, however, addition of multivalent cations, e.g. acidic aluminum compounds, is typically considered undesirable due to a strong suppression of swelling (Kulicke and Nottelmann 1989; Feng et al. 2010a).

Covalent crosslinking has emerged as the preferred means of converting suitable hydrophilic polymers into absorbent hydrogels, and many variants of this approach have shown excellent potential for the preparation of superabsorbents (Kabra et al. 1998; Liu et al. 2002; Sannino et al. 2003; Buchholz and Graham 2005; Chang et al. 2008; Zohuriaan-Mehr and Kabiri 2008; Laftah et al. 2011; Pohjanlehto et al. 2011; Kono and Fujita 2012). Notably, the highest levels of swelling are often detected when the extent of crosslinking is just sufficient to cause phase separation (Wach et al. 2001). Optimization of the crosslinking thus involves a balance between achieving high absorbency while also achieving suitable mechanical strength and minimizing the amount of material that dissolves into the bulk of the liquid when a highly absorbent hydrogel is in use (Wach et al. 2001; Guilherme et al. 2005).

Types of Superabsorbent Hydrogels

Acrylic-based hydrogels

To help understand the context of superabsorbent hydrogels that contain cellulosic components, it should be noted that the commercially dominant SAPs do not contain cellulose-based polymers. Acrylic-based and synthetic hydrogels are by far the dominant products manufactured and sold (Askari et al. 1993; Buchholz and Peppas 1994; Matushek 2008; Zohuriaan-Mehr and Kabiri 2008; Laftah et al. 2011). In general terms, acrylic-based SAPs are most often prepared from aqueous solutions containing acrylic acid (AA) and acrylamide (Am) monomers. These are typically crosslinked with N,N’-methylenebisacryamide (MBA) or divinylsulfone (DVS).

Carboxymethylcellulose (CMC)-based hydrogels

Among cellulose-derived materials used in SAPs, CMC holds a lead position (Oppermann 1995; Barbucci et al. 2000; Bin et al. 2000; Fei et al. 2000; Wach et al. 2001; Liu et al. 2002; 2005a; Bao et al. 2012). Crosslinking of the homopolymer can be achieved by gamma irradiation (Bin et al. 2000; Fei et al. 2000; Wach et al. 2001; Liu et al. 2002, 2005a), by chloromethylene iodide (Barbucci et al. 2000), or by polyvalent carboxylic acids such as citric acid (Demitri et al. 2008). Hydrogels formed by mutual gelation or covalent crosslinking of CMC and other water-soluble polymers have been studied (El Salmawi 2007). For instance, superabsorbents have been prepared with combinations of CMC and polyvinyl alcohol (El Salmawi 2007). Alternatively, CMC-based hydrogels have been prepared with the use of highly platy montmorillonite clay (“nanoclay”) to achieve a crosslinking effect (El Salmawi and Ibrahim 2011).

Hydroxyethylcellulose (HEC) and other cellulose-derivative-based hydrogels

In an analogous manner to CMC, other cellulose ethers or esters have been used as a basis for forming hydrogels, either alone or in combination with other polymers (Harsh and Gehrke 1991; Oppermann 1995; Kabra et al. 1998; Bourges et al. 2002; Wach et al. 2002; Lionetto et al. 2003; Demitri et al. 2008; Joshi et al. 2008; Dai and Kadla 2009; Bao et al. 2012). Bao et al. (2012) observed higher absorbency in superabsorbents made from graft copolymers of CMC, in comparison with graft copolymers from hydroxypropyl methyl cellulose, methyl cellulose, and hydroxyethyl cellulose. Like CMC, other derivatives of cellulose and the hydrogels produced from them have been shown to be biodegradable (Lionetto et al. 2003).

Copolymers of CMC by reaction with monomers

Considerable research has been devoted to the development of copolymers in which CMC is reacted with acrylic monomers and various crosslinking agents to form SAPs (Kuwabara and Kubota 1996; Pourjavadi and Mahdavinia 2005; Pourjavadi and Ghasemzadeh 2006; Abd El-Mohdy 2007; Ibrahim et al. 2007; Pourjavadi et al. 2007a-b, 2008, 2009, 2010; Suo et al.2007; Ali et al. 2008; Chen et al. 2008; Yang et al. 2009; El-Din et al. 2010; Feng et al. 2010a; Lin et al. 2010; Wang and Wang 2010; Bao et al. 2011; Eldin et al. 2011; Huang et al. 2011; Liu et al. 2011; Shu et al. 2011; Varaprasad et al. 2011; Wang et al. 2011c). Water absorbency values generally in the range of 40 to 1400 g/g have been reported in such studies, with the highest value reported by Pourjavadi et al. (2008). Copolymers incorporating CMC have also been prepared with polyvinyl alcohol (El Salmawi 2007), hydroxyethyl cellulose (Oppermann 1995; Lionetto et al. 2003; Sannino and Nicolais 2005; Demitri et al. 2008; Dai and Kadla 2009), and with hydroxypropylethyl cellulose (Faroongwarng and Sukonrat 2008).

Cellulose-synthetic graft copolymer hydrogels

Although underivatized cellulose is not nearly as hydrophilic as CMC, several investigators have prepared hydrogels by extensive grafting of other monomers onto cellulose, generally employing suitable organic solvent systems (Yoshinobu et al. 1992; Wu et al. 2008; Chang et al. 2009b; Kim and Mun 2009; Buyanov et al. 2010; Feng and et al. 2010a; Anirudhan et al.2011; Shu and Zhao 2011; Wu et al. 2012). Generally, water absorbency values in the range of 60 to 1800 have been reported for such hydrogels, though Kim and Mun (2009) reported a high value of 2500 g/g. As an alternative approach, Kono and Fujita (2012) showed that it is possible to form a hydrogel from a solution of native cellulose in an ionic liquid, employing a widely used crosslinking agent.

Hydrogels based on cellulose with other natural-based polymers

As one might expect based on environmental concerns, many researchers have investigated the preparation of “all-natural-based” hydrogels incorporating cellulose or its derivatives (Harsh and Gehrke 1991; Barbucci et al. 2000; Bin et al. 2000; Fei et al. 2000; Wach et al. 2001, 2002; Hirsh and Spontak 2002; Liu et al. 2002, 2005a; Rodriguez et al. 2003; Sannino et al.2003, 2004a,b; Sannino and Nicolais 2005; Yoshimura et al. 2006; Demitri et al. 2008; Faroongwang and Sukonrat 2008; Pourjavadi et al. 2008, 2010; Chang et al. 2009, 2010; Dai and Kadla 2009; El Salmawi and Ibrahim 2011; Salam et al. 2011a,b; Wang et al. 2011b; Kono and Fujita 2012; Pan and Ragauskas 2012; Kono et al. 2013). For instance, a hydrogel based on a combination of cellulose and CMC can be prepared from a solution of the ingredients in a NaOH/urea solvent system (Chang et al. 2010).

Major Covalent Crosslinking Agents

Superabsorbent hydrogels that incorporate cellulose-based components, such as CMC, require crosslinking, and many of the crosslinking agents are the same as those used in SAPs not containing any cellulose-based component. Table 1 provides an overview of the most important cross-linking systems that have been employed in studies of superabsorbent hydrogels that have incorporated cellulose-based compounds.

In addition to the agents listed in Table 1, crosslinking also can be achieved by use of dialdehyde cellulose (Han et al. 2010), making it possible to achieve crosslinking with multifunctional amines on lyocel fabric. Swelling was decreased by such treatment, which had utility for imparting crease resistance to clothing.

Interpenetrating

One type of hydrogel that merits special mention employs pairs of non-reacting polymer networks, i.e. interpenetrating networks. In such systems each network relies on a different crosslinking mechanism, and the resulting hydrogels are formed in-situ in such a way that, at least in principle, the two networks can remain intensively entangled. Several reported examples have exhibited high absorption capacity for water (Myung et al. 2008; Chang et al. 2009b; Chang and Zhang 2011; Varaprasad et al. 2011; Liu et al. 2013).

There is reason to doubt, however, that an interpenetrating structure is essential in achieving the improved performances reported for such systems. None of the cited articles provides a convincing explanation of why interpenetrating networks would be expected to absorb more fluid. Logic would suggest that such complex, entangled structures would result in more resistance to swelling in comparison to having just one network, in which the crosslinking degree and molecular mass had been suitably optimized. It seems likely that the cellulose-based components of such systems may act as “spacers”, partly suppressing the crosslinking within the acrylamide-based component. Further research is also needed to find out whether both networks are truly continuous in such systems.

Table 1. Major Crosslinking Agents Used in Studies of SAP Preparation

Polyelectrolyte complexes

A unique type of hydrogel can be formed through the use of pairs of oppositely charged polyelectrolytes via the formation of polyelectrolyte complexes (PECs). At least one of the polyelectrolytes could be a cellulosic material such as carboxymethyl cellulose (CMC) (Feng et al. 2006; Feng and Pelton 2007) or hemicellulose (Salam et al. 2011b). Though PECs generally do not show promise for the absorption of large amounts of pure water, various studies have shown favorable results for the absorption of saline solutions, possibly representing urine (Paula et al. 2002; Feng et al. 2006; Feng and Pelton 2007). Despite the fact that under many circumstances the ionic interactions are sufficient to maintain the integrity of such gels, some researchers have reported enhancement of hydrogel capabilities when covalent crosslinkers have been applied in the preparation of such PEC hydrogels (Chen and Tan 2006; Shang et al. 2008; Salam et al. 2011b). As will be discussed in more detail later, the ability of some such hydrogels to increase their absorbency with increasing salinity can be attributed to a salt-induced weakening of the ionic pair-wise interactions that hold PECs together.

Products Combining Cellulosic Fibers with SAPs

A cursory inspection of currently available disposable absorbent products reveals a combination of cellulosic fibers and superabsorbent hydrogel materials, along with various barrier layers, permeable layers, elastic elements, and closures (Nagorski 1994). Within such products, conventional cellulosic fibers, i.e. the fluff pulp, play a variety of roles. Though, as mentioned earlier, bleached kraft fibers can absorb about 12 g water per g or solid (Zohuriaan-Mehr and Kabiri 2008), this is not enough to compete with the much higher water-holding capacity of the SAPs. Rather, the fluff pulp is used primarily to provide structure and wicking (Schuchardt and Berg 1991). In particular, the addition the fluff pulp can prevent gel blocking, which is the blocking of liquids by a complete wall of saturated SAP particles that are not separated sufficiently by the fibers. Masoodi et al. (2012) carried out studies directed at examining the combined effects of cellulose fiber and CMC hydrogel powder with respect to the wicking and uptake of water. Numerical simulation was used to model the phenomenon observed in the experiment. Though conventional assumptions of flow through capillary structures appear to be valid, the permeability was found to decrease with time (Schuchardt and Berg 1991; Wiryana and Berg 1991; Mosoodi et al. 2012). It was proposed in the cited studies that swelling of the hydrogel particles tended to seal off channels within the composite structure.

Weerawarna and Bing (2006) claimed a process in which superabsorbent gels were chemically attached to suitable fibers. Schuchardt and Berg (1991) describe various practical ways in which SAPs have been incorporated into cellulosic pad structures, including the various shapes and sizes of hydrogel particles. The distribution of SAPs within the structure is clearly an important consideration that will merit further study.

Integration of Cellulosic Solids into SAPs

An alternative way in which cellulose can be incorporated into highly absorbent products is by including solid cellulose in the mixture during gelation of the SAP. This approach has been applied by Larsson et al. (2010, 2011), with a focus on nanocellulose product applications. These authors observed that the effects of adding nanocellulose, in terms of both swelling and mechanical properties, was approximately the same as the addition of the same mass of conventional crosslinking agent. This suggests that the cellulosic fibrils, by being relatively long, tended to act as reinforcement, forming a nanocomposite structure (Hubbe et al. 2008; Pandey et al. 2013).

The effects of adding nanocellulose during the preparation of SAPs, as just described, are strikingly in contrast to various results that have been obtained when alkali montmorillonite, i.e.“nanoclay”, or other finely divided minerals have been added to the formulation during the preparation of SAPs (Pourjavadi and Ghasemzadeh 2006; Pourjavadi et al. 2008; Dai and Kadla 2009; Zhang et al. 2008b; 2009; Wang and Wang 2010; Wang et al. 2010, 2011a,b,d; Anirudhan et al. 2011; Bao et al. 2011; Chang and Zhang 2011; El Salmawi and Ibrahim 2011; Kabiri et al. 2011). In some cases the inclusion of nano-sized mineral particles during hydrogel preparation yielded a substantial increase in water absorption capacity (Pourjavadi et al. 2008; Zhang et al. 2008b; 2009; Wang and Wang 2010; Wang et al. 2010, 2011a,b,d; El Salmawi and Ibrahim 2011), whereas in some other cases the nanominerals tended to suppress absorbency (Pourjavadi and Ghasemzadeh 2006; Dai and Kadla 2009). A possible mechanism to account for the enhancement of absorption, in the cases indicated, could be a debonding effect, allowing pools of fluid to occupy openings that are made available by poor bonding of the adjacent SAP material to a suitably sized mineral surface. In addition, progressive exfoliation of montmorillonite may provide additional capacity for the holding of fluid (Knudson 1993). Alternatively, in cases where the minerals happen to be better bonded to the matrix, they could be expected to restrain swelling and imbibition.

CHEMICAL EFFECTS

Various published studies have addressed the question, “In what ways can the chemical composition of absorbent material affect its absorbency?” This section first considers publications dealing with the absorbency of cellulose and related biomaterials, followed by various copolymer derivatives, hydrogels, and combined structures.

Cellulose as an Absorbent Material

The chemistry of cellulose itself clearly can contribute to absorbency on account of the numerous hydrophilic –OH groups throughout its structure. The most widely used tissue and towel products are typically made from delignified wood fibers; in other words they are mainly comprised of cellulose and hemicellulose (Gigac and Fiserova 2008). As documented in Table A, the absorbency of ordinary delignified cellulosic fibers often lies in the range between 2 and 10 g/g, based on tests with pure water (Shepherd and Xaio 1999; Gigac and Fiserova 2008; Zhang et al. 2008a; Beuther et al. 2010). Mechanical pulp fibers, which contain essentially all of the relatively hydrophobic and rigid lignin that was originally present in the wood, tend to absorb less water in comparison to kraft or sulfite (delignified) fibers (Eriksson et al. 1991). Breier (1997) reported corresponding values of 0.5 g/g for cotton fiber and 0.6 to 1 g/g for various regenerated cellulose products.

The somewhat higher reported performance of wood-derived fibers, especially after delignification, probably can be attributed to their mesoporous nature and/or internal delamination as a result of mechanical treatments. Also, as will be discussed in detail later in this article, drying of cellulosic material can make a great difference in terms of its pore structure and capacity to absorb water.

Cellulose crystallinity

Though the outermost surfaces of cellulose crystalline regions are able to interact with water, the crystalline domains within the cellulosic material are not expected to imbibe any water molecules (Bergenstrahle et al. 2008). Rather, it is the amorphous regions of cellulose, along with spaces occupied by the non-crystalline hemicelluloses, that are expected to swell and accommodate water (Heiner and Teleman 1997; Heiner et al. 1998; Bergenstrahle et al. 2008; Kocherbitov et al. 2008). These regions are represented schematically in Fig. 1. In support of this type of simplified model, it has been shown that the treatment of cellulose with ZnCl2 solution is able to disrupt the crystallization to some extent, an effect that can greatly increase the amount of water absorption (Williams and Stannet 1979; Vitta et al. 1986).

Fig. 1. Representation of crystalline regions and amorphous regions within the nanostructure of delignified fibers from wood

Absorption in the amorphous regions of cellulose, as well as within the domains containing hemicellulose, can be expected to involve strongly bound water, i.e. a form of hydration (Joubert et al. 1959; Gert 1996). In support of this statement, a strong exotherm has been observed accompanying the first increment of water uptake onto completely dry cellulose (Kocherbitov et al. 2008). Water that is strongly bound to cellulose has also been characterized as “hard to remove water” as measured by thermogravimetric analysis or as freezing or non-freezing bound water as measured using differential scanning calorimetry (Park et al. 2006, 2007; Sen et al. 2012).

In the case of conventional cellulosic fiber materials (not considering hydrogels), a majority of the retained water within a pad or sheet product is likely to reside between fibers rather than within them (Jerkeman 1970). Issues relative to the bulkiness of cellulosic absorbent materials will be discussed in a later section.

Hemicellulose and Absorbency

The swelling of cellulosic fibers has been found to correlate positively with the proportion of hemicellulose (Eriksson et al. 1991). The positive contribution of hemicel-lulose to swelling can be partly attributed to its irregular structure, with branching and side groups (e.g. as shown in Fig. 1). The irregularity of hemicellulose tends to discourage its crystallization (Billmeyer 1971; Brydson 1975). Peura et al. (2008) observed levels of crystallinity only in the range 16 to 19% in films cast from purified xylan. Likewise, Gabrielii et al. (2000) and Gröndahl et al.(2004) observed somewhat higher levels of crystallinity in films from aspen wood xylan prepared under conditions optimized to promote crystallization. A second factor that contributes to the hydrophilic nature of hemicelluloses is the content of carboxyl groups, which can dissociate to the carboxylate form at intermediate to high conditions of pH (Lindström and Carlsson 1982). Indeed, Salam et al. (2011a) showed that hemicellulose could be rendered even more hydrophilic by derivatizing it to further increase the content of carboxyl groups.

Extraction of hemicellulose from kraft fibers with alkaline solutions has been found to increase the water absorptivity of the fibers (Lund et al. 2012). This is somewhat surprising, given the fact that hemicelluloses have a distinctly hydrophilic nature. The explanation appears to lie in the conversion of other ionic forms of the remaining hemicellulose to the sodium form, which is known to enhance swelling (Lindström and Carlsson 1982). The importance of hemicellulose was also shown by Laine and Stenius (1997), who demonstrated a positive relationship between fiber ionic charge and swelling with water. Recently several researchers have attempted to take advantage of the hydrophilic nature of hemicelluloses by using them as a key component of highly absorbent hydrogels (Gabrielli and Gatenholm 1998; Gabrielli et al. 2000; Meena et al. 2011; Peng et al. 2011; Pohjanlehto et al. 2011; Salam et al. 2011a; Yang et al. 2011; Karaaslan et al. 2010, 2011, 2012).

Lignin and Absorbency

Compared to the polysaccharides in wood, lignin is more hydrophobic. Obataya et al. (2007) found that delignification of wood increased its tendency to swell in water. Likewise, as reported by Parham and Hergert (1980), delignified kraft pulps can be expected to absorb significantly more water than mechanical pulps, in which most of the native lignin still remains after pulping. Further demonstration of the relative hydrophobicity of lignin has been shown by studies in which lignin or its derivatives were precipitated onto cellulose, thus decreasing the substrate’s wettability by water (Sipponen et al. 2010; Li et al. 2011; Payne et al. 2012). One way to overcome the adverse effect of lignin on swelling is to heat lignified fibers above approximately 70 oC (Eriksson et al. 1991). Such heating, in the presence of water, was observed to soften lignin sufficiently so that the water uptake was increased.

Extractives and Absorbency

Because the term “extractives” usually implies that a substance can be removed from cellulosic material with an organic solvent (Back and Allen 2000), it follows that most such substances are likely to resist water wetting and penetration. When microfibrillated cellulose materials of different lignin and extractives contents were exposed to an organic solvent to remove any extractives, the water-cellulose interactions as measured by water contact angle all increased, with the largest increases observed for the material with the most extractives initially (Spence et al. 2010). However, a search of the literature revealed a paucity of studies attempting to correlate levels of extractives and extents of swelling or absorbency. As already noted, Lund et al. (2012) observed an increased swelling of softwood kraft fibers after their extraction with NaOH. Though the authors attributed this effect to the removal of hemicelluloses, it seems likely that at least part of the effect may have been due to removal of certain extractive materials, e.g. the salts of fatty and resin acids.

Ström and Carlsson (1992) found that an adverse effect of extractives on the water-wettability of kraft pulps could be overcome by treatment with oxygen plasma. The effect was attributed to oxidation of the exposed lignin. Related results have been published by Uner (2009), who found that corona discharge treatment of paper resulted in increased absorbency. Another way to increase wetting and absorbency involves use of suitable surfactants, i.e. rewetting agents (Shepherd and Xiao 1999).

Hydrophilic Substituent Groups and Absorbency

The presence and increasing level of hydrophilic groups in a cellulose-based material can contribute to water uptake in a number of ways, e.g. through hydrogen bonding with the water, by electrostatic effects (if the groups are charged), and by somehow interfering with the densified structure of the biomaterial (Lepoutre et al. 1973). Effects of several kinds of water-loving ionic groups are considered below.

The relationship between two key types of hydrophilic groups and the transient hydrogen bonds within liquid water is illustrated in Fig. 2. Hydroxyl groups exposed at the surface of an absorbent material can be expected to form hydrogen bonds with adjacent water molecules, as shown in the left frame of the figure. The hydrogen bonding extends, in a continuously changing pattern, throughout the whole liquid phase (Drost-Hansen 1969). The right-hand frame of the figure emphasizes the role of bound ionic groups, which can be expected to have a strong affinity for the polar molecules of water. As shown, the bound negatively charged groups can be expected to interact strongly with the positively charged pole of adjacent water molecules. The water molecules tend to orient themselves in a transient pattern, consistent with their polar nature.

Fig. 2. Representation of hydrogen bonding among the molecules of liquid water and also associated with hydrophilic groups. Left: Hydrogen bonding associated with hydroxyl groups bound to a substrate. Right: Interaction and orientation of polar H2O molecules with ionized groups such as carboxylate or sulphonate groups, all associated with the transient hydrogen bonded structure of liquid water

Oxidation

By low levels of oxidative treatment it is possible to increase the proportion of carboxylic acid and/or aldehyde groups on the surfaces of cellulosic fibers. Such oxidative treatment (e.g.TEMPO-mediated oxidation) has been shown to greatly increase water absorbency of softwood kraft pulps (Brodin et al. 2012; Brodin and Theliander 2012). Periodate oxidation has been shown to offer an alternative means of achieving similar effects (Zhang et al. 2008a).

Carboxymethylation and succinylation

Chemical derivatization is another way to increase the number of carboxyl groups on cellulosic surfaces. The idea is to covalently attach groups containing carboxyl functions to the cellulose. In particular, it has been found that carboxymethylation of fibers can greatly increase absorbency even when the treatment level is low enough to avoid solubilization of the source cellulosic material (Schuchardt and Berg 1991). As already noted, however, it is more common to employ extensive carboxymethylation of cellulose, an approach that allows use of fully soluble CMC as a basis for forming superabsorbent hydrogels through crosslinking (Zohuriaan-Mehr and Kabiri 2008; Sannino et al. 2009; Chang and Zhang 2011). Carboxylic acid groups also can be attached to cellulosic fibers by means of reaction with succinic anhydride or maleic anhydride; such treatment likewise has been shown to increase absorbency (Yoshimura et al. 2005, 2006).

Sulfonation

Zhang et al. (2008a) showed that sulfonate groups can be especially effective to increase the water retention values of bleached hardwood kraft fibers. It was noted that because sulfonic acid is a strong acid, a high proportion of such groups tend to retain their negative charge at any given moment, even when the charge density of a fiber surface is relatively high. The acid groups remain dissociated throughout the pH range where such fibers are likely to be used. Hou et al. (2008) also showed that sulfonation improved absorbency.

Sulfonate groups also have been shown to be advantageous in the preparation of high-performing superabsorbent hydrogels. For instance, sulfated starch has been used as a polyelectrolyte component in the preparation of high-performing SAPs (Lim et al. 2001). The improved performance, relative to similar hydrogels containing just carboxylate groups, again was attributed to a tendency of the strong acid groups to remain fully dissociated despite the relatively high levels of substitution in the starch-based polyelectrolyte.

Phosphorylation

Saito et al. (1994a, 1995) showed improvements in absorbency analogous to those of the sulfate group when using phosphorylated lignocellulosic materials. The phosphate was characterized as “a water-attracting group” (Saito et al. 1995). However, in their earlier article the authors had concluded that effects of preparation conditions on surface area and porous structure were more important relative to the content of ionizable groups (Saito et al. 1994a). In a third study the same authors attributed the increased absorbency to inter-crystalline swelling (Saito et al. 1994b).

Hydrophobic Substituent Groups and Absorbency

Cellulose-based materials are in general hydrophilic, and it has been found that a de-swelling effect can be achieved by derivatization with hydrophobic groups. Such an effect was shown most clearly by Koroskenyi and McCarthy (2001), who used acetylation to systematically decrease the swelling behavior of glucomannan. Obataya et al. (2007) showed that acetylation increased the ability of lignin-containing fibers to absorb non-polar solvents. One exception to the rule has been observed when cellulose fibers are derivatized at a low level, e.g. 5% substitution of the hydroxyl groups by acetyl groups; in such cases Ehrnrooth et al. (1977) reported increased swellability, which was attributed to the reduced crystallinity of the partially substituted cellulose.

Hydrogel Attributes and Absorbency

Considerable attention has been devoted to the absorbency of superabsorbent hydrogels in light of their ability to take up 10 to over 1000 times their mass in water. Various factors involving the structure of hydrogel composition and molecular structure appear to control absorbency. Some of the key factors, as described below, are the molecular mass of the component polymers, the density of crosslinking, and the nature of the chemical species used in the crosslinking. In addition, the presence and density of ionizable groups on polymer chains, which are physically unable to diffuse out of the gel, provide a driving force for osmotic effects, inducing more water to enter the gel.

Molecular mass

High molecular mass has been proposed as being a key requirement of polymers that are intended as the main components of superabsorbent hydrogels (Yoshimura et al. 2006). Thus, Kono and Fujita (2012) observed increased absorbency with increasing molecular mass of the cellulose used to prepare a crosslinked CMC hydrogel from an ionic liquid solution. In principle, if the molecular mass is relatively low, then there will be a greater likelihood of dissolution and loss of the polyelectrolyte from the hydrogel during its use. Though a higher degree of crosslinking could be used to overcome such deficiencies, such an approach would not be expected to yield high levels of absorbency (see next).

Crosslinking

In the absence of any crosslinking, many of the most widely used superabsorbent products would simply dissolve when placed in water. Though the highest swelling performance is often observed at the lowest levels of crosslinking (Bin et al. 2000; Wach et al. 2001, 2002; Liu et al. 2005a; Demitri et al. 2008; Han et al. 2010; Pourjavadi et al. 2010; Kono and Fujita 2012), it is generally observed that higher levels of crosslinking (or other reinforcement) are needed to achieve the needed mechanical resilience and resistance to dissolution of superabsorbent products (Guilherme et al. 2005; Myung et al. 2008; Tanaka et al. 2005; Schexnailder and Schmidt 2009; Zohuriaan-Mehr et al. 2010).

Somewhat contrary results were reported by Barbucci et al. (2000). Although these authors observed a monotonic decrease in swelling with increase in crosslinking up to a 50% level, they observed a higher swelling at “100% crosslinking”. The anomalous behavior was attributed to disruption of hydrogen bonding by the crosslinking agent. Anomalous increases in swelling with increasing crosslinking also have been observed when the crosslinking agent is much more hydrophilic than the polymer chain. Thus, Oliveira et al. (2013) showed that increasing crosslinking of cellulose acetate (a hydrophobic polymer) with a di-anhydride increased water sorption. It was noted that the curing of a dianhydride can form two ester bonds and two free carboxyl groups, the latter of which increase the ionic character of the hydrogel at intermediate pH values (Kono and Zakimi 2013; Oliveira et al. 2013).

Numerous studies have addressed the goal of finding an optimum level of crosslinking for various types of superabsorbent gels (Ali et al. 2008; Pohjanlehto et al. 2011; Kono and Fujita 2012). Kono et al. (2013) observed that the balance between simple grafting and crosslinking reactions is sometimes sensitive to the choice of solvent system. As one means to control this balance, Lionetto et al. (2005) showed that the degree of crosslinking can be monitored during the process by evaluation of ultrasonic wave propagation.

Chang et al. (2008) showed that the type of crosslinking, i.e. covalent-type vs. physical crosslinking, also can make a big difference relative to results. For the observed system it was found that chemical crosslinking yielded higher absorbency. In the cited case the chemical crosslinking approach made it possible to avoid use of multivalent cationic agents. As will be discussed later, such species generally tend to reduce the swelling of typical hydrogels.

Cross-link length

Several innovative studies have shown that, at least in principle, higher absorbency of hydrogels can be achieved by using longer, flexible crosslinking agents (Xiao and Meshitsuka 2001; Sannino et al. 2003; Chang and Zhang 2011; Kabiri et al. 2011; Pohjantehto et al. 2011). In practice this is achieved by inserting a “spacer” between the two reactive groups of a suitable crosslinking agent (Chang and Zhang 2011). The spacer within the crosslinked structure of a SAP can be expected to allow further expansion and therefore greater uptake of aqueous fluid. It is notable that analogous results were obtained when paper was treated with high- vs. low-mass polyvalent carboxylic acid species; the longer molecules gave rise to greater tensile strength and folding endurance (Yang and Xu 1998; Xu and Yang 1999).

There is an interesting parallel between the effects of spacers and the effects of certain very fine mineral products, which were mentioned earlier (Pourjavadi et al. 2008; Zhang et al.2008b; 2009; Wang and Wang 2010; Wang et al. 2010, 2011a,b,d; El Salmawi and Ibrahim 2011). Such nano-minerals, if poorly adhered to the adjacent polymeric material, can be expected to allow the opening up of fluid-filled spaces within the hydrogel. The presence of void spaces in such nanocomposite structures is supported by observations of relatively high opacity in some cases (Dai and Kadla 2009). It seems likely that some of the high levels of water uptake reported for SAPs based on interpenetrating networks (e.g. Çaykara et al. 2006; Chang et al. 2009b) may be due to one network interfering with the crosslinking of the other network during its preparation.

Ionizable Groups and Absorbency

Hydrogels that contain carboxylic groups generally exhibit increasing absorbency with increasing pH, especially within the pH range of about 3 to 7; the following examples are from studies involving cellulose-based materials in SAPs (Barbucci et al. 2000; Harsh and Gehrke 1991; Sannino and Nicolais 2005; Pourjavadi et al. 2008; Anirudhan et al. 2011; Bao et al.2011; Wang et al. 2011a). The effect has been attributed to dissociation of the groups to the carboxylate form (Barbucci et al. 2000). As noted by Harsh and Gehrke (1991), such effects of pH are consistent with a principle that absorbency increases with increasing charge density. While most such studies have employed pH as the main independent variable to manipulate the charge density, an alternative approach can be achieved by adjusting the composition of the polyelectrolytes used in forming the hydrogel. Thus, Wu et al. (2012) determined that absorbency increases with increasing level of charged monomer groups up to a 70% of acrylic acid in a cellulose graft copolymer hydrogel with acrylamide monomers and a crosslinker. Considering the importance of the topic, more research is recommended.

Particle Size of Hydrogel

Based on limited available data, the size of hydrogel particles can affect the rate of absorption of water, but the effect on absorption capacity appears to be weak (Tanaka and Fillmore 1979). The cited study dealt with acryalamide gels formed at a 5% concentration of polymer in water. The cited authors observed that the rate of swelling was inversely proportional to particle size, as would be expected according to a diffusion-limited process. Bao et al. (2011) found only a weak and inconsistent relationship between equilibrium absorption and particle size in the case of superabsorbent hydrogels based on CMC, acrylamide copolymer, and montmorillonite nanoparticles.

STRUCTURAL EFFECTS

In principle, structural aspects of various hydrogels can affect not only the absorption capacity but also the rate of filling. The focus here is on the effects of various cellulosic components on absorbent structure and performance in different situations. One can consider three size ranges in which important effects may be manifested, i.e. due to the presence of cellulosic fibers, due to nanocellulose, and due to the inherent differences in stiffness of different types of polymer chains within hydrogels.

Macroscopic Structural Effects

Cellulosic fiber use

Fluff pulp, which is usually provided as bleached softwood kraft fibers, is typically a main component of highly absorbent products (Jerkeman 1970; Parham and Hergert 1980; Nagorski 1994; Buchholz et al. 2005; Lund et al. 2012). Although, as noted earlier, the fibers themselves can accommodate some liquid (Zohuriaan-Mehr and Kabiri 2008), most of the absorption capacity of such products can be attributed to the inclusion of SAPs (Aliouche et al. 2001). The fibers are used in order to provide integrity to the product (Brodin and Theliander 2012) as well as to accommodate more rapid wicking of fluid throughout the absorbent and to prevent blocking (Buchholz et al. 2005).

According to Buchholz et al. (2005) a network of cellulosic fibers intertwined through a loose matrix of SAPs in an absorbent product can be expected to restrict the swelling of the SAPs. A constraining pressure of about 2 kPa was estimated in the cited work. It follows that the fibers within a SAP structure may decrease the absorption capacity; however, the fibers might be needed to maintain structural integrity and to allow quicker distribution of liquid through the product. A typical composition for a disposable diaper has been cited as 43% fluff pulp, 27% SAP material, and the remainder being various plastic materials and adhesives (Edana 2005).

Refining of fibers and absorbency

Refining refers to the repeated mechanical compression and shearing of a slurry of cellulosic fibers in the gap between the raised bars of a stator and a rotor (Smook 1992). The usual intention of refining is to cause internal delamination of the cell wall structure and thereby to increase the flexibility and conformability of the wet fibers. Refined fibers can be expected to develop a greater relative bonded area, hence achieving higher strength of paper products (Page 1969; Lowe et al. 2007). However, when high absorbency is the key goal, studies have shown that the extent of refining should be kept low (Gigac and Fiserova 2008). The cited authors observed that the manner of refining, in addition to the applied energy, can be important; a high specific edge load (generally obtained by using refiner plates with a relatively coarse pattern) resulted in higher absorbency of the resulting tissue paper. These effects can be attributed to a more voluminous packing of fibers that have become bent as result of harsh but limited refining treatment. Due to larger void volumes, more water was able to be absorbed into the tissue paper.

Bulk

Consistent with the preceding discussion, factors that tend to reduce the packing density of a structure can be expected to favor more rapid flow into products made from cellulosic fibers. Though mechanical pulp fibers individually tend to have lower absorbency than bleached kraft fibers, they also tend to be stiffer and less compressible, often yielding a bulkier structure of paper. The relationship between fiber stiffness and the resulting paper’s apparent density was demonstrated in a fundamental manner by Steadman and Luner (1985). Accordingly, Brodin et al. (2012) observed higher absorbency for pads composed of chemithermomechanical pulp (CTMP) fibers in comparison with kraft fibers. In the case of kraft pulps, Lund et al. (2011) showed that increased bulk could be achieved by crosslinking the fibers with a diepoxide. Such treatment would find application in products that need to resist wet compaction, as might happen in the case of a wet diaper holding up the weight of a child.

Another way to increase the bulkiness of a dry paper structure is to treat the fiber suspension with a suitable cationic surfactant before the drying process. Fatehi et al. (2010) showed that treatment with such “debonding agents” rendered the resulting tissue paper less dense. Likewise, Shepherd and Xiao (1999) showed that surfactant treatment of wet pulp resulted in an increased effective pore size and hydrophilicity of fibers once the paper had been dried. Cationic surfactants yielded the largest increase in the rate of wetting.

Another factor that is sometimes overlooked when considering the absorption capacity of a cellulose-based product is the swelling of the fibers themselves upon wetting, an effect that can change the dimensions of the structure as a whole. Bristow (1972) introduced this concept as a means of accounting for the absorption capacities of fiberboard products. Factors affecting the swelling of cellulosic fibers have been considered in various studies (Jayme and Büttel 1964, 1968; Scallan and Carles 1972; Gigac and Fiserova 2008; Zhang et al. 2008a; Cheng et al.2010; Lund et al. 2011).

Channeling

Cellulosic fiber webs, especially if the fibers have a relatively coarse and bulky nature, appear well suited to achieving a relatively rapid uptake of water (Beuther et al. 2010). Thus, as noted by Rezai and Warner (1997), combinations of SAPs and cellulosic fibers can be used in an attempt to achieve the best results from each component – relatively rapid wicking by the fibers and high absorbency contributed by the hydrogel. Schuchardt and Berg (1991) showed that, to a first approximation, well-established equations of flow through packed beds can be used to model the flow of aqueous fluids through mixtures of cellulosic fibers and superabsorbent hydrogel particles. However, it was apparent that swelling of the SAPs tended to progressively restrict the flow as a function of time. A correction term, i.e. a permeability factor, was defined to account for this effect. To make it less likely that swollen SAP particles can seal the passages adjacent to a cellulosic fiber, one could select fibers having an irregular cross-section (Pramanick and Crouse 1998). For instance, certain viscose fibers have been observed to be multi-lobal in cross-section (Schuster et al. 2006). The use of twisted, curly fibers for absorbent products has been patented (Sun and Lindsay 2003; Hu and Ko 2006).

As an alternative strategy to achieve high flow rate within a relatively large volume of superabsorbent material, certain authors have prepared superabsorbent fibers. These fibers tend to pack less densely than the same mass of rounded particles. For example, Feng et al. (2010b) spun blends of cross-linked polyacryonitrile and cellulose acetate as a means of producing highly absorbent fibers. Related results were achieved by Goetz et al. (2011) by extensive grafting of softwood kraft fibers with poly(methyl vinyl ether-co-maleic acid) and polyethylene glycol. Both studies achieved SAP-like levels of water absorbency, while retaining the bulky bed structure inherent in the use of suitably stiff fibers.

Pore structure

Another way in which to consider aspects related to the bulky structure of cellulose-based sorbents is to examine the effects of pore spaces. Such spaces can vary greatly in size, size distribution, and connectivity. Pores also can be quantified in terms of their net volume relative to the solid content. For instance, Beuther et al. (2010) estimated that a typical sheet of tissue paper can have a relative void volume greater than 0.9, consistent with the ability to pick up as much as ten times its mass in aqueous fluids. Brodin et al. (2012) used porosimetry to quantify the pore size distribution within an absorbent foam; by this means it was shown that inclusion of relatively stiff CTMP fibers resulted in greater preservation of pore volume when a mixed absorbent structure was compressed. Studies by Sannino et al. (2004b, 2005) showed that the porous nature of hydrogel foams could be systematically adjusted by varying the conditions of precipita-tion and foam formation.

At the nano-scale, as noted already, it has been proposed that a significant amount of water contained in a hydrogel may be present as small “lakes” rather than as chemically bound water (Gelin et al. 2007). In other words, the distribution of water may be uneven, when considered on a nano-scale. Hydrogels formed in the presence of nano-sized micelles of surfactant molecules have been shown to have enhanced porosity; such effects have been shown to nearly double the absorption capacity for water in some cases (Shi et al. 2013). In the case of chemically pulped fibers, effects related to porosity also extend to the nano-scale. A mesoporous structure within various kinds of cellulose material has been well documented (Stone and Scallan 1968; Schuster et al. 2006). Also, it has been proposed that grafting of such fibers with hydrophilic polymer chains may serve to brace adjacent microfibrils apart from each other, thus contributing to the swellability and ultimately to the volume available for sorption of fluids (Lepoutre et al. 1973).

At a molecular scale it has been proposed that the inherent stiffness of the cellulose polymer chain can lead to a firm hydrogel structure that is less susceptible to collapse of foam structures (Chang et al. 2009a; Zhang et al. 2010). As a likely explanation, the apparent stiffness may be due to cellulose’s tendency to self-associate into microfibrils, i.e. semi-crystalline domains, during the process of gelation (Chen et al. 2006; Kondo et al. 2008). A substantial presence of crystalline domains has been detected even in the case of hydrogels based on hemicellulose (Karaaslan et al. 2010, 2012). Nishio and Manley (1988) showed that the presence of polyvinyl alcohol tended to suppress the formation of crystallinity in cellulose when it was regenerated from an ionic liquid solution. It is reasonable to expect that such microfibrils and partially crystalline character play a role, especially when hydrogels are prepared from solutions of either cellulose itself or of derivatives such as CMC having a low enough degree of substitution as to still permit local, well-organized self-association of chain segments (Thuresson and Lindman 1999; Saito et al. 1994a,b; Bhandari et al. 2012). Bhandari et al. (2012) found an inverse relationship between the swelling tendency and the degree of crystallinity within cellulose films regenerated from ethanol-water-NaOH solutions.

Nanocellulose and Absorbency

During recent years there has been a high level of research activity devoted to the preparation and characterization of very finely separated cellulosic fibrillar or elongated crystalline particles (Klemm et al. 2011). Likewise, there has been related patenting activity (Duran et al. 2012; Charreau et al. 2013). Cheng et al. (2010) showed that certain nanocellulose materials by themselves achieved water retention values in the range of 3.4 to 3.5 g/g. Hua et al. (2011) patented a production system for nanocellulose and described their use in absorbent products.

Nanocellulose materials can be formed into highly porous, absorbent structures by freeze-drying of their aqueous suspensions. Thus, Pääkkö et al. (2008) showed that native cellulose nanofibrils, obtained by enzymatic and mechanical processing of bleached sulfite softwood fibers, yielded strong and voluminous aerogels upon freeze-drying. Immersion of an aqueous suspension of nanofibrillated cellulose in liquid propane, followed by vacuum pumping, yielded an aerogel structure that shrunk by only 7%. A practical, short-cut method was developed in which aerogels were formed by direct vacuum pumping (and induced freezing) of the initial aqueous suspensions. The latter method caused 20% shrinkage. Brodin et al. (2012) explored the use of microfibrillated cellulose in combination with macroscopic cellulose fibers in a freeze-dried structure; absorbencies in the range of 21 to 65 g/g were observed. The composite structure displayed better mechanical stability of its foam structure, as well as higher absorbency. Oxidation was used in the cited work to further increase the absorbency of the fiber component of the structure. Kettunen et al. (2001) reported an absorbency of 35 g/g in the case of freeze-dried nanocellulose aerogels.

Nanocellulose also has been incorporated into hydrogel products (Abbas et al. 2011; Karaaslan et al. 2011; Gane et al. 2012). The mechanical integrity of such hydrogels is generally improved by such addition (Karaaslan et al. 2011), suggesting that the elongated nanocellulose particles can have a function similar to that of a cross-linking agent, or probably more precisely as nano-reinforcement of the gel structure. A tight, relatively defect-free attachment between cellulose nanofillers and a hydrogel matrix has been demonstrated by a high transparency of such nanocomposite hydrogels (Dai and Kadla 2009). This was in contrast with the effect of nanoclay usage, which resulted in opaque hydrogels.

Maintaining meso-pore structure

Cellulosic nanomaterials as part of a hydrogel structure may help to alter or maintain the structure of pores. Spagnol et al. (2012a) found that incorporation of cellulose nano-whiskers (i.e.long crystals) tended to increase the pore size of starch-acrylic hydrogels. Relatively low levels of nanocellulose addition tended to create defects in the structure, thus decreasing the elastic modulus. However, amounts greater than 5% yielded a net increase in strength, suggesting the presence of a contiguous network of cellulosic elements within the matrix (Favier et al. 1997; Chakroborty et al. 2006a,b; Hubbe et al. 2008).

Effects on swelling

Effects of the presence of nanocellulose on the swelling of SAP composites has been considered in several studies (Buyanov et al. 2010; Larsson et al. 2010, 2011; Laftah et al. 2011; Spagnol et al. 2012a). In the review article by Laftah et al. (2011), the findings of such studies were summarized by stating that nanofibers can either increase or decrease absorbency in different cases. The hydrophilic nature of cellulosic fibers contributes to the absorbency of the structure as a whole, yet the more rigid nature of the resulting network structure can contribute to decreased absorbency. This may explain an observed complexity of the relationship between nanocellulose content and absorbency (Larsson et al. 2011). Anisotropic swelling was observed by Buyanov et al. (2010) in nanocellulose-filled hydrogel structures that had been pre-compressed during their formation. Spagnol et al. (2012a) reported that although nanocellulose inclusion at relatively high levels decreased absorbent capacity, it also resulted in greater pore size. The increase in pore size was judged to be an important advantage in terms of the absorption rate and the overall performance of the absorbents. The cited authors found, however, that at a low content of nanocellulose (5 to 10%) the water content was the highest. This trend was confirmed in subsequent work (Spagnol et al. 2012b,c).

EFFECTS OF FLUID COMPOSITION ON ABSORBENCY

In addition to the chemical and structural aspects of the sorbent material, as discussed in the previous two sections, the absorbency characteristics of cellulose-based materials can also be profoundly affected by the nature of the aqueous fluid to be absorbed. In addition to providing practical information, the goal of this section is to summarize evidence that can be used to support various theoretical concepts, which are addressed in later sections.

Effects of Salinity

Inspection of the last two data columns in Table A reveals a strong contrast between the amounts of pure water taken up by various cellulose-based absorbents vs. the corresponding amounts of saline solution (often evaluated with 0.9% NaCl). In typical cases the latter value is less than half of the absorption capacity in the case of pure water. Especially sharp contrasts in capacity, when comparing deionized vs. saline water, have been reported in the case of superabsorbent hydrogels (Lim et al. 2001; Suo et al. 2007; Zhang et al. 2008b, 2009; Kim and Mun 2009; Liang et al. 2009; Pourjavadi et al. 2008, 2009; Chang et al. 2010; Bao et al. 2011; Liu et al. 2011; Wang et al. 2011a,c,d; Wu et al. 2012).

For certain end-uses it would be highly desirable to maintain a high level of absorbency even in the presence of intermediate levels of salt ions. A number of authors have reported progress in development of somewhat “salt-tolerant” hydrogels (Kulicke and Nottelmann 1989; Yang et al. 2009; Wu et al. 2012). Yang et al. (2009) attributed the observed salt-tolerance to graft polymerization between CMC and polyacrylamide. Rezai and Warner (1997) reported that even though salt greatly decreased the absorbency of acrylonitrile-grafted wood pulp fibers, such fibers still absorbed three times as much water as their ungrafted counterparts. Also, various hydrogels comprised of hemi-cellulose in combination with chitosan have shown favorable performance in their relative ability to absorb saline solutions (Salam et al. 2011b).

Fält and Wågberg (2003) as well as Li and Dai (2004) observed an interesting effect of salt on the swelling of highly purified cellulosic fibers. The samples had been prepared with the complete exclusion of salt ions. Swelling actually tended to increase with the first added amounts of such salts as Na2SO4 and NaCl. The effect was attributed to an increase in effective pH within the mesopore spaces within the fibers. Related mechanisms have been further considered (Hubbe 2006; Hubbe et al. 2007a), as a way to account for effects of salt concentrations on the streaming potential of cellulosic fibers.

Effect of pH

Conventional kraft fibers used in papermaking are known to exhibit increased water retention with increasing pH (Jayme and Büttel 1964; Lindström and Carlsson 1982). Such effects have been observed most clearly in the case of relatively high-yield pulps, i.e. fibers that contain substantial amounts of hemicellulose and other species bearing carboxylic acid groups (Lindström and Carlsson 1982). In comparison, such effects in the case of bleached kraft pulps tend to be small (Lindström and Carlsson 1982; Hubbe and Panczyk 2007a).

In the case of hydrogels having cellulosic content, numerous studies have reported increased swelling with increasing pH, at least up to the neutral point (Vitta et al. 1986; Barbucci et al.2000; Sannino and Nicolais 2005; Chen and Tan 2006; Ali et al. 2008; Wang and Wang 2010; Anirudhan et al. 2011; Bao et al. 2011; Eldin et al. 2011; Wang et al. 2011b,d). Certain studies (Liu et al. 2005a; Ibrahim et al. 2007; Pourjavadi et al. 2007b; Akar et al. 2012; Shi et al. 2013; Wang et al. 2008, 2013) showed a distinct maximum swelling at neutral pH; a possible explanation for this finding is that increasing amounts of ions must be added to such a system to achieve pH values significantly different from 7; such ions would be expected to have the same effect corresponding to the previous subsection (Effects of Salinity). Odd results, roughly opposite to those just cited, were obtained by Varaprasad et al. (2011), though no explanation was given. Wang et al. (2010) developed a system based on hydroxyethylcellulose-g-poly(acrylic acid) with very finely divided attapulgite mineral that achieved consistent high swelling within the broad pH range between 4 and 10.

Evidence of an additional contribution to de-swelling has been observed in some hydrogel systems at the low end of the pH range investigated. Barbucci et al. (2000) observed a strong contraction of hydrogels at pH values well below the pKa value of the constituent carboxyl groups. The effect was attributed to the formation of hydrogen bonded pairs between adjacent protonated carboxylic acid groups.

Shang et al. (2008) developed a creative system in which the bending of a gel body could be used as an indication of the strength and direction of an imposed electric field. According to the authors, the bending of the gel was attributable to slight pH differences on either side of the hydrogel, resulting from the half-reactions at the electrodes used to generate the electric field.

Effects of Valency and Identity of Positive Ions

Numerous studies have shown greater decreases in swelling with increasing valency at a constant concentration of cations in the aqueous solution to which a hydrogel is exposed (Vitta et al. 1986; Kulicke and Nottelmann 1989; Feng and Wang 2010a; Bao et al. 2011; Akar et al. 2013; Chang et al. 2012; Li et al. 2012; Raafat et al. 2012). The effect has been attributed to reversible complexation between the cationic species and one or more carboxylate groups in the hydrogel (Bao et al. 2011; Raafat et al. 2012). In addition, multivalent cationic species can be expected to have a relatively large effect in reducing ionic double layer forces of repulsion associated with the bound carboxylate groups of a typical hydrogel (Hiemenz 1977). As noted by Vitta et al. (1986), by acting as a bridge between more than one carboxylate group in a hydrogel, a multivalent cation can act as a crosslinking agent, thus decreasing swelling.

Though not as obvious or important as the effect of valency, many studies have detected differences in the effects on swelling between cations of the same valency. Two studies dealing with SAPs showed a greater suppression of swelling in the presence of the divalent magnesium ion, in comparison with the same concentration of calcium ion (Bao et al. 2011; Li et al.2012). The same authors showed that monovalent sodium ion generally suppressed swelling to a greater degree than the potassium ion. In the case of acrylic-grafted sulfite fibers, Vitta et al. (1986) reported the following sequence of increasing water retention values at a fixed concentration of the following ions:

Al3+ < Cu2+ < Zn2+ < Ca2+ < Mg2+ < H+ < Na< K+ < Li+

It should be stressed that all of the effects related to different cations of the same valency were small relative to the effect of valency. It is interesting, though, that within the individual studies cited, the results tended to follow regular progressions according to the tabulated size of the respective ionic species.

Surfactants and Absorbency

Surfactants have been shown to affect absorbency in three ways. First, in the case of ordinary paper, addition of various surfactants has been shown to facilitate more rapid wetting in some cases (Hodgson and Berg 1988; Shepherd and Xiao 1999). Such treatments were able to overcome a “wetting delay” phenomenon that is often observed when water is spread onto initially-dry paper. Increased absorbency was observed by Shepherd and Xiao (1999) following treatment of the paper with surfactant.

Second, addition of a cationic surfactant before preparation of paper has been found to yield a bulkier structure (Mann and Leah 1987; Fatehi et al 2010). The debonding effect of such treatment resulted in dry-strength decreases in the range of 14 to 19%. Poffenberger et al. (2000) showed that, due to the hydrophobic nature of typical cationic surfactants, the net effect of debonding treatment during the production of tissue paper may be a reduction in the uptake of water. However, by the careful selection of debonding agents having a relatively hydrophilic character, such decreases can be avoided.

Third, in the case of SAPs, addition of cationic surfactant has been shown to result in a marked decrease of absorbency (Wang and Wang 2010; Wang et al. 2011c). Such an effect is consistent not only with the partially hydrophobic nature of the surfactants employed, but also with the expected electrostatic interactions. Wang and Wang (2010) noted that changes in swelling can be affected by screening of the negative charges within the SAP due to the cationic charge of the surfactants, changes in hydrogen bonding interactions, and a hydrophobic interaction between the alkyl tails of the surfactant molecules. In related work, Gawel et al. (2012) showed that the nature of a conventional anionic hydrogel was transformed upon the addition of a cationic polyelectrolyte. The surfactant treatment caused the hydrogel to have absorption behavior similar to that of a polyelectrolyte complex type of hydrogel, as was discussed earlier. In the case of ordinary kraft fibers, Swerin et al. (1990) showed that addition of cationic polymers caused a loss of swelling. Effects of cationic polymers on the swelling of cellulosic materials and anionic hydrogels will be considered in a later section dealing with theoretic aspects of swelling capacity.

Adsorbed Polyelectrolytes

The term “adsorption” can be defined as a process by which a chemical species (such as a polyelectrolyte) becomes attached to a surface – either on the apparent external surface of a material or on surfaces within a porous structure. A study by Ström and Kunnas (1991) demonstrated that polyelectrolytes adsorbed onto papermaking fibers can make a significant contribution to their water retention value. Cationic polyelectrolytes of high charge density were shown to decrease the amount of water held within the pulp fibers after their centrifugation to remove free water. The effect was attributed to the ability of highly-charged cationic polymers to accelerate dewatering of paper pulp; this was a surprising finding, since water retention value tests are usually intended to provide information about equilibrium conditions. A related study by Hubbe and Panczyk (2007b) confirmed the effect of low- to moderate-mass, high-charge-density cationic polymers on reducing the water retention of kraft pulps. By contrast, sufficiently high levels of very-high-mass cationic polymers tended to increase water retention. The latter effect was attributed to the water associated with the polyelectrolytes themselves in their adsorbed state.

Temperature

Effects of temperature are of both theoretical and practical interest (Miyamoto et al. 1995). Some authors have called thermal changes in absorbency a “smart” feature, since they may allow specialized applications of a thermo-sensitive hydrogel (Pourjavadi et al. 2007b; Yang et al. 2011). Typical temperature-responsive hydrogels exhibit markedly decreased absorbency with increasing temperature (Harsh and Gehrke 1991; Kuwabara and Kubota 1996; Çaykara et al 2006; Chang et al. 2009b). Decreases in absorbency as high as a factor of five (Kuwabara and Kubota 1996) or ten (Çaykara et al 2006) have been reported when comparing absorbency at about 20 oC to the correspond-ding value at about 40 oC. The strongest thermal effects are for SAPs incorporating relatively high proportions of poly(N-isopropylacrylamide); these generally exhibit a strong decrease in absorbency at temperatures near to physiological conditions (Kuwabara and Kubota 1996; Liu et al. 2004; Chang et al. 2009b; Yang et al. 2011). For example, Kuwabara and Kubota (1996) used a two-step procedure to graft a 51/49% acrylic acid/N-isopropylacyrlamide onto CMC at a grafting level of 205%; water absor-bency fell from about 14 g/g at 10 oC to 3 g/g at 50 oC. Çaykara et al. (2006) prepared semi-interpenetrating networks of poly-[(N-tert-butylacrylamide)-co-acrylamide] with hydroxypropylcellulose; water sorption was observed to fall from as high as 19 g/g at 5 oC down to about 2 g/g at 40 oC and above. Chang et al. (2009b) reported a decline in absorbency from 18 g/g to 5 g/g over a temperature rise from 10 oC to 60 oC in the case of an interpenetrating network of one part cellulose to three parts of poly-NiPAM. Likewise, in the case of SAPs comprising carboxymethylchitosan and poly-NiPAM, Ma et al. (2007) observed a decrease in buffer solution absorption from 19 g/g at 25 oC to 2 g/g at 35 oC. An opposite thermal effect, with moderately increased swelling of hydrogels with increasing temperature, has been reported less frequently (Liu et al. 2005a; El-Din et al. 2010). Hirsh and Spontak (2002) showed that the thermo-responsive characteristics can be attributed to changes in the elastic properties of such hydrogels.

External Pressure and Absorbency

In various applications of absorbent materials it may be important to resist the release of water upon compression. In fact, one of the strategies in the design of highly absorbent products is to employ a relatively stiff fiber matrix such that compressive forces acting on SAPs in the product are minimized (Buchholz et al. 2005). According to the same source, as has been noted, the presence of fibers also can restrict swelling. Guilherme et al. (2005) evaluated the stress-strain behavior of hydrogels as a function of temperature. It was shown that higher temperatures led to increased stress values in the cases considered. This was associated with the absorption of water, i.e. osmotic pressure. Sannino et al. (2004b) described the development of a test device capable of comparing the effects of compression on the absorption capacity of SAPs. Based on the results shown, it was clear that even a moderate applied pressure of about 300 kPa was sufficient to reduce the swelling with water by 90% in the case of CMC/HEC hydrogels that had been equilibrated (at different levels of compression) for 24 hours.

ACCOUNTING FOR OBSERVED SORPTION CAPACITIES

As has been shown in previous sections, the amount of aqueous fluid taken up by cellulose-based absorbents can depend strongly on the chemical composition and structure of the absorbent, and also on the characteristics of the liquid. Theoretical explanations that have been proposed to account for absorption capacities will be considered here. Some notable review articles that emphasize theoretical issues relevant to the extent of swelling are available (Scallan 1983; Brannon-Peppas and Peppas 1990; Kazanskii and Dubrovskii 1992; Ganji et al.2010).

Quantification of Sorption Capacities

As a first step in reviewing theories to account for absorption capacities and rates, it is important to consider how absorbency is tested. Some notable advances in measure-ment procedures have been described, as listed in Table 2.

Table 2. Methods for Quantification of Absorption Capacity or Rate

When testing the water sorption capacity of cellulosic fibers, Welo et al. (1952) observed that a simple weight-loss method, with free evaporation under ambient conditions, was easy and repeatable. The materials were first fully swollen with an excess of water, then weighed periodically. It was found, for each specimen, that weight loss was linear with temperature up to a critical point, and thereafter slowed down. The critical point was identified with the loss of all external water, leaving behind only water of swelling. The water at that critical point was used as an approximation of the water-holding capacity of the fibers. More recently, a similar concept was developed using isothermal drying in a thermogravimetric analyzer; the non-external water was named hard-to-remove water (Park et al. 2006).

The Donnan Equilibrium and Osmotic Pressure

When an absorbent material takes up a liquid, either that liquid needs to displace another fluid (such as air), or the material needs to swell to accommodate the new liquid. This section will consider osmotic pressure and various studies showing how osmotic pressure can account for such swelling, especially when a cellulosic material or a hydrogel is placed in water having few or no salt ions (Grignon and Scallan 1980; Vitta et al. 1986; Oppermann 1995; Sannino et al. 2003; Buchholz et al. 2005; Ganji et al. 2010). The same principles also apply to cellulosic fibers, though usually to a lesser extent. Part of the justification for considering cellulosic materials in the context of hydrogels involves the presence of ionizable bound groups, mainly carboxyl groups, within their structure (Herrington and Petzold 1992). Such groups contribute to swelling (Grignon and Scallan 1980; Lindström and Carlsson 1982) and ion exchange phenomena (Rudie et al. 2006; Hubbe et al. 2011) within cellulosic fibers.

One of the most important principles underlying swelling is that, under equilibrium conditions, there will be a fixed relationship between the concentrations of ions inside and outside of a swellable, gel-like material (Grignon and Scallan 1980). In other words a Donnan equilibrium condition will become established between the ion concentrations inside and outside of the hydrogel. If one assumes, for instance, that an aqueous system containing a suspended anionic hydrogel phase contains the ions OH, H3O+, Cl, and Na+, then the Donnan equilibrium can be expressed in the form,

(1)

where the subscript “g” indicates location within the gel phase and the subscript “s” indicates location within the solution phase, respectively.

A characteristic feature of a typical hydrogel is that ionizable groups are bound to polymer chains within the gel phase. The first analysis of the relationship between such charged groups and swelling was based on ordinary gelatin and the effects of acidity in a surrounding bulk solution (Proctor 1914). Grignon and Scallan (1980) were the first to extend the same concept to the swelling of pulp fibers; they noted that the amorphous polysaccharide regions within pulp fibers can be regarded as being hydrogels. In either case, the tendency of the solid to swell upon placement in deionized water can be visualized as a process whereby water rushes into the hydrogel zones in order to make the value of become closer to one. As water progressively enters the gel phase, the ionic concentration associated with the bound ionizable groups becomes diluted; however, in salt-free systems, such dilution is never enough to eliminate the contrast in ionic strength between the inside and the outside of hydrogel zones.

Once equilibrium has been established, the osmotic pressure tending to swell the hydrogel phases within the mixture can be calculated based on the following equation (Ganji et al. 2010),

 (2)

where is the osmotic coefficient of the gel phase and is the corresponding quantity for the external aqueous solution. Likewise, is the concentration of the ith species in the gel phase and is the corresponding value in the external solution phase. At equilibrium it is understood that the osmotic pressure will be in balance with elastic forces tending to limit swelling in the material (Grignon and Scallan 1980). The following equation can be used to estimate the osmotic coefficient of the gel phase in Eq. 2 (Ganji et al. 2010),

 (3)

where np is the molarity of the polymer, nm is the molarity of the monomer, and is the degree of ionization as a fraction.

Entropy of Mixing

A second key factor accounting for the swelling of a hydrogel with water involves the entropy of mixing (Sannino et al. 2003; Ganji et al. 2010). Such effects can be understood based on Eq. 4, for which a negative value of the change in Gibbs Free Energy (G) of a system indicates that a process will occur spontaneously:

G = H – T S (4)

In this expression, H is the change in heat energy (enthalpy), T is the absolute temperature, and S is the change in entropy. The term S is sometimes called a measure of the disorderliness of a system. In principle, energy is required to separate a mixture into pure subphases; by contrast, miscible materials will spontaneously form a single, disorganized phase with the release of energy when placed together.

Entropy can play a significant role in the swelling of hydrogels because a system in which the gel material is infused with water has greater randomness (higher entropy) than the separated (dry) hydrogel and water phases. The contribution to swelling pressure due to the entropy of mixing can be estimated from Eq. 5 (Ganji et al. 2010),

 (5)

where is the polymer volume fraction, V1 is the molar volume of the solvent, and is the interaction parameter.

Figure 3 illustrates another way in which changes in entropy tend to contribute to the swelling of a typical hydrogel material when it is placed in water (Khokholv and Nyrkova 1992). Here it is assumed that the gel material contains ionizable acid-type groups, such as –COO or –SO4etc. The left-hand frame of the figure represents the gel material in its dry form as the sodium salt of the poly-anionic gel. In such form, the component ions have very limited freedom of motion, being bound by their ionic bonds and the contracted, solidified nature of the gel. As illustrated by the right-hand frame of the figure, the addition of water, along with the swelling of the gel, implies more degrees of freedom. For instance, sodium ions may effectively “switch places” as counter-ions for the bound anionic groups. As indicated in the upper right of the figure, such an exchange even can occur with salt ions from the bulk. Further contributions to disorder and freedom of motion within such a system are associated with chain mobility and the relatively long (in atomic terms) average distance between oppositely charged ions in an aqueous system having very low electrical conductivity (Hiemenz 1977).

Fig. 3. Illustration of increased disorder and freedom of motion of a counter-ion (e.g. sodium) in a typical hydrogel that contains acid groups (e.g. carboxylate). Left: Dry form as sodium salt of polycarboxylate hydrogel; Right: Water-swollen form, in which the counter-ions are free to exchange with each other, to diffuse within the field of electrostatic forces, and also for the polyelectrolyte chain segments to undergo some constrained motion.

Hydration and Hydrogen Bonding Effects

A third key element to account for the swelling of hydrogels, in addition to the osmotic effects and entropy contributions just described, is hydrogen bonding. The relatively strong interactions between water molecules and hydrogel materials promote hydration and ionization, such that the osmotic and entropic effects can operate. The importance of hydrogen bonding relative to hydrogel swelling was perhaps best demonstrated by Koroskenyi and McCarthy (2001), who studied the highly hydrophilic polymer konjac glucomannan. In its native state the gelatinous material absorbed over 105 times its mass of water. By contrast, after full acetylation, removing all of the –OH groups, the same material had a water absorbency of only 1 g/g. Another clue that hydrogen bonding is intimately related to gel behavior comes from infra-red analysis. Sammon et al. (2006) detected distinct shifts in the wavelengths of infrared absorbance maxima that were coincident with gelation of hydroxylpropylmethylcellulose (HPMC) as a function of temperature.

Regarding the state of water within a typical hydrogel, various views have been expressed in the literature. According to Karaaslan et al. (2012), most of the water within a swollen hydrogel should be regarded as bound water. In other words, its properties can be expected to differ from those of bulk water in terms of such behaviors as freezing tendencies and temperatures. Accordingly, Faroongwarng and Sukonrat (2008) observed that the freezing point of water was depressed within hydrogels. Gelin et al. (2007) estimated that only about 10% of water present in a typical hydrogel acts like bulk water. Their work with measurement of ion diffusion rates, in addition to freeze-fracturing of hydrogels, led to the conclusion that the bulk-like water within hydrogels tended to be confined to isolated “lakes” within the pore structure. This finding suggests that various reported porous hydrogels or “foams” might be expected to contain substantial bulk-like water in addition to chemically bound water (Kabra et al. 1998; Omidian et al. 2005; Lu et al. 2007; Chang et al. 2009a,b; Sehaqui et al. 2010; Salam et al. 2011a). Joshi et al. (2008) proposed that the relatively sharp gelation transition that takes place upon heating of methylcellulose solutions can be attributed to a thermal breakdown of the water structure, allowing coalescence of the polymer chains. Venditti and coworkers have demonstrated that freezing and non-freezing bound water and bulk water are present in cellulosic fibers and can be measured with differential scanning calorimetry (Park et al. 2007). The heat of vaporization of the bound water can be almost 30% higher than bulk water. For fibers that cannot participate in hydrogen bonding, such as polypropylene, bound water and increased heat of vaporization of waters do not exist (Park et al. 2007).

Progress in understanding the nature of water within cellulosic materials has been achieved through advances in spectroscopic methods. Froix and Nelson (1975), using nuclear magnetic resonance relaxation times, concluded that free water, in addition to chemically bound water, existed within cellulosic fibers even in “dry” paper in equilibrium with relative humidity values above 0.1. Using NMR and ESR methods, Capitani et al. (1990) came to the conclusion that water domains within cellulose somehow permeate cellulose domains, such that the two kinds of domains interact by hydrogen bonding. Using calorimetric methods, Maloney et al. (1998) found evidence that bulk-like water was present within the larger mesopores of never-dried kraft fibers after their placement in water; by contrast, only bound water was present within the cell walls of previously dried cotton. Park et al. (2006) showed that such considerations can be used to estimate the dimensions of nano-sized water-filled pores within cellulosic materials.

Elastic Modulus

As proposed by Proctor (1914), the extent of swelling of a typical hydrogel in pure water or various electrolyte conditions can be modeled in terms of a balance between osmotic forces and restraining elastic forces due to the swollen material. Scallan and coworkers extended these concepts to cellulosic fibers (Grignon and Scallan 1980; Scallan and Tigerström 1992). In either case, pure water can be expected to diffuse into a hydrogel until a combination of partial dilution of the effective ionic concentration within the gel, plus elastic stretching of the gel achieves a balance of forces (Vitta et al. 1986; Ganji et al. 2010). The viscoelastic properties of various hydrogels, sometimes as a function of their state of swelling, have been studied (Guilherme et al. 2005; Feng et al. 2006; Chang et al. 2009a; Dai and Kadla 2009; Larsson et al. 2010). Notably, Larsson et al. (2011) and Spagnol et al. (2012a) showed that the inclusion of microfibrillated cellulose or cellulose nanocrystals within hydrogels had a strong reinforcing effect, essentially increasing the modulus and limiting the amount of swelling. Yoshinobu et al. (1992) and Lionetto et al. (2003) showed that there was a relationship between the degree of crosslinking and the elastic modulus of various hydrogels. As already noted, Oppermann (1995) proposed that inclusion of cellulose derivatives within a hydrogel can result in stiffening of the gel. Such an effect follows from the inherent stiffness of cellulose, which may be related to local association of cellulose segments.

An elastic contribution to pressure within a hydrogel, resulting from the configurational entropy of polymer chains, can be estimated if one makes various assumptions concerning the nature of the hydrogel. The stretching of polymer segments associated with the swelling of the gel implies a restriction in the number of motions of individual groups and segments along the chain, i.e. a loss of configurational entropy when the hydrogel material is stretched. The lead terms in an equation to estimate the resulting contribution to pressure (tending to resist expansion of the gel) are as follows (Ganji et al. 2010),

 (6)

where is the concentration of the polymer chains at the onset of gel formation.

Network Volume Constraints

Cross-linking effects

As was noted earlier, the highest recorded values of swelling tend to be associated with the lowest practical (and sometimes impractical) extents of crosslinking (Bin et al. 2000; Wach et al. 2001, 2002; Liu et al. 2005a; Demitri et al. 2008; Han et al. 2010; Pourjavadi et al. 2010; Kono and Fujita 2012). The basis for this relationship is illustrated in Fig. 4. As discussed by Ganji et al. (2010), it is possible to estimate the maximum extent of swelling of a hydrogel based on the density of crosslinking, and more specifically the lengths of the characteristic polymer segments between adjacent crosslink points. The quantity , represented by a typical segment in Fig. 4, is defined as the average or typical molecular mass between adjacent crosslink points. The quantity represents the linear correlation distance between the crosslinks. The volume fraction of the polymer in a swollen gel is then given by the following expression,

v2,s Vp /Vg = (1/2) / [ (Qm /1) + (1/2) ] (7)

where Vp is the polymer volume within the material, Vg is the volume of the swollen gel phase as a whole, 2 is the density of the polymer, 1 is the density of the solution, and Qm is the swollen volume on a mass basis. The degree of crosslinking, X, can be defined as

 (8)

where Mo is the characteristic molecular mass of the initial polymer chains before crosslinking. In cases where the polymer-solvent interactions have been tabulated, the typical dimensions of spaces between crosslink points can be estimated from (Ganji et al. 2010),

 (9)

where l is the bond length between carbon atoms, and Cn is the tabulated ratio for the system, as derived by Flory (1953). It should be noted that while the cited work did not take electrostatic effects into account, such interactions can be expected to profoundly affect polymer conformations (Fleer et al. 1993) and hence swelling.

Fig. 4. Representation of a hydrogel in its somewhat condensed (left) and highly swollen (right) states. Solid dots indicate cross-linked points of the polyelectrolyte chains (shown as coiled or wavy lines). The characteristic molecular weight of segments between crosslinks () and the correlation distance between crosslink points () are illustrated by considering one segment.

Cellulosic Fibers and Hydrogel Swelling

Various authors have suggested that a network of cellulosic fibers, i.e. a paper-like structure or a loose network of nanofibrillated pulp or fluff pulp fibers filled with hydrogel particles, may play a role in maintaining the structural integrity in a combined absorbent structure that contains SAPs (Buchholz et al. 2005; Brodin et al. 2012). It follows that such fibers may also restrict the swelling of the SAP material upon introduction of aqueous fluid. The concept is illustrated schematically in Fig. 5.

Fig. 5. Illustration of the concept that a network of macroscopic cellulosic fibers in an absorbent product might tend to constrain the swelling of hydrogel particles. Left: Representation of dry condition; Right:Swollen condition

As was noted earlier, when nanocellulose elements are incorporated into hydrogels, the observed result is generally a marked decrease in absorbency, often accompanied by an increase in the mechanical strength of the gel material (Dai et al. 2009; Larsson et al. 2010, 2011; Spagnol et al. 2012a). In an analogous manner to the mechanism just described to account for the restraining effects of fluff pulp fibers on SAP swelling, some authors have attributed the effects of nanocellulose on SAP absorbency to a “reinforcement” effect (Larsson et al. 2011). Indeed, when cellulosic nanofibrils or crystallites have been compounded with various plastics, the resulting “nanocomposites” are typically much stiffer than the matrix plastic material by itself (Azizi Samir et al. 2005; Hubbe et al. 2008). However, before placing too much weight on any such analogy, it is important to emphasize the huge size difference between a softwood kraft fiber and the smallest classes of cellulosic nanomaterials. Figure 6 provides a schematic comparison. The lengths of the nanocellulose entities, in some cases, may be as much as 10,000 times smaller than a typical fluff pulp fiber. As shown, one of the likely roles of the nanocellulose may be to function as cross-linking points. It is assumed here that polymer segments within the hydrogel may associate locally with the cellulosic surfaces. Such behavior would be consistent with observations that the effects of adding a crosslinking agent or adding nanocellulose are sometimes nearly equivalent (Dai and Kadla 2009; Larsson et al. 2010).

Fig. 6. Illustration emphasizing the great size difference between fluff pulp fibers vs. cellulosic nanocrystal, a difference that suggests that their roles with respect to reinforcement, crosslinking, or other effects might also be different

Drying and Absorbency

A profound loss of swelling capacity is known to occur when delignified pulp fibers are dried (Stone and Scallan 1966; Weise and Paulapuro 1999). Loss of fiber absorbency and the relative effects of loss of conformability of the wetted fibers, especially when fibers are dried multiple times, is a major contribution to reduced paper strength when the material is recycled (Weise and Paulapuro 1996; Hubbe et al. 2007b).

A unique example of loss of absorbency upon drying was documented by Mizutani et al. (1999), who studied the water absorbency of never-dried cotton fibers starting very soon after their biosynthesis in the bowl of the living plant. At the earliest measurable stage, the never-dried cotton fibers were associated with 2.5 times their mass in water. Upon drying of such fibers, they were then only able to absorb water to an extent of about 0.8 times their mass. The effect was attributed to the fact that in the continuously wet condition, while sealed within the fresh cotton bowl, the water present between the adjacent cellulosic microfibrils in the young cotton prevented those fibrils from coming together and hydrogen bonding directly with each other. As shown by Stone and Scallan (1966, 1968), the loss of absorbency upon drying of cellulose fibers generally can be attributed to a partly irreversible closure of submicroscopic pores. Strong capillary forces acting during the drying process (Campbell 1959) can pull the walls of some pores into close contact, and the regularity of the cellulose structure can result in a zipper-like formation of multiple, repeating hydrogen bonds, essentially causing the adjacent surfaces to coalesce into a single cellulosic entity (Hult et al. 2001; Pönni et al. 2012). Though additional refining has been shown to restore much of the bonding strength that is lost when kraft fibers are dried, the corresponding effects related to restoration of porosity tend to be more limited (Klungness and Caulfield 1982).

Efforts to suppress the loss of absorbency (and other valuable traits such as bonding potential) when kraft fibers are dried have been limited in success (Zhang et al. 2002). Early work by Higgins and McKenzie (1963) showed only one kind of treatment – the drying of cellulose fibers in the presence of high concentrations of sugar – that was effective in avoiding the irreversible loss of swelling potential. Work by Lindström and Carlsson (1982) showed that some of the damage could be avoided by drying under neutral to alkaline pH conditions – but only if the pulp also had a relatively high level of carboxylate groups, i.e. a high hemicellulose content. Subsequent work has confirmed the favorable effects of hemicellulose (Cao et al.1998), which may be attributed not only to its content of carboxyl groups, but also to its somewhat irregular chain structure that tends to resist crystallization. The importance of hemicellulose for swelling of birch kraft pulps was shown dramatically by Moss and Pere (2006). These researchers showed that intentional removal of most of the xylan component from the fibers almost eliminated the ability of such fibers to swell when they were placed back into water.

To avoid the usual deswelling effect of drying, freeze-drying is one strategy that has been successfully employed. Such an approach has been used to prepare cellulose-based “aerogels” that maintain very high porosity, bulk, and absorption capacities (Saito et al. 1995; Pääkkö et al. 2008; Kettunen et al. 2011; Theliander et al. 2011; Brodin et al. 2012). Another successful approach has been to transfer the swollen cellulosic material to an acetone solution before it is dried in air. Sannino et al. (2010) used such an approach to preserve the swollen nature of nanoporous crosslinked cellulose gels when they were dried.

Polyelectrolyte Complexes as a Smart Restraint on Hydrogel Swelling

In view of recent success in achieving highly favorable levels of absorption of saline solution through the use of polyelectrolyte complex hydrogels (Salam et al. 2011b), coupled with the tendency of such complexation to restrain the absorption of pure water (Dogu et al. 2009), this section will consider theoretical aspects of such absorbency. Wood-derived polysaccharides and their derivatives such as xylans and carboxymethylcellulose (CMC) can be readily incorporated into such systems, in combination with various cationic polymers such as chitosan (Salam et al. 2011a,b; Ayoub et al. 2013).

The tendency of polyelectrolyte complexes (PECs) to contain relatively little water within their core structure, under conditions where the amounts of oppositely charged bound ionizable groups are in approximate balance, can be understood based on the mechanism shown in Fig. 7. As shown in the figure, there are bound anionic groups on one set of polyelectrolyte chains and bound cationic groups on the other set. In each case the bound charges are balanced by an equal and opposite number of counter-ions. (Note that this is in addition to the various other monomeric ions of each charge in the system. Such ions have to sum up to equal normality of each charge). Upon mixing the two polyelectrolyte solutions, a complex is formed (Michaels 1965). As shown in the figure, the cooperative interactions involved in multiple ionic interactions between the oppositely charged polyelectrolytes tends to drive the complexation to completion. It has been proposed that complexation might occur in a zipper-like fashion or in a disorganized manner (Michaels 1965). In either case, the resulting ion pairs within the PEC can remain stable, such that the charges within the system are essentially self-neutralized. As illustrated in the figure, one of the results, therefore, is that the counter-ions can diffuse away into the bulk of solution. Some authors have credited the resulting increase in entropy as being the most important factor in promoting PEC formation (Michaels et al. 1965; Cundall et al.1979; Argüelles-Monal et al. 1990; Baker et al. 1995).

Fig. 7. Mechanism of polyelectrolyte complex (PEC) formation in which ion pairing between bound groups on the respected polyacid and polybase chains make it possible for the monomeric counter-ions to diffuse away, with a large increase in entropy of the system.

Antipolyelectrolyte effect

The tendency of polyelectrolyte complexes and polyampholyte systems to swell with water upon addition of salt has been called the “antipolyelectrolyte effect” (Valencia and Pierola 2007; Dogu et al. 2009; Wang et al. 2010). Because this effect appears to be consistent with the behaviour of the hemicellulose-based PECs just described, some background will be given here. As an underlying explanation for the effect, it has been shown that the addition of salt tends to weaken the electrostatic interactions within a PEC (Dragan and Cristea 2001). Increased absorbency with increasing salinity of aqueous solutions has been observed in several cases in which oppositely charged polyelectrolytes were employed in the preparation of cellulose- or hemicellulose-based hydrogels (Paula et al. 2002; Feng and Pelton 2007; Valencia and Pierola 2007; Wang et al. 2010; Shang et al. 2008; Salam et al. 2011a,b; Gawel et al.2012). Such weakening of the pair-wise associations of ionic groups on the polymer chains allows a partial reversal of the process illustrated in Fig. 7, thus drawing more nomomeric ions, along with some associated water, into the interior of the hydrogel phase.

Fascinating behavior related to the sorption of monovalent salt ions was observed in the course of recent studies with PEC-type hydrogels formed from biopolymers having relatively high charge densities (Ayoub et al. 2013). The ability of such hydrogels to remove sodium and chloride ions from solution, beyond what would have been predicted from the concentration of those ions in solution, was attributed to the process described in Eq. 10,

RCOO+NH4R’ + NaCl RCOO Na+ + R’NH4+ Cl (10)

where the dashed line is used to indicate ion-pair interactions within the hydrogel before introduction of the saline solution. The “R” and “ R’ ” indicate the two respective polyelectrolyte chains. As shown in Eq. 10, addition of NaCl to the material is expected to result in weakening of the complex such that the salt ions behave as counter-ions to the charged polyelectrolyte chains. The effectiveness of the PEC-type hydrogels to enable the removal of NaCl from solution is notable because most reported applications of cellulosic-based materials for ion exchange have involved removal of divalent metal ions (Rudie et al. 2006; Hubbe et al. 2011; Kumar et al. 2011). In such applications it is typical for the divalent ions to displace monovalent ions within the substrate due to their greater binding constants. The process represented by Eq. 10 is unique due to the fact that PEC-based hydrogels can be readily obtained in almost monomer-free condition; upon precipitation of the PEC during its formation, it is possible to decant the salt-containing solution, optionally replacing it with deionized water. Subsequently, upon introduction of a sufficiently concentrated solution of monovalent salt, the PEC interactions becomes weakened to such an extent that Eq. 10 becomes favorable.

ACCOUNTING FOR OBSERVED SORPTION RATES

In addition to the absorption capacity, another main concern of users of absorbent products is the rate of absorption. Several reviews dealing with absorption rates have appeared. In the case of ordinary paper products (not superabsorbent), several key studies have focused on theoretical issues (Bristow 1967; Aspler et al. 1987; Schuchardt and Berg 1991). Related numerical simulation work has been reported (Masoodi et al. 2012). The following articles have considered rates of absorption in the case of hydrogels (Kabra and Gehrke 1994; Zohariaan-Mehr et al. 2010; Wang et al. 2011c). Finally, Berg and coauthors have considered the rates of wicking into structures that are contained by fibers and hydrogel materials (Schuchardt and Berg 1991; Wiryana and Berg 1991).

Tests to Measure Absorption Rates

Practical tests of rates of absorbency of cellulose-based materials often involve rapid immersion of material packed into a cage, followed by evaluation of weight gain (Aberson 1969). The cited author noted that the results of such tests can be difficult to interpret. To overcome this problem, a new test was developed in which dry fluff pulp is exposed only on one side (Aberson 1969). As was noted in Table 2 (in previous section), rates of sorption can also be studied by dielectric spectroscopy (Gelin et al. 2007), wicking rates (Beuther et al. 2010), and the time for a water drop to be absorbed (Ström and Carlsson 1992).

For ordinary paper and paperboard products, various TAPPI methods are available (TAPPI 1996). For instance, TAPPI Method T411 describes the gravimetric Cobb test in which the sample is exposed to the liquid from one side for a set period (e.g. 120 s), after which the liquid is poured off and a blotter is used to remove excess liquid. TAPPI Method 491 is a related test in which the whole sample is submerged in the test liquid for a specified period (e.g. 10 min), and the sample is blotted with a standard roller before being reweighed. TAPPI Method T432 is intended for highly absorbent paper products such as blotter paper, tissue, and towel. A specified volume of liquid is gently placed on the sample, and one measures the amount of time that passes before the disappearance of specular reflection from the smooth surface of the liquid. TAPPI Method 561 is a dynamic test in which one continuously monitors the mass of liquid remaining in a reservoir attached to a siphon tube in contact with the paper specimen. Related tests have been adopted for evaluation of textiles (ASTM 1983). For instance, test TS-017 is based on the time in second it takes distilled water to climb either 20 mm or 30 mm up a vertical strip of cloth that has been dipped into a beaker. Test TS-018 measures the time for complete sorption and disappearance of a drop of water having a specified size.

For evaluation of events during very short exposure times of paper samples and aqueous solutions, two main procedures have been developed. According to a test developed by Bristow (1967), a narrow strip of paper is mounted on a slowly rotating wheel. A “headbox” with a precisely measured volume of colored fluid and a narrow slit opening at its base is lowered until it touches the sample. The length of the colored streak transferred to the paper can be used to calculate the volumetric uptake per unit of contact time. Related results have been obtained more recently with a device that measures the speed of transmission of ultrasonic vibrations through a water bath as a function of the time after sudden placement of a sheet of paper perpendicular to those waves (Pan et al. 1988; Leks-Stepien and Khadzhynova 2005).

Viscous Resistance Limiting Rates of Flow through Porous Material

A key factor that tends to reduce the rate with which an aqueous fluid passes into cellulosic-based materials, including hydrogels, is the fluid’s viscosity. The magnitude of viscous resistance also depends on the sizes and lengths of the passages through which the fluid needs to flow. These relationships have been captured in Darcy’s law and its various refinements by subsequent investigators (Darcy 1856; Masoodi and Pillai 2010). In its simple form, Darcy’s law can be expressed as in Eq. 11,

 (11)

where Q is the flow rate through a porous packed bed of uniform cross-section, k is a permeability constant, A is the cross-sectional area of the bed, Pb is the outlet pressure, Pa is the inlet pressure, is the fluid’s dynamic viscosity, and L is the linear distance from the inlet to the outlet.

When applying Darcy’s law to predict the rate of wicking in paper, Masoodi and Pillai (2010) noted that the model tended to work best if it was assumed that the material swelled in the course of wetting in order to fully make room for the liquid. The permeability of initially dry paper appears to vary as a function of time during the wicking of aqueous fluid.

Various approaches can be used to estimate the value of the coefficient k in the case of paper-like structures (Kozeny 1927; Carman 1938, 1939; Carrier 2002; Chapuis and Auberton 2003; Hubbe and Heitmann 2007). For instance, a simple laminar hydrodynamic model was used to derive the following expression,

 (12)

where k is the permeability (distance of penetration per unit of time), is a unit mass of the fluid, is the fluid’s dynamic viscosity, CK-C is the Kozeny-Carman coefficient (usually taken to be about 5), So is the specific surface area per unit displacement volume of particulate material, and e is the fractional void volume (Carrier 2002; Chapuis and Auberton 2003). Ingmanson (1952, 1953) applied similar concepts to the case of cellulosic fibers, taking into consideration the bulky structures formed by fibers, but also considering the compressibility of cellulose. In general, all such models imply that a given applied pressure is likely to result in greater flow with increasing pore size and/or decreasing effective surface area of the packing material. This general prediction has been confirmed, for instance, by studies in which the effective pore size within paper structures was increased by forming the paper in the presence of surfactants (Shepherd and Xiao 1999).

Another approach that has been used to model the kinetics of swelling of hydrogels is based on a visco-elastic model (Omidian et al. 1998). Variation of some parameters that can improve the superabsorbent properties can lead to considerable differences in swelling kinetics, which can be fitted to a Voigt-cell (spring and dashpot) model, as

Stp(1 – e-1/r) (13)

where St is the amount of swelling (g/g) at any moment, p is the power parameter (g/g), t is time (s), and r is the rate parameter. Swelling at different times can be measured in distilled water. Then, the results obtained at equilibrium and the swelling rate of the individual samples can be fitted into the above equation to find the values of the rate and power parameters. Another useful parameter, based on the same model, is the time required to reach 63% of full capacity (Omidian et al. 1998).

Capillarity and the Lucas-Washburn Equation

Capillary forces, due to the surface tension of water, are known to play a major role in the absorption of water and other fluids into porous materials (Buchholz et al. 2005). Lucas (1918) and Washburn (1921) showed that capillary forces can account for the rise of liquids within straws and related porous systems. It has been suggested that the lumens of cellulosic fibers can play a similar role in the wicking of water (Mizutani et al. 1999). Capillary forces also have been credited with explaining the rapid absorption of aqueous fluids into some highly porous hydrogels (Omidian et al. 2005). Sannino et al. (2003) proposed that capillary forces also can increase the amount of water held within swollen hydrogels. In addition, it is reasonable to expect also that capillary forces are involved in cases where fibers having irregular cross-sections are designed to provide channels for wicking within SAP media (Schuchardt and Berg 1991; Pramanick and Crouse 1998).

Figure 8 illustrates some essential features of the Lucas-Washburn model of capillary absorption. As shown, the capillaries are represented as smooth, cylindrical pores of a selected size. In the simplest case one assumes that the material is comprised of open-cell pores, such that the pressure of “Air” shown in the figure is always equal to the ambient pressure of the bulk liquid. As indicated by Eq. 14, the capillary pressure exerted by the interfacial tension at the meniscus is balanced by a viscous force due to the flow within tight passages as water continues to progress into the pore.

The integrated form of the Lucas-Washburn equation, Eq. 15, indicates that the distance of penetration into the porous material can be expected to be proportional to the square-root of time. Another implication of Eq. 14 is that the rate of absorption is expected to increase with increasing pore size. This prediction is reflected in the more rapid absorption of water into coarse, bulky paper (Aberson 1969).

Fig. 8. Idealized representation of porous material having uniform, smooth, cylindrical pores, with a liquid being applied to the outer surface at time equal to zero

Information about contact angles and pore sizes is needed in order to apply the Lucas-Washburn equation to practical systems. The contact angle of water on ordinary kraft pulp fibers has been measured as zero to 52 degrees (Young 1976) or about 56 to 64 degrees (Aberson 1969) in different cases. Since these values are less than 90 degrees, such fibers clearly fall within the category of “wettable” (Adamson and Gast 1997).

The Lucas-Washburn equation has been widely used to fit data related to the absorption of water and other fluids into paper. Two major deviations from the Lucas-Washburn equation were revealed by Bristow (1967), who developed a test method that has since received widespread use by the paper research community (for instance, see Aspler et al. 1987). Evidence of each of these deviations can be seen in Fig. 9, which is a replotting of data from Bristow’s 1967 publication. The horizontal axis in Fig. 9 is plotted on a “square root of time” axis, matching the form of the Lucas-Washburn equation (Eq. 15). Looking first at the original data (square symbols), it is apparent that about 11 mL/m2 of water transferred to the paper surface immediately, essentially independent of time. Bristow attributed that feature of the data to the filling of roughness at the paper surface. The other deviation involved an apparent “time delay”. For instance, in Fig. 9 a linear rise in transferred liquid vs. the square-root of time can be seen following a delay of about 0.2 seconds. Schuchardt and Berg (1991) observed a corresponding wetting delay of about 0.7 seconds in the case of pads made of fluff-grade softwood kraft fiber.

A question that apparently never has been addressed in the literature is whether the original data in Fig. 9 could be made to fit to the Lucas-Washburn equation by subtracting the delay time. To test that idea, Bristow’s data has been recalculated after subtracting 0.2 seconds from each of the reported time durations. The diamond-shaped symbols correspond to the adjusted time quantity before calculating the square-roots and replotting the data. As shown, the data remained in reasonable agreement with a straight line. Moreover, the left-hand end of that line corresponding to adjusted zero time was close to the level of the amount of transfer attributed by Bristow (1967) to the volume capacity of the roughness at the paper surface.

Fig. 9. Data of Bristow (1967) as originally reported and also replotted with a time delay of 0.2 seconds incorporated before the calculation of the square root of time

To explain the existence of a time delay, it has been suggested that humidity adjacent to an advancing droplet of water causes hydration of the cellulosic material and/or reorientation of various chemical groups associated with the cellulosic fibers, extractives, or additives used in papermaking (Lyne and Aspler 1982). Such changes are expected to result in progressive lowering of the contact angle such that the meniscus is then able to advance. It is likely that such effects are compounded by the highly porous nature of paper materials, involving multiple size scale of porosity; therefore it is expected that the effects due to somewhat unfavorable initial contact angles are amplified by the inherent difficulty of spreading a liquid onto a surface that is covered with fine pores (Cassie and Baxter 1944).

Eklund and Salminen (1987) extended Bristow’s model to take into account externally applied pressure of the wetting fluid on the rate of absorption into paper. Though this approach was used in an effort to model such processes as the coating of paper, the same principles could well apply to various situations in which absorbent products are used. The cited authors also took into account such effects as air pressure within dead-end pores, vapor-phase transport of moisture, and the swelling for fibers.

Schuchart and Berg (1991) found that the model of Lucas and Washburn fit very well when applied to the absorption of non-swelling liquids such as octane. In the case of water, as mentioned earlier, the permeability of absorbents tended to decrease due to the swelling of superabsorbent hydrogel particles, thus partially blocking flow passages in the material.

Numerical simulation of capillary-induced absorption into paper-like media was carried out by Masoodi et al. (2012). Their model included the swelling of the medium as well as the time-varying permeability.

Concepts related to capillarity can be used as a basis for the design of absorbent products. For instance, Aberson (1969) proposed that the pore spaces within a well-designed sorbent material ought to be no larger than the size that will provide enough suction to oppose gravity and hold onto all of the absorbed liquid. Smaller pores, though they would hold onto the fluids even more strongly, would be expected to offer greater resistance to flow, thus probably reducing the initial rate of absorption.

Hodgson and Berg (1988) showed that in some practical cases the rate of wetting of paper products by aqueous liquids could be increased by use of surfactants. It was proposed that the surfactants acted as wetting agents, decreasing the contact angle of water. Note that according to Eq. 14, such an effect would be predicted to enhance the rate of absorbency. On the other hand, surfactants can be expected to decrease the value of the interfacial tension in the same equation, which would tend to have the opposite effect. Thus the net effect can be expected to depend on various details, such as the structure of the surfactant and its relative affinity for the aqueous phase vs. various interfaces (Simoncic and Roman 2007; Tag et al. 2012).

Kettunen et al. (2001) demonstrated an innovative system in which a thin, photoactive, temporarily hydrophobic TiO2 film prevented water from being absorbed into aerogels formed from nanocellulose; upon illumination the materials became fully wettable (very low contact angle) and acquired the ability to take up 16 times their weight in water.

Diffusion as a Rate-Limiting Step in SAP Absorbency

Diffusion is a mass transport phenomenon that involves the movement of individual molecules due to concentration gradients. The driving force behind diffusion is related to the system tending towards a lower Gibbs free energy. Diffusion plays a role in the dissolving of a material, the movement of a liquid through a liquid, of a gas through a liquid, of a gas through a solid, of a gas through a gas, or of a liquid through a solid. As noted by Kabra and Gehrke (1994), in the case of non-porous absorbent hydrogels, diffusion can be expected to constitute the rate-limiting process governing swelling. Thus, the rates of swelling of porous gels can be on the order of one thousand times faster (see also Omidian et al. 2005). Also, as noted earlier, Tanaka and Fillmore (1979) found that the rate of swelling of individual gel particles was inversely proportional to the square of their size; this finding is suggestive of a diffusion-controlled mechanism due to the fact that the area of each gel particle is proportional to the square of the diameter.

As an initial approach, diffusion into a hydrogel particle might be modeled in a simple way, assuming isotropic properties and neglecting of any changes in those properties during the process of swelling. Equation 16 (Peppas and Hariharan 1994) represents how the second law of diffusion, according to Fick (1855), could be applied in such a model,

 (16)

where C is the concentration of water, t is time, x is the distance of permeation into the material, and v is the convective velocity of the fluid. It should be noted that Eq. 16 applies only in the case of a simple “slab” geometry.

A useful form of the rate equation is obtained when the exponent 0.5, which follows from Fick’s derivation, is replaced by an adjustable value n (Ganji et al. 2010),

 (17)

where Qt is the mass of fluid absorbed per unit of sorbent mass at time equal to t, and  is the value of Qt after a long time. When using this expression to fit experimental data, the system is considered to be Fickian if the fitted value of n equals 0.5 or is somewhat lower than 0.5 (Ganji et al. 2010). Values higher than 0.5 but less than unity indicate anomalous systems in which a combination of mechanisms may govern the rate of absorbency. A value of 1 implies that the rate of fluid uptake is governed only by the rate of relaxation of polymer segments, which is understood to be taking place much more slowly in comparison to the diffusion processes.

Berens and Hopfenberg (1978) noted that although Eq. 17 can be made to fit data corresponding to the initial swelling of hydrogels (up to about 60% of completion), strong deviations tend to be observed at longer times. The following equations were therefore proposed as a more realistic model to fit typical data (Ganji et al. 2010),

where k2 is the empirically-derived rate constant. The values of the constants A and k2 are determined from the slope and intercept of a plot of  versus time.

A key problem inherent in using a simple Fickian model to predict absorption rates of hydrogels is that such an approach does not account for major changes that take place in the gel structure and in the polyelectrolyte segment mobility as a result of the swelling process (Peppas and Sahlin 1989; Singh and Weber 1996; Afif and Grme 2002; Rajagopa 2003; Liu et al.2005b). To overcome this deficiency, Peppas and Hariharan (1994) proposed non-Fickian diffusion models in which an outer, swollen layer of each hydrogel particle was presumed to gradually progress inward toward the center of the hydrogel.

A cylindrical model, as shown in Fig. 10, was used to represent the situation in an ideal case. The rate of diffusion of water within the swollen layer was presumed to be much higher than the same value within the non-swollen core of each gel. The model depicted in Fig. 10 is consistent with a “slow reorientation of polymer chains”, as proposed by Ganji et al. 2010.

Fig. 10. Model of partly swollen hydrogel as a segment of an infinite cylinder, to enable derivation of a non-Fickian diffusion-based rate of absorbency. The model is based on an assumed much greater diffusion rate within a rubbery (already swollen) layer.

Peppas and Sahlin (1989) developed the following equation to account for such systems:

Qt = Q (k1 t + k2 0.5 ) (20)

More complex systems have been considered by invoking second-order diffusion kinetics (Yin et al. 2002). Li and Tanaka (1990) developed an approach taking account of gradients of stress within the hydrogel. Refinements related to effects of such details as variable surface concentrations, history-dependent diffusion rates, stresses between different parts of a gel, and relaxation of polymer segments have been further discussed by Ganji et al. (2010).

LIFE CYCLE AND PRODUCT DEVELOPMENT ISSUES

In view of the high volume of current usage of absorbent products, compounded by the waste inherent in the one-time usage and the non-renewable nature of their petroleum-based components, it is important to consider approaches that can be used to improve the environmental compatibility of absorbent products in future years. Some key elements that have been considered, in this regard, involve biodegradability of SAPs, life cycle analysis, and product development strategies.

Within this section, a life cycle perspective will be used to identify key areas where cellulose-based super absorbents could potentially reduce environmental impacts of absorbent products. Additionally, key areas of uncertainty surrounding environmental impacts will be discussed. A brief discussion of the basics of life cycle assessment is given below, followed by a discussion related to superabsorbents with or without the inclusion of cellulose-based components.

Life Cycle Assessment Applied to Cellulosic SAPs

Life cycle assessment (LCA) can be defined as a way to assess potential environ-mental impacts associated with a product, process, or service; this generally entails making an inventory of the relevant inputs of materials and energy, as well as releases to the environment, evaluating various potential environmental impacts associated with items on the inventory, and then weighing the results of the analysis in order to make more appropriate decisions (EPA 2012). The use of LCA provides additional informa-tion surrounding all the environmental impacts of a product, process, or service to decision makers. By examining a variety of environmental impacts, burden shifting can be minimized through systems thinking. For example, burden shifting can occur when only greenhouse gases (GHG) or only solid wastes are examined. The technique of LCA can be broken down into four distinct steps, (1) goal and scope definition, (2) inventory analysis, (3) impact assessment, and (4) interpretation. For more complete LCA guidance, ISO 14044 provides explanations and a more detailed procedure for performing such studies. The EPA also has many helpful documents online on how to perform LCAs (EPA 2012).

In the following discussion, diaper production and use will be considered as an example, as there is a significant body of literature surrounding diapers, and they employ traditional SAPs that could be partially or wholly replaced in some products by cellulose-based SAPs.

Product composition

Diapers have the primary function of relieving the issues of incontinence by absorbing and containing solids and fluids. Diapers were originally made from cotton and were washable and reusable. Over the past 50 years, the use of reusable diapers has been replaced, in part, with disposable diapers (Edana 2005). According to Edana, disposable diapers have several main components that work together to remove liquids from the skin and store them until disposal. In this construction, there is a soft nonwoven surface that facilitates moisture removal and is comfortable to the skin. Under the top layer, there is a distribution layer of material that transfers the liquids to the super-absorbent layer. Within the core of the diaper, the superabsorbents are made of SAPs and cellulose pulp store liquids. The internal materials of the diaper are contained by a breathable but aqueous fluid-resistant polyethylene nonwoven fabric or film to prevent moisture transfer from the diaper to surrounding surfaces. These different components require raw materials that are made from a variety of materials. According to Edana (2005), cellulose fluff pulp is the largest component of diapers at 43%, and super absorbents are the second largest component at 27%, as shown in Fig. 11.

Within this discussion, the superabsorbents making up approximately 27% of a diaper will be considered for substitution with a blend of cellulose- and petroleum-based SAPs. Assuming that half the superabsorbent is cellulose-based, the total renewable material content of the diaper would then increase to 56.5%. Since raw materials can account for 84% of fossil fuel use and 64% of global warming potential of a typical diaper (Weisbrod and Van Hoof 2012), this material substitution may be significant. Additionally, Mirabella et al. 2013 found that petroleum-based polymer substitution with biobased polymer (polylactic Acid, PLA) may reduce some environmental impact categories on a cradle-to-gate basis.

Fig. 11. Average diaper material composition, percentages shown on a per weight basis (based on data from Edana 2005)

Carbon captured in cellulose SAPs

SAPs are traditionally made from petroleum-based polymers, which are a non-renewable resource. Cellulose, when sourced from sustainably managed forests, represents a material that is not only renewable, but also removes CO2 from the environ-ment through photosynthesis. Incorporating cellulose-based materials into products can achieve significant reductions in GHG emissions due to the carbon removed from the atmosphere contained in the product. Thus, by incorporating cellulose-based SAPs into a diaper, the material would be effectively storing CO2 that was once in the atmosphere and may be credited for this in an LCA. The cellulose-based SAP does contain carbon sourced from the atmosphere; however, the fossil fuel and chemical use in the production of the cellulose-based SAP will also play a significant role in the overall global warming potential of the diaper (Johnson 1994). Without a full LCA, it is impossible to understand the full implications of this substitution; however, the GHG emissions would likely decrease relative to usage of the current product mix.

Product Use and Reuse

Towards a launderable enhanced sorbent

Based on a consumer survey (Smith and Sheeran 1992) it would appear that there could be strong demand for products that can combine favorable features of current disposable absorbent products together with features of conventional absorbents. In particular, it would be desirable to combine high absorbency levels together with suitability for washing and repeated usage. Up to this point it appears that key challenges facing the development of such products involve suitability for washing and drying.

Beskisiz et al. (2009) showed that some textiles comprising superabsorbent fibers can be laundered and even dry-cleaned without excessive loss of their component materials and absorbent nature. The specialized fabric exhibited higher absorbency than conventional fabric; however, it also required a longer drying time after laundering.

The high levels of water held within conventional SAPs after their use, in addition to the gelatinous nature of the swollen material, appear to present a serious challenge in terms of both washing and drying. What is needed, it would appear, is some kind of mechanism by which the swollen nature of the hydrogels can be reversed, such that most of the imbibed liquid can be released during the washing process. In particular, to minimize the time and cost of drying, one would want any SAPs to be in a mostly non-swollen condition before the start of a conventional drying operation. As noted earlier in this article, various researchers have demonstrated thermally-sensitive hydrogels that incorporate a cellulosic component (Harsh and Gehrke 1991; Kuwabara and Kubota 1996; Çaykara et al 2006; Chang et al. 2009b). Under laboratory conditions many such SAPs have been shown to release 80 to 90% of their water content upon heating to 40 oC (Kuwabara and Kubota 1996; Çaykara et al 2006; Ma et al. 2007). It would seem reasonable that such systems could be further developed, along with adjustments in the capabilities of washing machines, such that home laundering of “thermally smart” reusable superabsorbent products could become a practical reality.

Laundering superabsorbents could decrease landfill waste and reduce raw material usage; however, laundering requires large quantities of water and energy (Aumônier and Collins 2005). Many LCA studies have been carried out for disposable and reusable cloth diapers (Aumônier and Collins 2005; Olive 2004). These studies generally indicate that disposable diapers have higher energy use in the production stage, while reusable diapers are most energy-intensive during the use phase when laundering. By introducing a laundering step to a “semi disposable” diaper, the overall energy usage would likely increase. Factors such as number of reuses possible, wash water temperature, water content of diaper entering the drying system, and fluid flow properties from the diaper are critical to understanding the environmental implications of a reusable disposable diaper.

The “thermally smart” superabsorbents could potentially decrease the energy used in laundering and specifically drying. However, significant quantities of energy would be required to heat wash water to 40 oC or another temperature level required, which may offset the drying benefits of reduced water content entering the dryer. As learned from other textile LCA studies, washing soiled garments in cold water followed by line air-drying can be regarded as the laundering option having the lowest environmental impacts (Levi Strauss & Co. 2009; O’Brien et al. 2009). It is unlikely that such a laundering option would be possible with a superabsorbent diaper.

Product End of Life Considerations

Biodegradability of absorbent materials

Several studies have recommended the use of cellulose-based materials in absorbent products in order to achieve suitable biodegradability (Cooke 1990; Lokhande and Gotmare 1999; Fei et al. 2000; Dutkiewicz 2002; Lionetto et al. 2003, 2005; Chang et al. 2008; Demitri et al. 2008; Singha and Rana 2011). In particular, various studies have demonstrated biodegradable character of hydrogels that contain a cellulose-derived component (Bin et al. 2000; Wach et al. 2001, 2002; Lionetto et al. 2003; Sahoo et al. 2005; Yoshimura et al. 2005, 2006; Wang et al. 2008; Feng et al. 2010a; Kono and Fujita 2012; Xie et al. 2012). By contrast, it has long been clear that SAPs based on acrylamide generally have limited biodegradability (Wichterle and Lim 1960). Yoshimura et al. (2005, 2006) showed that ester-based cellulosic absorbents, prepared with maleic anhydride, offered particular promise in terms of enhanced biodegradability. On the other hand, Sahoo et al. (2005) showed that grafting sometimes can be used as a way to decrease the excessive biodegradability of starch present in the mixture from which SAPs are being prepared.

Despite the demonstrated superior performance of superabsorbent hydrogel materials, it is clear from the product selection available at a typical supermarket that cotton-based products continue to attract substantial customer demand. Some of this demand might be prompted by the fact that typical cotton-based products can be laundered and used multiple times, thus offering a perceived cost savings. In addition, it is likely that customers’ environmental concerns contribute to the purchase decision (Smith and Sheeran 1992). For instance, a disposable liner for diaper-type products, composed entirely of “natural materials” has been patented (Flores Gonzales 2010). The core material was comprised of cotton. Macaulay et al. (2004) carried out a pilot study evaluating the washing and reuse of absorbent garments for adults with urinary incontinence. Guidelines were given regarding usage of such items. Performance and utility issues were compared relative to disposable incontinence products.

Despite the popular belief that biodegradable materials are inherently better for the environment, there are some situations and limiting factors which complicate this issue. Biodegradable products are often associated with decreased landfill space requirements and material degradation over time. This biodegradation requires oxygen, microbes, and other environmental conditions. When biodegradable products end up in landfills, they generally require as much volume as the non-biodegradable alternative. Additionally, when biodegradable products are buried in oxygen-deprived landfills, they degrade slowly, if at all (EPA 2012). If biodegradation does occur, it will be largely anaerobic, producing methane that is collected or burned in most landfill scenarios in the USA. In some landfills, particularly common in developing countries, methane is not collected, resulting in large contributions to global warming potential, because methane has more than 20 times the global warming effect of CO(Bare 2011). Furthermore, Levis and Barlaz (2011) indicated that biodegradable materials in fact can produce higher end-of-life GHG emissions with similar upstream emissions, resulting in a higher net global warming potential across the product life cycle.

Biodegradable products that are engineered to decompose rapidly often decompose before the methane collection systems used in landfills have been made operational; as a result such products can emit GHGs directly to the environment (Levis and Barlaz 2011). However, when biodegradable diapers are collected and processed in a composting operation, Colón et al.(2011) found that 50% of the carbon contained in the diaper was degraded to CO2.

The effects of biodegradable products on landfill space are controversial as well. Once a landfill is capped, the space required beneath the top cap layer is set. If the internal components of the landfill degrade, the size of the landfill will not necessarily decrease. The contents may become less dense as mass of material is removed as CO2 and methane; however, the landfill will not likely shrink in size (foot print or height).

On the other hand, one clear benefit of using biodegradable materials does not involve the landfill but rather the inevitability of litter left to decay on the ground. Biodegradable materials left on the surface of the ground degrade at higher levels due to proper moisture content, oxygen levels, and microbial activity. In some applications, biodegradable materials would reduce the quantity of human wastes in the environment; however, one may question this benefit in the case of diapers as they are not traditionally thought to be a litter problem.

Engineering of Sorbent Products for the Application

In general, the design of new highly absorbent products to better meet the needs of consumers must take account of such issues as compact dimensions, dependability of performance, and health effects (Fader et al. 2010). As one looks to the future development of sorbent products, typical sorbent performance measurements will remain important. For example, reducing the costs of absorbents will open the use of the materials to wider applications as soil amendments and other high-volume low-value applications. This represents an opportunity for superabsorbents made with natural polymers such as cellulose, starch, and hemicellulose. According to Kono and Fujita (2012), the projected cost of manufacturing SAPs based on carboxymethylcellulose has tended to discourage implementation of production of superabsorbents based on cellulose. However, high cost was also cited as a key concern for SAP usage in general (Liang et al. 2009). Furthermore, functional improvements such as higher density, increased mechanical strength, or great absorbency per unit mass may lead to the use of sorbent in new products or new product forms. While progress along these traditional lines will lead to innovations, it is perhaps more important to consider the role of sorbents in a broader sense and how novel functionality can lead to new engineered products.

A universal absorbent can be defined as a material that can take up a wide variety of liquids. While this class of products is limited in number, some every-day materials such as wood fibers, ground wood pulp, or even bark will act as universal absorbents holding both polar and non-polar liquids within the structure of the materials. It may be possible to engineer universal suber-absorbents that take up both polar and non-polar liquids by constructing a block co-polymer in such a way that there are domains of polar and non-polar material (Loh 2002). These types of materials may or may not have to be crosslinked to prevent solubility and their subsequent functioning as suber-absorbents. A material constructed as in Fig. 12 would have domains that are insoluble in polar liquids and domains that are insoluble in non-polar liquids.

Fig. 12. Block co-polymer containing both polar and non-polar domains. A: Immersion of material into non-polar solvent causes the non-polar regions to swell and absorb liquid, while polar regions prevent the dissolution of the polymer. B: Immersion of material into polar solvent causes the polar regions to swell and absorb liquid, while non-polar regions prevent the dissolution of the polymer.

The result would be a material that is capable to taking up many different types of liquids. Furthermore, if a material such as this is not crosslinked, then it could potentially provide a “smart” functionality. For instance, it could be optimized to absorb a polar liquid, then when having a non-polar liquid added, to dissolve and release the polar liquid. This could prove useful in many fields including the oil and gas industry.

The ability of materials to respond to environmental stimuli is also of high potential value. Previously in this review it was shown that absorbency is affected by a number of factors including temperature, pH, salinity, and pressure. While for current commercial absorbents the change in the absorptive properties might be relatively small, these responses could be enhanced through the addition of crosslinkers consisting of polymers that are highly responsive of various stimuli. Interestingly, when absorbents take up liquids, they are nearly always accompanied by an associated change in physical dimensions. This allows for one to create products that can respond physically to a chemical or chemi-physical interaction between the absorbent and a liquid. Thus, one may engineer these materials for micro and nanotechnology where movement is needed to be generated in response to a stimulus. In the literature, many examples exist of polymers that can respond to various stimuli by changing conformation (Schmaljohann 2006; Jeong and Gutpwska 2002). This change in conformation can be taken advantage of to create smart absorbents.

Figure 13 shows a schematic example of the effect of a stimuli-responsive polymer acting as a crosslinker in a superabsorbent material. As shown, such as stimuli-responsive material can act to increase or decrease absorption of liquids depending upon the stimuli. Integration of this functionality into absorbents will open the doors to using these materials for products and purposes never before considered.

Fig. 13. Schematic view of a lyophilic polymer crosslinked with a light-sensitive polymer. Upon exposure to light, the crosslink polymer conformation changes, allowing for the absorbent to take up a large fraction of liquid. If this is a reversible process, then removing the stimulus could result in the expulsion of the liquid.

Absorbents may also be used to selectively absorb materials from liquids, making them useful for multiple applications. For example, Ayoub et al. (2013) demonstrated the use of an absorbent composed of both anionic and cationic polymers for isolating dissolved salt from water. One could also design absorbents that bind or destroy biologically active agents such as viruses or bacteria. This would create an absorbent that not only binds liquids but also disinfects at the same time. The ability to uptake, retain, and disinfect liquids may find uses in a variety of applications and products.

The recyclability of the absorbents is a challenge due to the very nature of absorbents. Absorbents are generally designed to tenaciously retain the liquids they absorb. Thus, subsequently removing those liquids is often energy-intensive or overly time-consuming. The design of an absorbent that could be repeatedly used would potentially be beneficial from both an environmental standpoint and economic standpoint. This may occur by combining together the concept of universal absorbent and or smart absorbents. By creating an absorbent that will release its absorbed liquid upon a stimulus or change in liquid, it is conceivable that absorbent could be easily and conveniently recycled. For example, the conformation of a polymer may be controlled by exposure to certain wavelengths of light. If this polymer is used as a crosslinker in a superabsorbent, then the shrinking and swelling of the polymer could be controlled by the exposure to light. This simple light exposure could be used to expel the retained liquid from the absorbent. This would make recycling of an absorbent as simple as turning on a switch.

Ultimately, absorbents can be used in a wide variety of applications. This utility can be rapidly expanded by integrating novel functionalities into the absorbent. Much of the basic science exists to create these materials. It is simply a matter of economics and desire for such materials to be developed.

CLOSING COMMENTS

As shown by many of the authors cited in this article, the usage of cellulose-based materials in absorbent products represents an intriguing field for academic research and product development. There can be a feeling of magic when one considers the huge ratios that can be achieved between the amounts of aqueous fluid that can be absorbed by much smaller amounts of sorbent material. Several domains of science overlap, including new approaches to synthesis of novel composition and means of crosslinking. The physical chemistry aspects are also fascinating, involving new ways to think about contributions to entropy, Donnan equilibria, the effects of nanostructures on absorption kinetics, and the like. Moreover, absorbent materials are economically important and have large environmental implications. Thus, there is a chance to be a hero – or maybe even to earn some money – if one can come up with an improved practical solution to absorbing fluids in challenging situations.

Although much progress has been achieved in understanding the role of cellulose-based materials and copolymers in the absorption of aqueous fluids, many gaps in knowledge are still waiting to be filled. For example, although it has been claimed that interpenetrating networks of hydrophilic polymers hold promise for superior absorbency (Myung et al. 2008), the theoretical basis of such a claim, if valid, would need to be established. Although incorporation of cellulosic components into acrylamide-based absorbent products has potential to make them more biodegradable, it is unclear whether and how quickly such degradation might take place under typical conditions of landfilling. Relatively little has been reported about the effects of squeezing highly absorbent products in an effort to expel the imbibed aqueous fluid. Though resistance to squeezing is sometimes considered to be a characteristic of superabsorbent materials (Laftah 2011), it is clear that application of pressure can at least partly reverse the osmotically-drive effects (Sannino et al. 2004b; Li et al. 2013). Though the concepts of osmotic pressure and the Donnan equilibrium have provided a basis for understanding absorbency phenomena (Grignon and Scallan 1980), it is unclear how or whether these concepts can be applied to the different kinds of water that are found in practical systems (Maloney et al. 1998), i.e. non-freezing bound water, bound water having a depressed freezing point, or water within isolated pockets of a hydrogel where the freezing point may be similar to that of bulk water. The contributions of such factors as the entropy of mixing, the entropy due to the freeing up of ions, and the effects of ionic and van der Waals forces within hydrogels during fluid absorption need to be better quantified.

The distance between various simplified theoretical models and the practical realities of commercially viable sorbent products will need to be narrowed (Ganji et al. 2010). On the one hand, experiments can be contrived in which independent variables such as the distance between crosslink points are precisely engineered. On the other hand, there is a need for applications of more comprehensive mathematical approaches that incorporate details of the nanostructure of hydrogel materials in a more realistic way.

Various less-glamorous aspects of the usage of highly absorbent materials incorporating cellulosic materials urgently need research attention. Life cycle assessment, if it is to be done well, requires a lot of work to gather high-quality data and to accurately weigh the impact of various aspects of absorbent usage on the environment. Filling the absorbent with fluid appears to have received a disproportionate amount of the attention from academics. More attention needs to be given to what can be done with the absorbent products after they have been used – especially when they are contaminated with bodily fluids and smeared with excrement, etc. On May 25, 1961 US president John F. Kennedy challenged scientists and engineers to put a man on the moon and bring him safely back to Earth. A similar challenge could be given at this point for scientists and engineers to design a highly absorbent diaper such that it can be fully laundered, with no significant loss of either its dry mass or absorbency, and requiring little more energy during laundering than is needed for conventional items of reusable clothing.

REFERENCES CITED

Abbas, S., Bitis, R., Larsson, A., Larsson, M., and Malmgren, K. (2011). “Superabsorbent polymer composite comprising a superabsorbent polymer and cellulosic nanofibrils,” US Pat. 6,844,066.

Abd El-Mohdy, H. L. (2007). “Water sorption behavior of CMC/PAM hydrogels prepared by -irradiation and release of potassium nitrate as agrochemical,” Reactive & Functional Polymers 67, 1094-1102.

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

Adamson, A. W., and Gast, A. P. (1997). Physical Chemistry of Surfaces, Wiley, New York.

Afif, A. E., and Grme, M. (2002). “Non-Fickian mass transport in polymers,” J. Rheol. 46, 591-628.

Ahmed, I, Jones, A. M., and Smith, S. J. (2005). “Cellulose material with improved absorbency,” US Pat. 6,906,131 B2.

Akar, E., Altinisik, A., and Seki, Y. (2012). “Preparation of pH- and ionic-strength responsive biodegradable fumaric acid crosslinked carboxymethyl cellulose,” Carbohydr. Polymers90(4), 1634-1641.

Ali, A. E. H., El-Rehim, H. A. A., Kamal, H., and Hegazy, D. E. S. A. (2008). “Synthesis of carboxymethyl cellulose based drug carrier hydrogel using ionizing radiation for possible use as site specific delivery system,” J. Macromol. Sci. Pt. A – Pure Appl. Chem. 45(8), 628-634.

Aliouche, D., Ait-Amar, H., and Lahfati, K. (2001). “Experimental determination and interpretation of absorption capacities of liquids by biotextiles with complex absorbent structure – Influence of superabsorbent polymer gels,” Chem. Eng. J. 81(1-3), 317-322.

Aliouche, D., Sid, B., and Ait-Amar, H. (2006). “Graft-copolymerization of acrylic monomers onto cellulose. Influence on fibre swelling and absorbency,” Anal. Chimie – Sci. Mater.31(5), 527-540.

Anirudhan, T. S., Tharun, A. R., and Rejeena, S. R. (2011). “Investigation on poly(methacrylic acid)-grafted cellulose/bentonite superabsorbent composite: Synthesis, characterization, and adsorption characteristics of bovine serum albumin,” Indust. Eng. Chem. Res. 50(4), 1866-1874.

Argüelles-Monal, W., Gárciga, M., and Peniche-Covas, C. (1990). “Study of the stoichiometric polyelectrolyte complex between chitosan and carboxymethyl cellulose,” Polymer Bull. 23, 307-313.

Askari, F., Nafisi, S., Omidian, H., and Hashemi, S. A. (1993). “Synthesis and characterization of acrylic-based superabsorbents,” J. Appl. Polymer Sci. 50(10), 1851-1855.

Aspler, J. S., Davis, S., and Lyne, M. B. (1987). “The surface chemistry of paper in relation to dynamic wetting and sorption of water and lithographic fountain solutions,” J. Pulp Paper Sci. 13(2), J55-J60.

ASTM (1983). Technical Supplement: A Compilation of Procedures and Guidelines for Textile Products, AATCC C2C Committee and the ATSM Intl. Committee D13 on Textiles, AATCC, Research Triangle Park, NC, USA.

Athawale, V. D., and Lele, V. (2001). “Recent trends in hydrogels based on starch-graft-acrylic acid: A review,” Starch-Starke 53(1), 7-13.

Aumônier, S., and Collins, M. (2005). “Life cycle assessment of disposable and reusable nappies in the UK,” Environmental Agency, May.

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., published online, DOI: 10.1021/sc300166m.

Azizi Samir, M. A. S., Alloin, F., and Dufresne, A. (2005). “Review of recent research into cellulosic whiskers, their properties and their application in nanocomposite field,” Biomacromol. 6(2), 612-626.

Back, E. L., and Allen, L. (2000). Pitch Control, Wood Resin and Deresination, TAPPI Press, Atlanta, 392 pp.

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.

Baker, J. P., Blanch, H. W., and Prausnitz, J. M. (1995). “Swelling properties of acrylamide-based ampholytic hydrogels – Comparison of experimental with theory,” Polymer 1995, 36(5), 1061-1069

Bao, Y., Ma, J. Z., and Li, N. (2011). “Synthesis and swelling behaviors of sodium carboxymethyl cellulose-g-poly(AA-co-AM-co-AMPS)/MMT superabsorbent hydrogel,” Carbohyd. Polymers 84(1), 76-82.

Bao, Y., Ma, J. Z., and Sun, Y. G. (2012). “Swelling behaviors of organic/inorganic composites based on various cellulose derivatives and inorganic particles,” Carbohydr. Polymers 88(2), 589-595.

Barbucci, R., Magnani, A., and Consumi, M. (2000). “Swelling behavior of carboxymethyl cellulose hydrogels in relation to cross-linking, pH, and charge density,” Macromol. 33(20), 7475-7480.

Barcus, L. R., and Bjorkquist, W. D. (1997). “Fibers and pulps for papermaking based on chemical combination of poly(acrylate-co-itaconate), polyol and cellulosic fiber,” US Pat.5,698,074.

Bare, J. (2011). “TRACI 2.0: The tool for the reduction and assessment of chemical and other environmental impacts 2.0,” Clean Technologies and Environmental Policy 13(5), 687-696.

Berens, A. R., and Hopfenberg, H. B. (1978). “Diffusion and relaxation in glassy polymer powders. 2. Separation of diffusion and relaxation parameters,” Polymer 19, 489-496.

Bergenstrahle, M., Wohlert, J., Larsson, P. T., Mazeau, K., and Berglund, L. A. (2008). “Dynamics of cellulose-water interfaces: NMR spin-lattice relaxation times calculated from atomistic computer simulations,” J. Phys. Chem. B 112(9), 2590-2595.

Berthold, J., and Salmén, L. (1997). “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.

Beskisiz, E., Ucar, N., and Demir, A. (2009). “The effects of super absorbent fibers on the washing, dry cleaning and drying behavior of knitted fabrics,” Textile Res. J. 79(16), 1459-1466.

Beuther, P. D., Veith, M. W., and Zwick, K. J. (2010). “Characterization of absorbent flow rate in towel and tissue,” J. Eng. Fibers Fabrics 5(2), 1-7.

Bhandari, P. N., Jones, D. D., and Hanna, M. A. (2012). “Carboxymethylation of cellulose using reactive extrusion,” Carbohydr. Polym. 87(3), 2246-2254.

Billmeyer, F. W. (1971). Textbook of Polymer Science, Wiley-Interscience, New York.

Bin, F., Wach, R. A., Minomo, H., Yoshii, F., and Kume, T. (2000). “Hydrogel of biodegradable cellulose derivatives. I. Radiation-induced crosslinking of CMC,” J. Appl. Polymer Sci.78(2), 278-283.

Bourges, X., Weiss, P., Daculsi, P., and Legeay, G. (2002). “Synthesis and general properties of silated-hydroxypropyl methylcellulose in prospect of biomedical use,” Adv. Colloid Interface Sci. 99(3), 215-228.

Brannon-Peppas, L., and Peppas, N. A. (1990). “The equilibrium swelling behavior of porous and non-porous hydrogels,” In: Absorbent Polymer Technology, Brannon-Peppas, L., and Harland, R. S. (eds.), Studies in Polymer Science Ser. 8, Elsevier, Amsterdam.

Breier, R. (1997). “Die Strangveredlung von Lyocell – Aktueller Stand der Praxis,” Lenzinger Berichte 76, 108-111.

Bristow, J. A. (1967). “Liquid adsorption into paper during short time intervals,” Svensk Papperstidning 19(15), 623-629.

Bristow, J. A. (1972). “Swelling of fiber building boards on immersion in water,” Svensk Papperstidning 75(21), 847-852.

Brodin, F. W., Lund, K., Brelid, H., and Theliander, H. (2012). “Reinforced absorbent material: A cellulosic composite of TEMPO-oxidized MFC and CTMP fibres,” Cellulose 19(4), 1413-1423.

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

Brydson, J. A. (1975). Plastics Materials, Butterworth, Boston, loc cit. p. 59.

Buchholz, F. L., and Graham, A. T. (2005). Modern Superabsorbent Polymer Technology, Wiley-VCH, New York.

Buchholz, F. L., and Peppas, N. A. (1994). Superabsorbent Polymers. Science and Technology, ACS Symp. Ser. 573, Amer. Chem. Soc., Washington, DC.

Buchholz, F. L., Pesce, S. R., and Powell, C. L. (2005). “Deswelling stresses and reduced swelling of superabsorbent polymer in composites of fiber and superabsorbent polymers,” J. Appl. Polymer Sci. 98(6), 2493-2507.

Buyanov, A. L., Gofman, I. V., Revel’skava, L. G., Khripnov, A. K., and Tkachenko, A. A. (2010). “Anisotropic swelling and mechanical behavior of composite bacterial cellulose – poly(acrylamide or acrylamide-sodium acrylate) hydrogels,” J. Mechan. Behav. Biomed. Mater. 3, 102-111.

Cai, J., and Zhang, L. (2006). “Unique gelation behavior of cellulose in NaOH/urea aqueous solution,” Biomacromol. 7(1), 183-189.

Campbell, W. B. (1959). “The mechanism of bonding,” Tappi 42(12), 999-1001.

Cao, B., Tschirner, U., and Ramaswamy, S. (1998). “Impact of pulp chemical composition on recycling,” Tappi J. 81(12), 119-127.

Capitani, D., Emanuele, M. C., Bella, J., Segre, A. L., Attanasio, D., Focher, B., and Capretti, G. (1990). “1H NMR relaxation study of cellulose and water interaction in paper,” TAPPI J.82(9), 117-124.

Carman, P. C. (1938). “Fluid flow through granular beds,” Trans. Inst. Chem. Engrs. (London) 16, 168-188.

Carman, P. C. (1939). “Determination of the specific surface of powders. Part 2,” J. Soc. Chem. Ind. 58, 2-7.

Carrier, W. D. (2002). “Goodby, Hazen; Hello, Kozeny-Carman,” J. Geotech. Geoenviron. Eng. (Nov.), 1054-1056.

Cassie, A. B. D., and Baxter, S. (1944). “Wettability of porous surfaces,” Trans. Faraday Soc. 40, 546-551.

Çaykara, T., Sengiil, G., and Birlik, G. (2006). “Preparation and swelling properties of temperature-sensitive semi-interpenetrating polymer networks composed of poly(N-tert-butylacrylamide)-co-acrylamide] and hydroxypropyl cellulose,” Macromol. Mater. Eng. 291(9), 1044-1051.

Chang, C., Lue, A., and Zhang, L. (2008). “Effect of crosslinking methods on the structure and properties of cellulose/PVA hydrogels,” Macromol. Chem. Phys. 209(12), 1266-1273.

Chang, C. Y, Duan, G., and Zhang, L. N. (2009a). “Fabrication and characterization of novel macroporous cellulose-alginate hydrogels,” Polymer 50, 5467-5473.

Chang, C., Han, K., and Zhang, L. (2009b). “Structure and properties of cellulose/poly(N-isopropylacrylamide) hydrogels prepared by IPN strategy,” Polymer Advan. Technol. 22(9), 1329-1334.

Chang, C. Y., Duan, B., Cai, J., and Zhang, L. N. (2010). “Superabsorbent hydrogels based on cellulose for smart swelling and controllable delivery,” Eur. Polymer J. 46(1), 92-100.

Chang, C. Y., and Zhang, L. N. (2011). “Cellulose-based hydrogels: Present status and application prospects,” Carbohydrate Polymers 84(1), 40-53.

Chapuis, R. P., and Auberton, M. (2003). “On the use of the Kozeny-Carman equation to predict the hydraulic conductivity of soils,” Can. J. Geotech. 40, 616-628.

Charreau, H., Foresti, M. L., and Vazquez, A. (2013). “Nanocellulose patents trends: A comprehensive review on patents on cellulose nanocrystals, microfibrillated and bacterial cellulose,” Recent Patents Nanotechnol. 7(1), 56-70.

Chatterjee, P. K., and Gupta, B. S. (eds.) (2002). Absorbent Technology, Vol. 13 of Textile Science and Technology series, Elsevier, Amsterdam.

Chen, X. M., Burger, C., Fang, D. F., Ruan, D., Zhang, L., Hsiao, B. S., and Chu, B. (2006). “X-ray studies of regenerated cellulose fibers wet spun from cotton linter pulp in NaOH/thiourea aqueous solutions,” Polymer 47(8), 2839-2848.

Chen, Y., Liu, Y. F., and Tan, H. M. (2008). “Preparation of macroporous cellulose-based superabsorbent polymer through the precipitation method,” BioResources 3(1), 247-254.

Chen, Y., and Tan, H. M. (2006). “Crosslinked carboxymethylchitosan-g-poly(acrylic acid) copolymer as a novel superabsorbent polymer,” Carbohyd. Res. 341(7), 887-896.

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.

Colón, J., Ruggieri, L., Sánchez, A., González, A., and Puig, I. (2011). “Possibilities of composting disposable diapers with municipal solid wastes,” Waste Management & Research 29(3), 249-259.

Cooke, T. F. (1990), “Biodegradability of polymers and fibers – A review of the literature,” J. Polymer Eng. 9(3), 171-211.

Cundall, R. B., Lawton, J. F., Murray, D., and Phillips, G. O. (1979). “Polyelectrolyte complexes. 1. The effect of pH and ionic strength on the stoichiometry of model polycation – polyanion complexes,” Makromol. Chem. 180, 2913-2922.

Dai, Q., and Kadla, J. F. (2009). “Effect of nanofillers on carboxymethyl cellulose/ hydroethyl cellulose hydrogels,” J. Appl. Polymer Sci. 114(3), 1664-1669.

Darcy, H. (1856). Les Fontaines Publiques de la Ville de Dijon, Dalmont, Paris.

Demitri, C., Del Sole, R., Scalera, F., Sannino, A., Vasapollo, G., Maffezzoli, A., Ambrosio, L., and Nicolais, L. (2008). “Novel superabsorbent cellulose-based hydrogels crosslinked with citric acid,” J. Appl. Polymer Sci. 110(4), 2453-2460.

Deo, H. T., and Gotmare, V. D. (1999). “Acrylonitrile monomer grafting on gray cotton to impart high water absorbency,” J. Appl. Polymer Sci. 72(7), 887-894.

De Oliveira, W., and Glasser, W. G. (1996). “Hydrogels from polysaccharides. I. Cellulose beads for chromatographic support,” J. Appl. Polymer Sci. 60(1), 63-73.

Dogu, S., Kilic, M., and Okay, O. (2009). “Collapse of acrylamide-based polyampholyte hydrogels in water,” J. Appl. Polymer Sci. 113(3), 1375-1382.

Dragan, S., and Cristea, M. (2001). “Influence of low-molecular-weight salts on the formation of polyelectrolyte complexes based on polycations with quaternary ammonium salt groups in the main chain and poly(sodium acrylate),” Eur. Polym. J. 37, 1571-1575.

Drost-Hansen, W. (1969). “Structure of water near solid interfaces,” Ind. Eng. Chem. 61(11), 10-47.

Duran, N., Lemes, A. P., and Seabra, A. B. (2012). “Review of cellulose nanocrystals patents: preparation, composites and general applications,” Recent Patents Nanotechnol. 6(1), 16-28.

Dutkiewicz, J. K. (2002). “Superabsorbent materials from shellfish waste – A review,” J. Biomed. Mater. Res. 63(3), 373-381.

Edana. (2005). “Sustainability report: Baby diapers and incontinence products,” Report 2005. Accessed 6/3/13, http://www.edana.org/docs/default-source/default-document-library/sustainability-report-baby-diapers-and-incontinence-products-2005.pdf?sfvrsn=2.

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.

Eklund, D., and Salminen, P. (1987). “Water sorption in paper during short times,” Appita 40(5), 340-346.

El-Din, H. M. N., Alla, S. G. A., and El-Naggar, A. W. M. (2010). “Swelling and drug release properties of acrylamide/carboxymethyl cellulose networks formed by gamma irradiation,” Radiation Physics Chem. 79(6), 725-730.

Eldin, M. S. M., El-Sherif, H. M., Soliman, E. A., Elzatahry, A. A., and Omer, A. M. (2011). “Polyacrylamide-grafted carboxymethyl cellulose: Smart pH-sensitive hydrogel for protein concentration,” J. Appl. Polymer Sci. 122(1), 469-479.

Eldin, M. S. M., Omer, A. M., Soliman, E. A., and Hassan, E. A. (2013). “Super-absorbent polyacrylamide grafted carboxymethyl cellulose pH sensitive hydrogel: 1. Preparation and characterization,” Desalination Water Treat. 51(16-18), 3196-3206.

El Salmawi, K. M. (2007). “Application of polyvinyl alcohol (PVA)/carboxymethyl cellulose (CMC) hydrogel produced by conventional crosslinking or by freezing and thawing,” J. Macromol. Sci. Pt. A. Pure Appl. Chem. 44(4-6), 619-624.

El Salmawi, K. M., and Ibrahim, S. M. (2011). “Characterization of superabsorbent carboxymethylcellulose/clay hydrogel prepared by electron beam irradiation,” Macromol. Res. 19(10), 1029-1034.

EPA (2012). “The degradable debate,” http://www.epa.gov/osw/wycd/catbook/debate.htm. Accessed 6/3/2013.

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

Fader, M., Bliss, D., Cottenden, A., Moore, K., and Norton, C. (2010). “Continence products: Research priorities to improve the lives of people with urinary and/or fecal leakage,” Neurourology and Unrodynamics 29(4), 640-644.

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.

Fanta, G. F., Burr, R. C., and Doane, W. M. (1987). “Graft polymerization of acrylonitrile onto wheat straw,” J. Appl. Polymer Sci. 33, 899-906.

Farag, S., and Al-Afaleq, E. I. (2002). “Preparation and characterization of saponified de-lignified cellulose polyacrylonitrile-graft copolymer,” Carbohyd. Polym. 48(1), 1-5.

Faroongwarng, E., and Sukonrat, P. (2008). “Thermal behavior of water in the selected starch- and cellulose-based polymeric hydrogels,” Intl. J. Pharmaceutics 352(1-2), 152-158.

Fatehi, P., Outhouse, K. C., Xiao, H. N., and Ni, Y. H. (2010). “Debonding performance of various cationic surfactants on networks made of bleached kraft fibers,” Indust. Eng. Chem. Res.49(22), 11402-11407.

Favier, V., Canova, G. R., Shrivastava, S. C., and Cavaille, J. Y. (1997). “Mechanical percolation in cellulose whisker nanocomposites,” Polymer Eng. Sci. 37(10), 1732-1739.

Fei, B., Wach, R. W., Mitomo, H., Yoshii, F., and Kume, T. (2000). “Hydrogel of biodegradable cellulose derivatives. I. Radiation-induced crosslinking of CMC,” J. Appl. Polymer Sci.78(2), 278-283.

Feng, H., Li, J. A., and Wang, L. J. (2010a). “Preparation of biodegradable flax shive cellulose-based superabsorbent polymer under microwave irradiation,” BioResources 5(3), 1484-1495.

Feng, X., and Pelton, R. (2007). “Carboxymethyl cellulose:polyvinylamine complex hydrogel swelling,” Macromol. 40(5), 1624-1630.

Feng, X., Pelton, R., and Leduc, M. (2006). “Mechanical properties of polyelectrolyte complex films based on polyvinylamine and carboxymethyl cellulose,” Indust. Eng. Chem. Res.45(20), 6665-6671.

Feng, Y., Xiao, C. F., and Yang, B. C. (2010b). “Studies on a novel polyacrylonitrile copolymer/cellulose acetate blending water absorbency fiber,” Fibers Polymers 11(6), 838-842.

Fick, A. (1855). “V. On liquid diffusion,” The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science 10(63), 30-39.

Fink, H. P., Weigl, P., Purz, H. J., and Ganster, J. (2001). “Structure formation of regenerated cellulose materials from NMMO-solution,” Prog. Polymer Sci. 26, 1473-1524.

Fleer, G. J., Cohen Stuart, M. A., Scheutjens, J. M. H. M., Cosgrove, T., and Vincent, B. (1993). Polymers at Interfaces, Chapman & Hall, London.

Flores Gonzales, E. C. (2010). “Nappy or panty liner, contains disposable core made entirely from natural materials,” US Pat. 2005222545-A1.

Flory, P. J. (1953). Principles of Polymer Chemistry, Cornel Univ. Press, Ithaca, NY.

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

Gabrielii, I., and Gatenholm, P. (1998). “Preparation and properties of hydrogels based on hemicellulose,” J. Appl. Polymer Sci. 69(8), 1661-1667.

Gabrielii, I., Gatenholm, P., Glasser, W. G., Jain, R. K., and Kenne, L. (2000). “Separation, characterization and hydrogel-formation of hemicellulose from aspen wood,” Carbohyd. Polym.43(4), 367-374.

Gane, P. A. C., Schoelkopt, J., Gantenbein, D., and Schenker, M. (2012). “Process for production of nano-fibrillar cellulose gels,” US Pat. Appl. US2012/0107480A1.

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

Gawel, K., Gao, M., and Stokke, B. T. (2012). “Impregnation of weakly charged anionic microhydrogels with cationic polyelectrolytes and their swelling properties monitored by a high resolution interferometric technique. Transformation from a polyelectrolyte to a polyampholyte hydrogel,” Eur. Polymer J. 48(11), 1949-1959.

Gelin, K., Bodin, A., Gatenholm, P., Mihranyan, A., Edwards, K., and Strømme, M. (2007). “Characterization of water in bacterial cellulose using dielectric spectroscopy and electron microscopy,” Polymer 48(26), 7623-7631.

Gert, E. V. (1996). “Possibilities of nitric acid preparation of powder cellulose forms,” Cellulose 3(4), 217-228.

Gigac, J., and Fiserova, M. (2008). “Influence of pulp refining on tissue paper properties,” TAPPI J. 7(8), 27-32.

Goetz, L. A., Sladky, J. P., and Ragauskas, A. J. (2009). “Analysis of microwave vs. thermally assisted grafting of poly(methyl-vinyl ether co-maleic acid)-polyethylene glycol to birch kraft pulp,” Holzforschung 63(4), 414-417.

Goetz, L. A., Sladky, J. R., and Ragauskas, A. J. (2011). “Preparation of microwave-assisted polymer-grafted softwood kraft pulp fibers. Enhanced water absorbency,” J. Appl. Polymer Sci. 119(1), 387-395.

Grignon, J., and Scallan, A. M. (1980). “Effect of pH and neutral salts upon the swelling of cellulose gels,” J. Appl. Polymer Sci. 25(12), 2829-2833.

Gröndahl, M., Eriksson, L., and Gatenholm, P. (2004). “Material properties of plasticized hardwood xylans for potential application as oxygen barrier films,” Biomacromol. 5(4), 1528-1535.

Guilherme, M. R., Campese, G. M., Radovanovic, E., Rubira, A. F., Feitosa, J. P. A., and Muniz, E. C. (2005). “Morphology and water affinity of superabsorbent hydrogels composed of methacrylated cashew gum and acrylamide with good mechanical properties,” Polymer 46(19), 7867-7873.

Gupta, B. S. (1988). “The effect of structural factors on the absorbent characteristics of nonwovens,” Tappi J. 71(8), 147-152.

Gürdağ, G., Güçlü, G., and Özgümüş, S. (2001). “Graft copolymerization of acrylic acid onto cellulose: Effects of pretreatments and crosslinking agent,” J. Appl. Polym. Sci. 80(12), 2267-2272.

Han, S., Lee, M., and Kim, B. K. (2010). “Cross linking reactions of oxidized cellulose fiber. I. Reactions between dialdehyde cellulose and multifunctional amines on lyocell fabric,” J. Appl. Polymer Sci. 117(2), 682-690.

Harsh, D. C., and Gehrke, S. H. (1991). “Controlling the swelling characteristics of temperature-sensitive cellulose ether hydrogels,” J. Controlled Release 17(2), 175-185.

Heiner, A. P., Kuutti, L., and Teleman, O. (1998). “Comparison of the interface between water and four surfaces of native crystalline cellulose by molecular dynamics simulations,” Carbohydrate Res. 306(1-2), 205-220.

Heiner, A. P., and Teleman, O. (1997). “Interface between monoclinic crystalline cellulose and water: Breakdown of the odd/even duplicity,” Langmuir 13(11), 511-518; (26), 7305 (erratum).

Hiemenz, P. E. (1977). Principles of Colloid and Surface Chemistry, Marcel Dekker, New York.

Herrington, T. M., and Petzold, J. C. (1992). “An investigation into the nature of charge on the surface of papermaking wood pulps. I. Charge/pH isotherms,” Colloids Surf. 64(2), 97-108.

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.

Hirsh, S. G., and Spontak, R. J. (2002). “Temperature-dependent property development in hydrogels derived from hydroxypropylcellulose,” Polymer 43(1), 123-129.

Hodgson, K. T., and Berg, J. C. (1988). “The effect of surfactants on wicking flow in fiber networks,” J. Colloid Interface Sci. 121(1), 22-31.

Hou, Q. X., Liu, W., Liu, Z. H., and Bai, L. L. (2008). “Characteristics of wood cellulose fibers treated with periodate and bisulfite,” Indust. Eng. Chem. Res. 46(23), 7830-7837.

Hu, S.-H., and Ko, Y. C. (2006). “Method of producing twisted, curly fibers,” US. Pat. 6,984,447.

Hua, X., Lale, M., and Owston, T. (2011). “Cellulose nanofilaments and method to produce same,” US Pat. Appl. 2011/0277947A1.

Huang, L. J., Xu, T., Wang, H. T., and Wang, S. F. (2011). “Preparation of super absorbent polymer by carboxymethyl cellulose grafting acrylic acid using low-temperature plasma treatment,” Advan. Mater. 239-242, 2578-2583.

Hubbe, M. A. (2006). “Sensing the electrokinetic potential of cellulosic fiber surfaces,” BioResources 1(1), 116-149.

Hubbe, M. A., Chen, H., and Heitmann, J. A. (2009). “Permeability reduction phenomena in packed beds, fiber mats, and wet webs of paper exposed to flow of liquids and suspensions: A review,” BioResources 4(1), 405-451.

Hubbe, M. A., Hasan, S. H., and Ducoste, J. J. (2011). “Cellulosic substrates for removal of pollutants from aqueous systems: A review. 1. Metals,” BioResources 6(2), 2161-2287.

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.

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., 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.

Hubbe, M. A., Rojas, O. J., Lucia, L. A., and Sain, M. (2008). “Cellulosic nanocomposites: A review,” BioResources 3(3), 929-980.

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

Hult, E.-L., Larsson, P. T., and Iversen, T. (2001). “Cellulose fibril aggregation – An inherent property of kraft pulps,” Polymer 42(8), 3309-3314.

Ibrahim, M. M., and El-Zawawy, W. K. (2004). “Preparation, characterization, and properties of fast-responding bulk hydrogel and treated bleached banana pulp,” Polymers for Advanced Technologies 15(5), 275-283.

Ibrahim, S. M., El Salmawi, K. M., and Zahran, A. H. (2007). “Synthesis of crosslinked superabsorbent carboxymethyl cellulose/acrylamide hydrogels through electron-beam irradiation,” J. Appl. Polymer Sci. 104(3), 2003-2008.

Ingmanson, W. L. (1952). “An investigation of the mechanisms of water removal from pulp slurries,” Tappi 35(10), 439-448.

Ingmanson, W. L. (1953). “Filtration resistance of compressible materials,” Chem. Eng. Progr. 49(11), 577-584.

Jain, V., and Ni, Y. H. (2008). “Alkaline hydrolysis of methyl acrylate grafted cellulose fibers,” J. Appl. Polymer Sci. 107(6), 3885-3890.

Jain, V., Xiao, H. N., and Ni, Y. H. (2007). “Grafting of poly(methyl acrylate) onto sulfite pulp fibers and its effect on water absorbance,” J. Appl. Polymer Sci. 105(6), 3195-3203.

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

Jeong, B., and Gutpwska, A. (2002). “Lessons from nature: Stimuli-responsive polymers and their biomedical applications,” Trends in Biotechnology 20(7), 305-311.

Jerkeman, P. (1970). “Cellulose fibers as water-absorbent material,” Svensk Papperstidn. – Nordiska Cellulosa 73(2), 37-38.

Johnson, B. N. (1994). “Inventory of land management inputs for producing absorbent fiber for diapers: A comparison of cotton and softwood land management,” Forest Products Journal 44(6), 39-45.

Joshi, S. C., Liang, C. M., and Lam, Y. C. (2008). “Effect of solvent state and isothermal conditions on gelation of methylcellulose hydrogels,” J. Biomater. Sci.: Polymer Edn. 19(12), 1611-1623.

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.

Kabiri, K., Omidian, H., Zohuriaan-Mehr, M. J., and Doroudiani, S. (2011). “Superabsorbent hydrogel composites and nanocomposites: A review,” Polymer Compos. 32(2), 277-289.

Kabra, B. G., and Gehrke, S. H. (1994). “Rate-limiting steps for solvent sorption and desorption by microporous stimuli-sensitive absorbent gels,” In: Superabsorbent Polymers: Science and Technology, Buchholz, F. L., and Peppas, N. A. (eds.), ACS Symp. Ser., Vol. 573, 76-86.

Kabra, B., Gehrke, S., and Spontak, R. (1998). “Microporous responsive hydroxypropyl cellulose gels. 1. Synthesis and microstructure,” Macromol. 31(7), 2166-2173.

Karaaslan, A. M., Tshabalala, M. A., and Buschle-Diller, G. (2010). “Wood hemicellulose/chitosan-based semi-interpenetrating network hydrogels: Mechanical, swelling and controlled drug release properties,” BioResources 5(2), 1036-1054.

Karaaslan, M. A., Tshabalala, M. A., and Buschle-Diller, G. (2012). “Semi-interpenetrating polymer network hydrogels based on aspen hemicellulose and chitosan: Effect of crosslinking sequence on hydrogel properties,” J. Appl. Polymer Sci. 124(2), 1168-1177.

Karaaslan, M. A., Tshabalala, M. A., Yelle, D. J., and Buschle-Diller, G. (2011). “Nanoreinforced biocompatible hydrogels from wood hemicelluloses and cellulose whiskers,” Carbohydrate Polymers 86(1), 192-201.

Kazanskii, K. S., and Dubrovskii, S. A. (1992). “Chemistry and physics of agricultural hydrogels,” Advan. Polymer Sci. 104, 97-133.

Kettunen, M., Silvennoinen, R. J., Houbenov, N., Nykanen, A., Ruokolainen, J., Sainio, J., Pore, V., Kemell, M., Ankerfors, M., Lindstrom, T., Ritala, M., Ras, R. H. A., and Ikkala, O. (2011). “Photoswitchable superabsorbency based on nanocellulose aerogels,” Advan. Functional Mater. 21(3), 510-517.

Kim, B. S., and Mun, S. P. (2009). “Effect of Ce4+ pretreatment on swelling properties of cellulosic superabsorbents,” Polymers for Advanced Technologies 20(12), 899-906.

Kim, U., and Kuga, S. (2002). “Functionalization of cellulose by periodate oxidation,” Cell Commun. 56(1), 7-10.

Klemm, D., Kramer, F., Moritz, S., Lindström, T., Ankerfors, M., Gray, D., and Dorris, A. (2011). “Nanocelluloses: A new family of nature-based materials,” Angew. Chem. Intl. Ed.50(24), 5438-5466.

Klungness, J. H., and Caulfield, D. F. (1982). “Mechanisms affecting fiber bonding during drying and aging of pulps,” Tappi J. 65(12), 94-97.

Knudson, M. I. (1993). “Bentonite in paper: The rest of the story,” TAPPI Papermakers Conf., TAPPI Press, Atlanta, 141-150.

Kocherbitov, V., Ulvenlund, S., Kober, M., Jarring, K., and Arnebrant, T. (2008). “Hydration of microcrystalline cellulose and milled cellulose studied by sorption calorimetry,” J. Phys. Chem. B 112(12), 3728-3734.

Kondo, T., Koschella, A., Heublein, B., Klemm, D., and Heinze, T. (2008). “Hydrogen bond formation in regioselectively functionalized 3-mono-O-methyl cellulose,” Carbohydr. Res.343(15), 2600-2604.

Kono, H., and Fujita, S. (2012). “Biodegradable superabsorbent hydrogels derived from cellulose by esterification crosslinking with 1,2,3,4-butanetetracarboxylic dianhydride,” Carbohydr. Polymers 87(4), 2582-2588.

Kono, H., Fujita, S., and Oeda, I. (2013). “Comparative study of homogeneous solvents for the esterification crosslinking of cellulose with 1,2,3,4-butanetetracarboxylic dianhydride and water absorbency of the reaction products,” J. Appl. Polymer Sci. 127(1), 478-486.

Kono, H., and Zakimi, M. (2013). “Preparation, water absorbency, and enzyme degradability of novel chitin- and cellulose/chitin-based superabsorbent hydrogels,” J. Appl. Polymer Sci.128(1), 572-581.

Koroskenyi, B., and McCarthy, S. P. (2001). “Synthesis of acetylated konjac glucomannan and effect of degree of acetylation on water absorbency,” Biomacromol. 2(3), 824-826.

Kosemund, K., Schlatter, H., Ochsenhirt, J. L., Krause, E. L., Marsman, D. S., and Erasala, G. N. (2009). “Safety evaluation of superabsorbent baby diapers,” Regul. Toxicol. Pharmacol.53(2), 81-89.

Kozeny, J. (1927). “About capillaries conducting water in the earth,” Sitzber. Akd. Wiss. Wien, Math. Naurv. Kasse 136(Abl. 2a), 271-306.

Kulicke, W. M., and Nottelmann, H. (1989). “Structure and swelling of some synthetic, semisynthetic, and biopolymer hydrogels,” Adv. Chem. Ser. 223, 15-44.

Kumar, J., Balomajumder, C., and Mondal, P. (2011). “Application of agro-based biomasses for zinc removal from wastewater – A review,” Clean – Soil Air Water 39(7), 641-652.

Kuwabara, S., and Kubota, H. (1996). “Water-absorbing characteristics of acrylic acid-grafted carboxymethyl cellulose synthesized by photografting,” J. Appl. Polymer Sci. 60(11), 1965-1970.

Laftah, W. A., Hashim, S., and Ibrahim, A. N. (2011). “Polymer hydrogels: A review,” Polymer-Plastics Technol. Eng. 50(14), 1475-1486.

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.

Larsson, M., Stading, M., and Larsson, A. (2010). “High performance polysodium acrylate superabsorbents utilizing microfibrillated cellulose to augment gel properties,” Soft Mater. 8(3), 207-225.

Larsson, M., Zhou, Q., and Larsson, A. (2011). “Different types of microfibrillated cellulose as filler materials in polysodium acrylate superabsorbents,” Chinese J. Polymer Sci. 29(4), 407-413.

Leks-Stepien, J., and Khadzhynova, S. (2005). “Effect of wetting solution components on the rate of penetration into offset paper,” Przeglad Papierniczy 61(6), 337-340.

Lepoutre, P., Hui, S. H., and Robertson, A. A. (1973). “Water absorbency of hydrolyzed polyacrylonitrile-grafted cellulose fibers,” J. Appl. Polymer Sci. 17(10), 3143-3156.

Levi Strauss & Co. (2009). “A Product Life Cycle Approach to Sustainability,” Report, accessed 6/3/13, http://www.levistrauss.com/sites/default/files/librarydocument/2010/4/Product_Lifecyle_Assessment.pdf.

Levis, J. W., and Barlaz, M. A. (2011). “Is biodegradability a desirable attribute for discarded solid waste? Perspectives from a national landfill greenhouse gas inventory model,” Environmental Science & Technology 45(13), 5470-5476.

Li, D., Zhang, X. Y., Simon, G. P., and Wang, H. T. (2013). “Forward osmosis desalination using polymer hydrogels as a draw agent: Influence of draw agent, feel solution and membrane on process performance,” Water Res. 47(1), 209-215.

Li, H., Liu, H., Fu, S. Y., and Zhan, H. Y. (2011). “Surface hydrophobicity modification of cellulose fibers by layer-by-layer self-assembly of lignosulfonates,” BioResources 6(2), 1681-1695.

Li, Q., Ma, Z. H., Yue, Q. Y., Gao, B. Y., Li, W. H., and Xu, X. (2012). “Synthesis, characterization and swelling behavior of superabsorbent wheat straw graft copolymers,” Bioresour. Technol. 118, 204-209.

Li, S., and Dai, H. (2004). “The influences of inorganic salts on WRV of papermaking stock,” Proc. 2nd Intl. Symp. Techol. Pulping Papermaking Biotech. Fiber Plants (2nd ISTPPBFP), Nanjing Forestry Univ., Oct. 13-14, 434-437.

Li, Y., and Tanaka, T. (1990). “Kinetics of swelling and shrinkage of gels,” J. Chem. Phys. 92, 1365-1371.

Liang, R., Yuan, H. B., Xi, G. X., and Zhou, Q. X. (2009). “Synthesis of wheat straw-g– poly(acrylic acid) superabsorbent composites and release of urea from it,” Carbohydrate Polymers77(2), 181-187.

Liang, X. T., Huang, Z. Q., Zhang, Y. J., Hu, H. Y., and Liu, Z. J. (2013). “Synthesis and properties of novel superabsorbent hydrogels with mechanically activated sugarcane bagasse and acrylic acid,” Polymer Bull. 70(6), 1781-1794.

Lim, D. W., Whang, H. S., Yoon, K. J., and Ko, S. W. (2001). “Synthesis and absorbency of a superabsorbent from sodium starch sulfate-g-polyacrylonitrile,” J. Appl. Polymer Sci. 79(8), 1423-1430.

Lim, K. Y., Yoon, K. J., and Kim, B. C. (2003). “Highly absorbable lyocell fiber spun from celluloses/hydrolyzed starch-g-PAN solution in NMMO monohydrate,” Eur. Polymer J. 39(11), 2115-2120.

Lin, C. X., Zhan, H. Y., Liu, M. H., Fu, S. Y., and Huang, L. H. (2010). “Rapid homogeneous preparation of cellulose graft copolymer in BMIMCL under microwave irradiation,” J. Appl. Polymer Sci. 118(1), 399-404.

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.

Lionetto, F., Sannino, A., and Maffezzoli, A. (2005). “Ultrasonic monitoring of the network formation in superabsorbent cellulose based hydrogels,” Polymer 46(6), 1796-1803.

Lionetto, F., Sannino, A., Mensitieri, G., and Maffezzoli, A. (2003). “Evaluation of the degree of cross-linking of cellulose-based superabsorbent hydrogels: A comparison between different techniques,” Macromol. Symp. 200, 199-207.

Liu, J., Li, Q., Su, Y., Yue, Q. Y., Gao, B. Y., and Wang, R. (2013). “Synthesis of wheat straw cellulose-g-poly (potassium acrylate)/PVA semi-IPNs superabsorbent resin,” Carbohydr. Polym. 94(1), 539-546.

Liu, P., Peng, J., Li, J., and Wu, J. (2005a). “Radiation crosslinking of CMC-Na at low dose and its application as substitute for hydrogel,” Radiat. Phys. Chem. 72(5), 635-638.

Liu, P., Zhai, M., Li, J., Peng, J., and Wu, J. (2002). “Radiation preparation and swelling behavior of sodium carboxymethylcellulose hydrogels,” Radiat. Phys. Chem. 63(3-6), 525-528.

Liu, Q., Wang, X., and Kee, D. D. (2005b). “Mass transport through swelling membranes,” Intl. J. Eng. Sci. 43, 1464-1470.

Liu, Q. X., Xu, W. C., Lv, Y. B., and Li, J. L. (2011). “Preparation and properties studies of cellulose super absorbent polymer,” Advanced Textile Materials, Book Ser.: Advanced Materials Research, Vol 332-334, 1816-1819.

Liu, W., Zhang, B., Lu, W., Li, X., Zhu, D., Yao, K., Wang, Q., Zhao, C., and Wang, C. (2004). “A rapid temperature-responsive sol–gel reversible poly(N-isopropylacryl-amide)-g-methylcellulose copolymer hydrogel,” Biomater. 25(15), 3005-3012.

Liu, Z. X., Miao, Y. G., Wang, Z. Y., and Yin, G. H. (2009). “Synthesis and characterization of a novel super-absorbent based on chemically modified pulverized wheat straw and acrylic acid,” Carbohydrate Polymer 77(1), 131-135.

Loh, W. (2002). “Block copolymer micelles,” Encyclopedia of Surface and Colloid Science, Markcel Dekker, New York.

Lokhande, H. T., and Gotmare, V. D. (1999). “Utilization of textile loom waste as a highly absorbent polymer through graft co-polymerization,” Bioresour. Technol. 68(3), 283-286.

Lokhande, H. T., and Varadarajan, P. V. (1993). “A new approach in the production of non-wood-based cellulosic superabsorbents through the PAN grafting method,” Bioresour. Technol.45(3), 161-165.

Loria-Bastarrachea, M. I., Carrillo-Escalante, H. J., and Aguilar-Vega, M. J. (2002). “Grafting of poly(acrylic acid) onto cellulosic microfibers and continuous cellulose filaments and characterization,” J. Appl. Polymer Sci. 83(2), 386-393.

Lowe, R. M., Page, D. H., Waterhouse, J. F., Hsieh, J., Cheluka, N., and Ragauskas, A. J. (2007). “Deformation behavior of wet lignocellulosic fibers,” Holzforschung 61(3), 261-266.

Lu, G. D., Yan, Q. Z., Su, X. T., Liu, Z. Q., and Ge, C. C. (2007). “Progress of porous hydrogels,” Prog. Chem. 19(4), 485-493.

Lucas, R. (1918). “Rate of capillary ascension of liquids,” Kolloid Z. 23, 15-22.

Lund, K., de al Motte, H., Brelid, H., Westman, G., and Hanson, C. (2011). “Diepoxide treatment of softwood kraft pulp: Influence on absorption properties of fibre networks,” Cellulose18(5), 1365-1375.

Lund, K., Sjöström, K., and Brelid, H. (2012). “Alkali extraction of kraft pulp fibers: Influence on fibre and fluff pulp properties,” J. Eng. Fiber Fabr. 7(2), 30-39.

Lyne, M. B., and Aspler, J. S. (1982). “Wetting and the sorption of water by paper under dynamic conditions,” Tappi J. 65(12), 98-101.

Ma, J., Xu, Y., Zhang, Q., Zha, L., and Liang, B. (2007). “Preparation and characterization of pH- and temperature-responsive semi-IPN hydrogels of carboxymethyl chitosan with poly-(N-isopropyl acrylamide) crosslinking by clay,” Colloid Polym. Sci. 28, 479-484.

Macaulay, M., Clarke-ONeill, S., Fader, M., Pettersson, L., and Cottenden, A. (2004). “Are washable absorbents effective at containing urinary incontinence?,” Nursing Times 100(12), 58-62.

Maloney, T. C., Paulapuro, H., and Stenius, H. (1998). “Hydration and swelling of pulp fibers measured with differential scanning calorimetry,” Nordic Pulp Paper Res. J. 13(1), 31-36.

Mann, R. L., and Leah, P. J. (1987). “Fluff pulp and debonding agents,” Paper Southern Africa 7(6), 41-42.

Marci, G., Mele, G., Palmisano, L., Pulito, P., and Sannino, A. (2006). “Environmentally sustainable production of cellulose-based superabsorbent hydrogels,” Green Chem. 8(5), 439-444.

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.

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.

Matushek, M. (2008). The History of Super Absorbent Chemistry, M2 Polymer Technologies, Inc.

Meena, R., Lehnen, R., Schmitt, U., and Saake, B. (2011). “Effect of oat spelt and beech xylan on the gelling properties of kappa-carrageenan hydrogels,” Carbohydrate Polymers 85(3), 529-540.

Michaels, A.S. (1965). “Polyelectrolyte complexes,” Indust. Eng. Chem. 57(10), 32-40.

Michaels, A. S., Mir, L., and Schneider, N. S. (1965). “A conductometric study of polycation-polyanion reactions in dilute aqueous solution,” J. Phys. Chem. 69(5), 1447-1455.

Mirabella, N., Castellani, V., and Sala, S. (2013). “Life cycle assessment of bio-based products: A disposable diaper case study,” The International Journal of Life Cycle Assessment 18, 1036-1047.

Miyamoto, T., Long, M., and Donkai, N. (1995). “Preparation of new types of temperature-responsive cellulose derivatives,” Macromol. Symp. 99, 141-147.

Mizutani, C., Inagaki, H., and Bertoniere, N. R. (1999). “Water absorbancy of never-dried cotton fibers,” Cellulose 6(2), 167-176.

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.

Myung, D., Waters, D., Wiseman, M., Duhamel, P.-E., Noolandi, J., Ta, C. N., and Frank, C. W. (2008). “Progress in the development of interpenetrating polymer network hydrogels,” Polymers. Advan. Technol. 19(6), 647-657.

Nagorski, H. (1994). “Characterization of a new superabsorbent polymer generation,” In: Superabsorbent Polymers: Science and Technology, Buchholz, F. L., and Peppas, N. A. (eds.), ACS Symposium Series, Vol. 573, pp. 99-111.

Neumann, P., Blizzard, L., Grimmer, K., and Grant, R. (2004). “Expanded paper towel test: An objective test of urine loss for stress incontinence,” Neurourology Urodynamics 23(7), 649-655.

Nishio, Y., and Manley, R. (1988). “Cellulose-poly(vinyl alcohol) blends prepared from solutions in N,N-dimethylacetamide-lithium chloride,” Macromol. 21(5), 1270-1277.

Obataya, E., Shibutani, S., and Minato, K. (2007). “Swelling of acetylated wood II: Effects of delignification on solvent adsorption of acetylated wood,” J. Wood Sci. 53(5), 408-411.

O’Brien, K., Olive, R., Hsu, Y. C., Morris, L., Bell, R., and Kendall, N. (2009). “Life cycle assessment: Reusable and disposable nappies in Australia,” ALCAS–Australian Life Cycle Assessment Society, downloaded July 2013, s3.amazonaws.com/zanran_storage/www.crdc.com.au/…/44777470.pdf‎

Olive, R. (2004). “A life cycle analysis of disposable, reusable and compostable nappies under Brisbane conditions,” Dept. of Chemical Engineering, University of Queensland.

Oliveira, V. A., Veloso, T. C., Leao, V. A., dos Santos, C. G., and Botaro, V. R. (2013).
“Hydrogels of cellulose acetate crosslinked with pyromellitic dianhydride – Part 1: Synthesis and swelling kinetics,” Quimica Nova 36(1), 102-106.

Omidian, H., Hashemi, S. A., Sammes, P. G., and Meldrum, I. G. (1998). “A model for the swelling of superabsorbent polymers,” Polymer 39, 6697-6704.

Omidian, H., Rocca, J. G., and Park, K. (2005). “Advances in superporous hydrogels,” J. Controlled Release 102(1), 3-12.

Oppermann, W. (1995). “Superabsorbent materials based on cellulose,” Papier 49(12), 765-769.

Pääkkö, M., Vapaavuori, J., Silvennoinen, R., Kosonen, H., Ankerfors, M., Lindström, T., Berglund, L. A., and Ikkala, O. (2008). “Long and entangled native cellulose I nanofibers allow flexible aerogels and hierarchically porous templates for functionalities,” Soft Matter 4, 2492-2499.

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

Pan, S. B., and Ragauskas, A. J. (2012). “Preparation of superabsorbent cellulosic hydrogels,” Carbohydr. Polymers 87(2), 1410-1418.

Pan, Y.-L., Kuga, S., and Usuda, M. (1988). “Ultrasonic technique to study wetting and liquid penetration of paper,” TAPPI J. 71(5), 119-123.

Pandey, J. K., Nakagaito, A. N., and Takagi, H. (2013). “Fabrication and applications of cellulose nanoparticle-based polymer composites,” Polymer Eng. Sci. 53(1), 1-8.

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. (2006). “Changes in pore size distribution during the drying of cellulose fibers as measured by differential scanning calorimetry,” Carbohydr. Polym. 66(1), 97-103.

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.

Paula, H. C. B., Gomes, F. J. S., and de Paula, R. C. M. (2002). “Swelling studies of chitosan/cashew nut gum physical gels,” Carbohydr. Polym. 48, 313-318.

Payne, K. C., Jackson, C. D., Aizpurua, C. E., Rojas, O. J., and Hubbe, M. A. (2012). “Oil spill abatement: Factors affecting oil uptake by cellulosic fibers,” Environ. Sci. Technol. 46(14), 7725-2230.

Peng, X. W., Ren, J. L., Zhong, L. X., Peng, F., and Sun, R. C. (2011). “Xylan-rich hemicelluloses-graft-acrylic acid ionic hydrogels with rapid responses to pH, salt, and organic solvents,” J. Agricul. Food Chem. 59(15), 8208-8215.

Peppas, N. A., and Hariharan, D. (1994). “Dynamic swelling of ionic networks. Application to behavior of dry superabsorbents,” In: Superabsorbent Polymers Science and Technology, Buchholz, F. L., and Peppas, N. A. (eds.), ACS Symp. Ser. 573, Amer. Chem. Soc., Washington, DC.

Peppas, N. A., and Sahlin, J. J. (1989). “A simple equation for the description of solute release. III. Coupling of diffusion and relaxation,” Int. J. Pharm. 57, 169-172.

Peura, M., Karppinen, T., Soovre, A., Salmi, A., Tenkanen, M., Haeggstrom, E., and Serimaa, R. (2008). “Crystallization and shear modulus of a forming biopolymer film determined by in situ x-ray diffraction and ultrasound reflection methods,” J. Appl. Physics 104(2), article no. 023513; DOI: 10.1063/1.2951451.

Poffenberger, C., Deac, Y., and Zeman, W. (2000). “Novel hydrophilic softeners for tissue and towel applications,” TAPPI Papermakers Conf. & Trade Fair, TAPPI Press, Atlanta, 85-93.

Pohjanlehto, H., Setala, H., Kammiovirta, K., and Harlin, A. (2011). “The use of N,N ‘-diallylaldardiamides as cross-linkers in xylan derivatives-based hydrogels,” Carbohydrate Res.346(17), 2736-2745.

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

Pourjavadi, A., Amini-Fazl, M. S., and Ayyari, M. (2007a). “Optimization of synthetic conditions CMC-g-poly (acrylic acid)/celite composite superabsorbent by Taguchi method and determination of its absorbency under load,” Expr. Polymer Lett. 1(8), 488-494.

Pourjavadi, A., and Ghasemzadeh, H. (2006). “CMC-g-poly (sodium acrylate)/kaolin superabsorbent hydrogel composites: Synthesis, characterization and swelling behavior,” Polymers Polymer Compos. 14(7), 701-712.

Pourjavadi, A., Ghasemzadeh, H., and Mojahedi, F. (2009). “Swelling properties of CMC-g-poly (AAm-co-AMPS) superabsorbent hydrogel,” J. Appl. Polymer Sci. 113(6), 3442-3449.

Pourjavadi, A., and Mahdavinia, G. R. (2005). “Superabsorbency and swelling behaviour of partially hydrolyzed carboxymethylcellulose-g-PAAm hydrogels,” J. Polymer Mater. 22(3), 235-243.

Pourjavadi, A., Seidi, F., Salimi, H., and Soleyman, R. (2008.). “Grafted CMC/silica gel superabsorbent composite: Synthesis and investigation of swelling behavior in various media,” J. Appl. Polymer Sci. 108(5), 3281-3290.

Pourjavadi, A., Soleyman, R., Ghasemzadeh, H., and Salimi, H. (2010). “CMC/celite superabsorbent composites: Effect of reaction variables on saline-absorbency under load,” Iranian Polymer J. 19(8), 571-579.

Pourjavadi, A., Zohuriaan-Mehr, M. J., Ghasernpoori, S. N., and Hossienzadeh, H. (2007b). “Modified CMC. V. Synthesis and super-swelling behavior of hydrolyzed CMC-g-PAN hydrogel,” J. Appl. Polymer Sci. 103(2), 877-883.

Pramanick, A. K., and Crouse, B. W. (1998). “Application of deep grooved polyester fiber in composite high absorbent paper,” TAPPI Nonwovens Conf. Trade Fair, Petersburg, FL, 211-228.

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

Quddus, M. A. A. R., Rojas, O. J., and Pasquinelli, M. A. (submitted). “Molecular dynamics simulations of the thermal stability of crystalline cellulose surfaces coated with oleic acid,” Functional Materials from Renewable ResourcesACS Symp. Ser. 1107, 191-208; also see Quddus, M. “A systematic study of fatty acids interacting with crystalline cellulose surfaces via molecular dynamic simulations,” Doctoral dissertation, NC State Univ., Depts. of Fiber and Polymer Sci. and Forest Biomaterials, 2011.

Raafat, A. I., Eid, M., and El-Arnaouty, M. B. (2012). “Radiation synthesis of superabsorbent CMC based hydrogels for agriculture applications,” Nucl. Intrum. Methods Phys. Res. Sec. B – Beam Interac. Mater. Atoms 283, 71-76.

Rajagopa, K. R. (2003). “Diffusion through polymeric solids undergoing large deformation,” Mater. Sci. Technol. 19, 1175-1180.

Rezai, E., and Warner, R. R. (1997). “Polymer-grafted cellulose fibers. 1. Enhanced water absorbency and tensile strength,” J. Appl. Polymer Sci. 65(8), 1463-1469.

Rodriguez, R., Alvarez-Lorenzo, C., and Concheiro, A. (2003). “Cationic cellulose hydrogels: Kinetics of the cross-linking process and characterization as pH/ion-selective drug delivery systems,” J. Control Release 86(2-3), 253-265.

Roy, N., Saha, N., Kitano, T., and Saha, P. (2010). “Novel hydrogels of PVP-CMC and their swelling effect on viscoelastic properties,” J. Appl. Polymer Sci. 117(3), 1703-1710.

Rudie, A. W., Ball, A., and Patel, N. (2006). “Ion exchange of H+, Na+, Mg2+, Ca2+, Mn2+, and Ba2+ on wood pulp,” J. Wood Chem. Technol. 26(3), 259-272.

Sahoo, P. K., Mohapatra, R., Sahoo, A., DebSarkar, N., and Swain, S. K. (2005). “Characterization, biodegradation, and water absorbency of chemically modified tossa variety jute fiber via pulping and grafting with acrylamide,” Intl. J. Polymer Anal. Charact. 10(3-4), 153-167.

Saito, H., Sakaurai, A., Sakakibara, M., and Saga, H. (2003). “Preparation and properties of transparent hydrogels,” J. Appl. Polymer Sci. 90(11), 3020-3025.

Saito, N., Aoyama, M., Shimizu, Y., Takai, M., and Hayashi, J. (1995). “Super absorbent materials prepared from lignocellulosic materials by phosphorylation. 7. Characteristics of water absorbencies and alkaline treatments,” Mokuzai Gakkaishi 41(2), 179-185.

Saito, N., Shimizu, Y., Takai, M., and Hayashi, J. (1994a). “Super absorbent materials prepared from lignocellulosic materials by phosphorylation. 4. Fine-structure and water absorbency,” Mokuzai Gakkaishi 40(9), 937-942.

Saito, N., Shimizu, Y., Takai, M., and Hayashi, J. (1994b). “Super absorbent materials prepared from lignocellulosic materials by phosphorylation. 5. Crystalline-structure and water absorbency,” Mokuzai Gakkaishi 40(11), 1200-1207.

Salam, A., Pawlak, J. J., Venditti, R. A., and El-Tahlawy, K. (2011a). “Incorporation of carboxyl groups into xylan for improved absorbency,” Cellulose 18(4), 1033-1041.

Salam, A., Venditti, R. A., Pawlak, J. J., and El-Tahlawy, K. (2011b). “Crosslinked hemicellulose citrate-chitosan aerogel foams,” Carbohyd. Polymers 84(4), 1221-1229.

Salimi, H., Pourjavadi, A., Seidi, F., Jahromi, P. E., and Soleyman, R. (2010). “New smart carrageenan-based superabsorbent hydrogel hybrid: Investigation of swelling rate and environmental responsiveness,” J. Appl. Polymer Sci. 117(6), 3228-3238.

Sammon, C., Bajwa, G., Timmins, P., and Melia, C. D. (2006). “The application of attenuated total reflection Fourier transform infrared spectroscopy to monitor the concentration and state of water in solution of a thermally responsive cellulose ether during gelation,” Polymer 47(2), 577-584.

Sannino, A., Demitri, C., and Madaghiele, M. (2009). “Biodegradable cellulose-based hydrogels: Design and applications,” Materials 2(2), 353-373.

Sannino, A., Madaghiele, M., Conversano, F., Mele, G., Maffezzoli, A., Netti, P. A., Ambrosio, L., and Nicolais, L. (2004a). “Cellulose derivative-hyaluronic acid-based microporous hydrogels cross-linked through divinyl sulfone (DVS) to modulate equilibrium sorption capacity and network stability,” Biomacromol. 5(1), 92-96.

Sannino, A., Maffezzoli, A., and Nicolais, L. (2003). “Introduction of molecular spacers between the crosslinks of a cellulose-based superabsorbent hydrogel: Effects on the equilibrium sorption properties,” J. Appl. Polymer Sci. 90(1), 168-174.

Sannino, A., Mensitieri, G., and Nicolais, L. (2004b). “Water and synthetic urine sorption capacity of cellulose-based hydrogels under a compressive stress field,” J. Appl. Polymer Sci.91(6), 3791-3796.

Sannino, A., and Nicolais, L. (2005). “Concurrent effect of microporosity and chemical structure on the equilibrium sorption properties of cellulose-based hydrogels,” Polymer 46(13), 4676-4685.

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 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.

Schexnailder, P., and Schmidt, G. (2009). “Nanocomposite polymer hydrogels,” Colloid Polymer Sci. 287(1), 1-11.

Schmaljohann, D. (2006). “Thermo- and pH-responsive polymers in drug delivery,” Advanced Drug Delivery Reviews 58, 1655-1670.

Schuchardt, D. R., and Berg, J. C. (1991). “Liquid transport in composite cellulose-superabsorbent fiber networks,” Wood Fiber Sci. 23(3), 342-357.

Schuster, K. C., Suchomel, F., Manner, J., Abu-Rous, M., and Firgo, H. (2006). “Functional and comfort properties of textiles from TENCEL ® fibres resulting from the fibres’ water-absorbing nanostructure: A review,” Macromol. Symp. 244, 149-165.

Sehaqui, H., Salajková, M., Zhou, Q., and Berglund, L. (2010). “Mechanical performance tailoring of tough ultra-high porosity foams prepared from cellulose I nanofiber suspensions,” Soft Matter 6, 1824-1832.

Sehaqui, H., Zhou, Q., and Berglund, L. (2011). “High porosity aerogels of high specific surface area prepared from nanofibrillated cellulose (NFC),” Compos. Sci. Technol. 71(13), 1593-1599.

Sen, S. K., Baheti, V. K., Venditti, R. A., Pawlak, J. J., Park, S., and Bansal, M. C. (2012). “Cellulose microfibril-water interaction as characterized by isothermal thermogravimetric analysis and scanning electron microscopy,” BioResources 7(4), 4683-4703.

Shang, J., Shao, Z., and Chen, X. (2008). “Electrical behavior of a natural polyelectrolyte hydrogel: Chitosan/carboxymethylcellulose hydrogel,” Biomacromol. 9(4), 1208-1213.

Shet, R. T., and Wallajapet, P. R. R. (1997). “Manufacture of sulfonated cellulose with improved absorbent properties,” US Pat. 5703225.

Shepherd, I., and Xiao, H. N. (1999). “The role of surfactants as rewetting agents in enhancing paper absorbancy,” Colloids Surf. A – Physicochem. Eng. Aspects 157, 235-244.

Shi, S. N., Wang, W. B., and Wang, A. Q. (2013). “pH-responsive sodium alginate-based superporous hydrogel generated by an anionic surfactant micelle templating,” Carbohydr. Polym. 94(1), 449-455.

Shu, J., Li, X. J., and Zhao, D. B. (2011). “Microwave-irradiated preparation of super absorbent resin by graft copolymerization of cellulose and acrylic acid/acrylamide,” Manufacturing Processes and Systems, Book Ser.: Advanced Materials Research, Univ Wollongong, Vols. 148-149, pp. 799-802.

Silva, S. M. C., Pinto, F. V., Antunes, F. E., Miguel, M. G., Sousa, J. J. S., and Pais, A. A. C. C. (2008). “Aggregation and gelation in hydroxypropylmethyl cellulose aqueous solutions,” J. Colloid Interface Sci. 327(2), 333-340.

Simoncic, B., and Roman, V. (2007). “Wettability of cotton fabric by aqueous solutions of surfactants with different structures,” Colloids Surf. A.- Physicochem. Eng. Aspects 292(2-3), 236-245.

Singh, J., and Weber, M. E. (1996). “Kinetics of one-dimensional gel swelling and collapse for large volume change,” Chem. Eng. Sci. 51, 4499-4508.

Singha, A. S., and Rana, A. K. (2011). “Kinetics of graft copolymerization of acrylic acid onto Cannabis indica fibre,” Iranian Polymer J. 20(11), 913-929.

Sipponen, M. H., Pastinen, O. A., Strengell, R., Hyotylainen, J. M. I., Heiskanen, I. T., and Laakso, S. (2010). “Increased water resistance of CTMP fibers by oat (Avena sativa L.) husk lignin,” Biomacromol. 11(12), 3511-3518.

Smith, P. M., and Sheeran, K. D. (1992). “A profile of consumer preferences for baby diapers,” Forest Prod. J. 42(9), 65-71.

Smook, G. A. (1992). Handbook for Pulp & Paper Technologists, Second Ed., Angus Wilde PublicationsVancouver.

Spagnol, C., Rodrigues, F. H. A., Pereira, A. G. B., Fajardo, A. R., Rubira, A. F., and Muniz, E. C. (2012a). “Superabsorbent hydrogel nanocomposites based on starch-g-poly(sodium acrylate) matrix filled with cellulose nanowhiskers,” Cellulose 19(4), 1225-1237.

Spagnol, C., Rodrigues, F. H. A., Pereira, A. G. B., Fajardoa, A. R., Rubira, A. F., and Muniz, E. C. (2012b). “Superabsorbent hydrogel composite made of cellulose nanofibrils and chitosan-graft-poly(acrylic acid),” Carbohydr. Polym. 87, 2038-2045.

Spagnol, C., Rodrigues, F. H. A., Neto, A. G. V. C., Pereira, A. G. B., Fajardo, A. R., Radovanovic, E., Rubira, A. F., and Muniz, E. C. (2012c). “Nanocomposites based on poly(acrylamide-co-acrylate) and cellulose nanowhiskers,” Eur. Polymer J. 48, 454-463.

Spence, K. L., Venditti, R. A., Rojas, O. J., Habibi, Y., and Pawlak, J. J. (2010). “The effect of chemical composition on microfibrillar cellulose films from wood pulps: Water interactions and physical properties for packaging applications,” Cellulose 17, 835-848.

Steadman, R., and Luner, P. (1985). “The effect of wet fiber flexibility on sheet apparent density,” In: Papermaking Raw Materials, Trans. 8th Fundamental Res. Symp., Oxford, UK, Sept. 15-20, Mechanical Engineering Publ. Ltd., London, 311-358.

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 Symposium Held at Cambridge, Sept. 1965, Bolam, F. (ed.), Tech. Sec. British Paper & Board Makers’ Assoc., London, Vol 1., 145-174.

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 Carlsson, G. (1992). “Wettability of kraft pulps – Effect of surface composition and oxygen plasma treatment,” J. Adhes. Sci. Technol. 6(6), 745-761.

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.

Sun, T., and Lindsay, J. D. (2003). “Curly cellulose fiber for diapers, having preset wet curl value is obtained by applying polymeric reactive compound to cellulose fiber having preset curl value which is treated with intra-crystalline swelling agent,” US Pat. 2003108739-A1.

Suo, A. L., Qian, J. M., Yao, Y., and Zhang, W. G. (2007). “Synthesis and properties of carboxymethyl cellulose-graft-poly(acrylic acid-co-acrylamide) as a novel cellulose-based superabsorbent,” J. Appl. Polymer Sci. 103(3), 1382-1388.

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.

Tag, C. M., Toiviainen, M., Juuti, M., Rosenholm, J. B., Backfolk, K., and Gane, P. A. C. (2012). “The effect of isopropyl alcohol and non-ionic surfactant mixtures on the wetting of porous coated paper,” Transport in Porous Media 94(1), 225-242.

Tanaka, T, and Fillmore, D. J. (1979). “Kinetics of swelling of gels,” J. Chem. Phys. 70, 1214-1218.

Tanaka, Y., Gong, J. P., and Osada, Y. (2005). “Novel hydrogels with excellent mechanical performance,” Prog. Polymer Sci. 30(1), 1-9.

Tang, Y., Wang, X., Li, Y., Lei, M., Du, Y., Kennedy, J. F., and Knill, C. J. (2010). “Production and characterization of novel injectable chitosan/methylcellulose/salt blend hydrogels with potential application as tissue engineering scaffolds,” Carbohydrate Polymers 82(3), 833-841.

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

TAPPI (1996). TAPPI Standard Methods, TAPPI Press, Atlanta.

Theliander, H., Wernersson, F., Hansson, C., Gustafson, I., and Falk, T. (2011). “Absorbent article comprising absorbent porous foam of freeze-dried microfibrillated cellulose,” WO Patent2011090410, 28 July 2011.

Thuresson, K., and Lindman, B. (1999). “Association in nonionic cellulose ether solutions due to microcrystallites,” Colloids Surf. A – Physicochem. Eng. Aspects 159(1), 219-226.

Uner, B. (2009). “The effect of support media on corona treated paper sorption properties,” Scient. Res. Essays 4(10), 1024-1030.

Valencia, J., and Pierola, I. F. (2007). “Interpretation of the polyelectrolyte and antipolyelectrolyte effects of poly (N-vinylimidazole-co-sodium styrenesulfonate) hydrogels,” J. Polymer Sci. Part B – Polymer Phys. 45(13), 1683-1693.

Varaprasad, K., Vimala, K., Ravindra, S., Reddy, N. N., and Raju, K. M. (2011). “Development of sodium carboxymethyl cellulose-based poly(acrylamide-co-2-acrylamido-2-methyl-1-propane sulfonic acid) hydrogels for in vitro drug release studies of ranitidine hydrochloride an anti-ulcer drug,” Polym.-Plast. Technol. Eng. 50(12), 1199-1207.

Vitta, S. B., Stahel, E. P., and Stannet, V. T. (1986). “The preparation and properties of acrylic and methacrylic acid grafted cellulose prepared by ceric ion initiation. II. Water retention properties,” J. Appl. Polym. Sci. 32(7), 5799-5810.

Wach, R. A., Mitomo, H., Yoshii, F., and Kume, T. (2001). “Hydrogel of biodegradable cellulose derivatives. II. Effect of some factors on radiation-induced crosslinking of CMC,” J. Appl. Polymer Sci. 81(12), 3030-3037.

Wach, R. A., Mitomo, H., Yoshii, F., and Kume, T. (2002). “Hydrogel of radiation-induced cross-linked hydroxypropylcellulose,” Macromol. Mater. Eng. 287(4), 285-295.

Wang, C. G., He, L. X., Dong, X. G., Gao, X. P., Liu, W., Dong, X. C., Yuan, T., and Zhang, J. (2007). “Preparation and properties of superabsorbent by graft co-polymerization of acrylic-acid onto different celluloses of crops stems,” Chem. J. Chinese Univ. – Chin. 28(9), 1787-1790.

Wang, D., Song, Z. Q., and Shang, S. B. (2008). “Characterization and biodegradability of amphoteric superabsorbent polymers,” J. Appl. Polymer Sci. 107(6), 4116-4120.

Wang, J. L., Wang, W. B., and Wang, A. Q. (2010). “Synthesis, characterization and swelling behaviors of hydroxyethyl cellulose-g-poly(acrylic acid)/attapulgite superabsorbent composite,” Polymer Eng. Sci. 50(5), 1019-1027.

Wang, J. L., Wang, W. B., Zheng, Y. A., and Wang, A. Q. (2011a). “Effects of modified vermiculite on the synthesis and swelling behaviors of hydroxyethyl cellulose-g-poly(acrylic acid)/vermiculite superabsorbent nanocomposites,” J. Polymer Res. 18(3), 401-408.

Wang, J., Xu, L., Li, C. C., Yue, Z. Y., Zhai, M. L., and Li, J. Q. (2010). “Antipolyelectrolyte swelling of amphiphilic hydroxypropyl methylcellulose phthalate gels,” Colloids Surf. A – Physicochem. Eng. Aspects 356, 89-96.

Wang, W. B., and Wang, A. Q. (2010). “Nanocomposite of carboxymethyl cellulose and attapulgite as a novel pH-sensitive superabsorbent: Synthesis, characterization and properties,” Carbohydrate Polymers 82(1), 83-91.

Wang, W. B., Wang, J., Kang, Y. R., and Wang, A. Q. (2011b). “Synthesis, swelling and responsive properties of a new composite hydrogel based on hydroxyethyl cellulose and medicinal stone,” Composites. Pt. B – Eng. 42(4), 809-818.

Wang, W. B., Wang, Q., and Wang, A. Q. (2011c). “pH-responsive carboxymethylcellulose-g-poly(sodium acrylate)/polyvinylpyrrolidone semi-IPN hydrogels with enhanced responsive and swelling properties,” Macromol. Res. 19(1), 57-65.

Wang, W. B., Xu, J. X., and Wang, A. Q. (2011d). “A pH-, salt- and solvent-responsive carboxymethylcellulose-g-poly(sodium acrylate)/medical stone superabsorbent composite with enhanced swelling and responsive properties,” Express Polymer Lett. 5(5), 385-400.

Wang, Y. Z., Shi, X. N., Wang, W. B. and Wang, A. Q. (2013). “Synthesis, characterization, and swelling behaviors of a pH-responsive CMC-g-poly(AA-co-AMPS) superabsorbent hydrogel,” Turk. J. Chem. 37(1), 149-159.

Warner, R. R., and Rezai, E. (1997). “Polymer-grafted cellulose fibers. 2. Polymer localization and induced alterations in fiber morphology,” J. Appl. Polymer Sci. 65(8), 1471-1485.

Washburn, E. V. (1921). “The dynamics of capillary flow,” Phys. Rev. 17, 273-283.

Weerawarna, S. A., and Su, B. (2010). “Methods for the preparation of mixed polymer superabsorbent fibers containing cellulose,” US Patent 7,717,995 B2.

Weisbrod, A. V., and Van Hoof, G. (2012). “LCA-measured environmental improvements in PampersA (R) diapers,” Intl. J. Life Cycle Assess. 17(2), 145-153.

Weise, U., and Paulapuro, H. (1996). “The relationship between fiber shrinkage and hornification,” Papier 50(6), 328-333; reprinted in English in 1998 in: Prog. Paper Recycling 7(3), 14-21.

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

Welo, L. A., Ziifle, H. M., and Loeb, L. (1952). “Swelling capacities of fibers in water Part I: Desiccation rate measurements,” Textile Res. J. 22, 254-261.

Wichterle, O., and Lim, D. (1960). “Hydrophilic gels for biological use,” Nature 185, 117-118.

Williams, J. L., and Stannet, V. T. (1979). “Highly water absorptive cellulose by postdecrystallization,” J. Appl. Polym. Sci. 23(4), 1265-1268.

Wiryana, S., and Berg, J. C. (1991). “The transport of water in wet-formed networks of cellulose fibers and powdered superabsorbent,” Wood Fiber Sci. 23(3), 457-464.

Wu, F., Zhang, Y., Liu, L., and Yao, J. M. (2012). “Synthesis and characterization of a novel cellulose-g-poly(acrylic acid-co-acrylamide) superabsorbent composite based on flax yarn waste,” Carbohydrate Polymers 87(4), 2519-2525.

Wu, W. J., Pan, Z. G., and Peng, J. P. (2008). “Preparation of superabsorbent polymer from softwood pulp by grafting-copolymerization,” Proc. Intl. Conf. Pulping, Papermaking Biotechnol. 2008: ICPPB ’08, Jin, Y., Zhai, H., and Li, Z. (eds.), Vol. 2, pp. 529-532.

Xiao, C. M., and Gao, Y. K. (2008). “Preparation and properties of physically crosslinked sodium carboxymethylcellulose/poly(vinyl alcohol) complex hydrogels,” J. Appl. Polymer Sci.107(3), 1568-1572.

Xiao, Y. C., and Meshitsuka, G. (2001). “Development of high-retention water absorbent from cellulosic materials: Water absorbent from bleached kraft pulp,” J. Wood Sci. 47(5), 394-399.

Xie, L. H., Liu, M. Z., Ni, B. L., and Wang, Y. F. (2012). “Utilization of wheat straw for the preparation of coated controlled-release fertilizer with the function of water retention,” J. Agric. Food Chem. 60(28), 6921-6928.

Xu, G. G., and Yang, C. Q. X. (1999). “Comparison of the kraft paper crosslinked by polymeric carboxylic acids of large and small molecular sizes: Dry and wet performance,” J. Appl. Polymer Sci. 74(4), 907-912.

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,” Polymer J.38(8), 819-826.

Yanai, Y., Hamada, K., and Shimizu, Y. (2006). “The liquid ammonia treatment of cotton fibers – Comparison and combination with mercerization using a practical unit,” Sen-I Gakkaishi62(4), 81-88.

Yang, C. Q., and Xu, Y. F. (1998). “Paper wet performance and ester crosslinking of wood pulp cellulose by poly(carboxylic acids),” J. Appl. Polymer Sci. 67(4), 649-658.

Yang, F., Li, G., He, Y. G., Ren, F. X., and Wang, G. X. (2009). “Synthesis, characterization, and applied properties of carboxymethyl cellulose and polyacrylamide graft copolymer,” Carbohydrate Polymers 78(1), 95-99.

Yang, J. Y., Zhou, X. S., and Fang, J. (2011). “Synthesis and characterization of temperature sensitive hemicellulose-based hydrogels,” Carbohydrate Polymers 86(3), 1113-1117.

Yin, Y., Yang, Y., and Xu, H. (2002). “Swelling behavior of hydrogels for colon-site drug delivery,” J. Appl. Polymer Sci. 83, 2835-2842.

Yoshimura, T., Matsunaga, M., and Fujioka, R. (2009). “Alginate-based superabsorbent hydrogels composed of carboxylic acid-amine interaction: preparation and characterization,” E-Polymers, Article No. 080.

Yoshimura, T., Matsuo, K., and Fujioka, R. (2005). “Synthesis and characterization of superabsorbent polymers from various kinds of cotton balls,” Konbunshi Ronbunshu 62(7), 321-325.

Yoshimura, T., Matsuo, K., and Fujioka, R. (2006). “Novel biodegradable superabsorbent hydrogels derived from cotton cellulose and succinic anhydride: Synthesis and characterization,” J. Appl. Polymer Sci. 99(6), 3251-3256.

Yoshinobu, M., Morita, M., and Sakata, I. (1992). “Porous structure and rheological properties of hydrogels of highly water absorptive cellulose graft copolymers,” J. Appl. Polym. Sci.45(5), 805-812.

Young, R. A. (1976). “Wettability of wood pulp fibers. Applicability of methodology,” Wood Fiber 8(2), 120-128.

Zamani, A., Henriksson, D., and Taherzadeh, M. J. (2010). “A new foaming technique for production of superabsorbents from carboxymethyl chitosan,” Carbohyd. Polymers 80(4), 1091-1101.

Zhang, J., Jiang, N., Dang, Z., Elder, T. J., and Ragauskas, A. J. (2008a). “Oxidation and sulfonation of cellulosics,” Cellulose 15(3), 489-496.

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, Y. T., Fan, L. H., Cheng, L. H., Zhang, L., and Chen, H. L. (2009). “Preparation and morphology of high-performance exfoliated poly(sodium acrylate)/hydrotalcite nanocomposite superabsorbents,” Polymer Eng. Sci. 49(2), 264-271.

Zhang, Y. T., Fan, L. H., Zhao, P. Z., Zhang, L., and Chen, H. L. (2008b). “Preparation of nanocomposite superabsorbents based on hydrotalcite and poly(acrylic-co-acryla-mide) by inverse suspension polymerization,” Composite Interfaces 15(7-9), 747-757.

Zhang, Z., Wang, Z. C., Chang, Y., Wu, Z. C., Xu, J., and Lei, Z. Q. (2010). “Preparation of superabsorbent composites by graft copolymerization of acrylamide onto native cellulose,” Proceedings of the 7th National Conference on Chinese Functional Materials and Applications, pp. 531-533.

Zohuriaan-Mehr, M. J., and Kabiri, K. (2008). “Superabsorbent polymer materials: A review,” Iranian Polymer J. 17(6), 451-477.

Zohuriaan-Mehr, M. J., Omidian, H., Doroudiani, S., and Kabiri, K. (2010). “Advances in non-hygienic applications of superabsorbent hydrogel materials,” J. Mater. Sci. 45(21), 5711-5735.