You are on page 1of 10

Cellulose (2010) 17:11371146 DOI 10.

1007/s10570-010-9448-0

Inuence of lignin on cellulose-NaOH-water mixtures properties and on Aerocellulose morphology


Romain Sescousse Audrey Smacchia Tatiana Budtova

Received: 4 April 2010 / Accepted: 19 August 2010 / Published online: 1 September 2010 Springer Science+Business Media B.V. 2010

Abstract Microcrystalline cellulose and organosolv lignin, both dissolved in 8%NaOH-water, were mixed with the objective to study the inuence of lignin on the properties of cellulose solutions and on the morphology of dry porous materials. Mixture viscosity and gelation were investigated. Cellulose-lignin gels were regenerated in aqueous acid baths and dried under supercritical CO2 to obtain Aerocellulose, an aerogellike material. The presence of lignin in the mixture speeded up gelation. During regeneration part of lignin was washed out. This created large pores and channels in the dry materials. The overall results obtained showed that cellulose and lignin are not compatible in the solvent used. Keywords Cellulose Lignin Mixtures Gelation Porous materials Introduction Cellulose and lignin are the main components in a vegetal ber; together they are the most abundant

CEMEF is the member of the European Polysaccharide Network of Excellence (EPNOE), http://www.epnoe.eu. R. Sescousse A. Smacchia T. Budtova (&) riaux Mines ParisTech, Centre de Mise en Forme des Mate (CEMEF), UMR CNRS/Ecole des Mines de Paris 7635, BP 207, 06904 Sophia-Antipolis, France e-mail: Tatiana.Budtova@mines-paristech.fr

natural polymers. Being studied for more than one century, in our days they experience a renaissance due to the increased interest in making new materials from biomass-based polymers. One of the background questions is if cellulose and lignin are compatible and how the presence of lignin inuences the properties of cellulosic materials? To answer these questions is important for several reasons. On one hand, cellulose and lignin coexist in cell walls of bers making a sort of a complex multi-level composite. Each component brings its own properties: mechanical strength from crystalline regions of cellulose; hydrophobicity, binding and protection against microorganisms from lignin. The synergy of their properties, together with the ones of hemicellulose, results in what is known as natural bers with all their varieties and special properties. On the other hand, for practically all applications cellulose and lignin have to be separated. During the separation, cellulose and lignin are in contact being in solution or suspension state (in papermaking, for example); it is important to know if they interact or not. Finally, some applications suggest using lignin as a component blended with polysaccharides. If cellulose or its derivatives are mixed with lignin, will the nal material demonstrate the synergetic properties as it occurs in a natural ber? The answer depends on the type of lignin, common solvent and mixing conditions used. In overall, it seems that in literature two main trends coexist: one which indirectly shows that cellulose and lignin are incompatible, and the

123

1138

Cellulose (2010) 17:11371146

other, also indirectly, suggests the opposite. A very short overview of various points of view is given below. a) Coexistence of cellulose and lignin in a ber. Using various non-destructive experimental techniques it was shown that lignin is oriented preferentially parallel to the surface of the cell wall (Terashima and Seguchi 1988; Atalla and Agarwal 1985). Carbohydrates, in general, seem to form complexes with lignin as a unit of secondary cell wall structure (Chesson 1993). Molecular dynamics was used to study the growth and deposition of lignin on a cellulose surface in a cell wall. These studies supported the above mentioned experimental nding and demonstrated not only a certain spatial organization of lignin around cellulose microbrils, but also gave some evidences of links between two polymers (Houtman and Atalla 1995; Perez et al. 2004; Besombes and Mazeau 2005). However, the nature of these links was not specied. Ordered supramolecular model lignin structures were obtained experimentally when lignin was synthesized directly on cellulose acetate substrate (Micic et al. 2003). The ordering was interpreted as being due to substrate OH groups. b) Separation of cellulose and lignin. In papermaking the maximal removal of lignin is needed to improve the hydrogen bonding between clean cellulosic bers, to increase their swelling and avoid light absorption. The re-deposition of lignin onto cellulose bers during cooking, washing and bleaching is thus of great importance, and the question on the compatibility and interactions between cellulose and lignin arises. According to Maximova et al. (2001, 2004) who studied the absorption of lignin by cellulose bers, it is absorbed due to the capillary forces (imbibition) and can be easily removed by washing with water. They concluded that there was no true molecular adsorption of lignin onto bers under all conditions investigated (various pH and ionic strengths). c) Blends of lignin with cellulose and cellulose derivatives. The structure and properties of lms and bers made from lignin-cellulose derivative blends have been extensively studied in the nineties (see, for example, Rials and Glasser

1989, 1990; Dave et al. 1992; Dave and Glasser 1997). Fibers and lms were usually prepared by mixing polymers in a common solvent followed either by solvent evaporative drying (lms) or wet spinning (bers). In the case of organosolv lignin (OL)-hydroxypropyl cellulose (HPC) lms, miscible amorphous blends at low OL concentrations in the mixture and phase separation above 50% of lignin were observed (Rials and Glasser 1990). For the mixtures of OL with ethyl cellulose or with cellulose acetate butyrate (CAB) similar results were reported but lignin content at with phase separation occurred was lower, about 510%. As for bers spun from organosolv lignin-CAB mixtures and regenerated in water, SEM and TEM results showed pore size increase with the increase of OL content. Analyzing the above-mentioned results obtained on OL-cellulose derivatives mixtures (Glasser et al. 1998), authors claim that there is an evidence of strong intermolecular interactions between the two components. Cellulose and lignin were mixed in dimethylacetamide/LiCL solvent and mixture viscosity was studied by Glasser et al. 1998: an increase of mixture dynamic elastic modulus with the addition of lignin was recorded and it was explained by strong secondary interactions between the components. The goal of this work is to study how the addition of lignin inuences the properties of cellulose8%NaOH-water solutions and of dry porous material, Aerocellulose, made from these solutions. (79)% NaOH-water is cellulose solvent and there are a lot of attempts to use it for making lms, bres, membranes and sponges as far as this way of cellulose processing is much more environmentally friendly than viscose process. Certain simplications, as compared with the naturally occurring systems, were made: we used microcrystalline cellulose and organosolv lignin. We rst studied mixture viscosity and compared the result obtained with various cases known for polymer pairs forming interpolymer complexes or making an immiscible blend. Gelation of cellulose-8%NaOH solution in the presence of lignin was investigated. Cellulose-lignin dry porous samples were then prepared according to the procedure developed for making Aerocellulose, pure cellulose ultra-light aerogel-like materials (Gavillon and Budtova 2008;

123

Cellulose (2010) 17:11371146

1139

Sescousse and Budtova 2009). They were made from the gelled mixtures followed by cellulose regeneration in water or acid baths and drying in supercritical conditions. The morphology of ligno-Aerocellulose was studied and correlated with the rheological and regeneration observations.

Experimental part Materials Microcrystalline cellulose Avicel PH101 (cellulose in the following) with degree of polymerization of 170 as given by the manufacturer was purchased from SigmaAldrich. Organosolv lignin (lignin in the following) was in powder, obtained with Organocell process (Lindner and Wegener 1989, 1990) kindly provided by vTI-Institute for Wood Chemistry, Hamburg, Germany. The initial pulp was Norway Spruce. The molecular weight is Mw = 7,300 g mol-1 with polydispersity of 4.1, as determined with GPC using polyethylene glycol calibration. This lignin is soluble in aqueous solutions of neutral and high pH. Acetic acid was of 1 mol L-1 received from VWR; it was diluted with distilled water to desired concentrations to use as regenerating bath liquid. NaOH in pellets (98% purity) and acetone (98% purity) were also purchased from VWR. Distilled water was used for solution preparation and for regenerating bath. CO2 for drying in supercritical conditions was supplied by Air Liquide with a purity of 99.9%. Methods Samples preparation Cellulose-NaOH solutions were prepared as described elsewhere (Gavillon and Budtova 2008; Sescousse and Budtova 2009). Briey, an aqueous solution of 12% wt NaOH was cooled down to -6 C. Cellulose was swollen in distilled water in a certain proportion and kept at 5 C. 12% NaOHwater and swollen-in-watercellulose were mixed at -6 C with a stirring rate of 1,000 rpm for 2 h. The ready solutions of various cellulose concentrations in 8%NaOH-water were stored at 5 C to avoid aging.

Lignin was dissolved in aqueous 8%NaOH by stirring for a few hours at room temperature. Cellulose-lignin mixtures in 8%NaOHwater were prepared by simply mixing ready solutions in different proportions. All mixtures prepared were visually homogeneous and did not show any phase separation within 1 week storage time in refrigerator. Two types of mixtures were prepared: (a) series of mixtures with the total polymer concentration being constant and (b) mixtures with cellulose concentration being constant and lignin concentration (and thus total polymer concentration) varied. To prepare dry porous Aerocellulose-type samples, cellulose-lignin-8%NaOH-water mixtures were poured into cylindrical moulds with dimensions of about 35 cm height and 2 cm diameter and kept at 65 C for 2 h. These conditions were chosen to ensure cellulose gelation (Roy et al. 2003; Gavillon and Budtova 2008). The gels were then placed in regenerating bath of either distilled water, or 0.1 and 1 mol L-1 acetic acid. Bath liquid was regularly changed until pH did not vary anymore indicating that all NaOH is washed out. The samples were then rinsed with water to remove acid and then washed in acetone to remove water which is not compatible with CO2. Swollen-in-acetone samples were dried in supercritical CO2 conditions as described earlier (1 L autoclave, 80 bar, 35 C) by Gavillon and Budtova 2008; Sescousse and Budtova 2009. After depressurisation (4 bar per hour at 37 C), light brown samples of ligno-Aerocellulose were extracted and analysed. Experimental methods Flow and gelation of cellulose and lignin solutions and cellulose-lignin mixtures in 8%NaOH-water were studied using stress-controlled Bohlin Gemini rheometer. Steady state and dynamic shear experiments were carried out using cone-plate geometry equipped with a Peltier temperature control system. Solutions or mixtures were directly introduced on the plate after the preparation, and a layer of low viscosity silicon oil was put around to prevent water evaporation. Gelation of cellulose and celluloselignin mixtures was studied at 15 C at frequency 1 Hz and stress 1 Pa. Scanning electron microscopy (SEM) experiments were performed using A PHILIPS XL30 on samples

123

1140

Cellulose (2010) 17:11371146


0.1

covered with gold at the acceleration voltage of 15 kV. Porosity of ligno-Aerocellulose was measured in SAFT, Bordeaux, France using mercury intrusion with Micromeritics Autopore IV porosimeter. The details on data treatment will be given in Results section.

viscosity, Pa s

0 10 20
0.01

35

Results and discussion Cellulose-lignin mixtures in aqueous 8%NaOH In order to understand if and how cellulose and lignin macromolecules are interacting in the common solvent, the behaviour of the initial components in 8% NaOH-water has to be rstly characterised. The ow of cellulose-(7-9%)NaOH-water solutions and viscosityconcentration-temperature dependences have been studied extensively for various cellulose types and concentrations and in different conditions (Roy et al. 2003; Egal 2006; Gavillon and Budtova 2008). It was shown that above the overlap concentration solutions are gelling with time and temperature increase. Gelation time is exponentially temperature-dependent, as shown by Roy et al. 2003; Gavillon and Budtova 2008. The viscosity of cellulose solutions in conditions far from gel point is Newtonian below the overlap concentration and slightly shear thinning above it. The activation energy Eact cell for 15% solutions is about 1925 kJ mol-1, respectively (Roy et al. 2003; Egal 2006). The ow of lignin-8%NaOH-water was investigated for 620% lignin concentrations in the temperature interval from 0 to 35 C. In all cases studied lignin solutions behave like Newtonian uids; an example for 20% lignin-8%NaOH-water is shown in Fig. 1. This is what should be expected due to lignin low molecular weight. Viscosity-temperature dependence is illustrated in the Arrhenius plot (Fig. 2) for 15 and 20% lignin solutions; data for 9%NaOH-water taken from Roy et al. 2003 are shown for comparison. As expected, the activation energies of lignin solutions slightly increase with the increase of polymer concentration: Eact 15%lign = 16 2 kJ mol-1 Eact 20%lign = 18 2 kJ mol-1. One way to estimate the interactions between the components in a common solvent is to study how mixture viscosity varies with its composition. The
0.001 1 10 100

shear rate, 1/s


Fig. 1 Flow of 20%lignin-8%NaOH-water solution at various temperatures

-2 0.36 -3 0.38 0.4 0.42

1/RT, mol/kJ
0.44

-4
1

-5
2

-6
3

-7

-8

ln, ln(Pa s)

Fig. 2 Arrhenius plot for lignin-8%NaOH-water solutions for 20% (1) and 15% (2) lignin concentrations and also for 9%NaOH-water solution (3) (data taken from [16]). Solid lines are least square linear approximation

experimental data gexp obtained should be then compared with a theoretical dependence of mixture additive viscosity gadd calculated at different compositions which assumes the absence of special interactions between the components (i.e. polymer1 polymer2 interactions are the same as the interactions between the macromolecules of the same type). Various conclusions on polymer1polymer2 interactions can be made depending if the initial solutions

123

Cellulose (2010) 17:11371146

1141

Table 1 Conclusions on polymer1-polymer2 interactions that can be made from the comparison of experimental mixture viscosity gexp with the one calculated from the theoretical mixing rule gadd gexp [ gadd Dilute solutions gexp = gadd gexp \ gadd

Loose branched or a gel-like structure No interactions leading to new bonds Two options: (1) zip-type made of linked macromolecules of formation between polymer 1 and interpolymer complexes. Example: both types. Example: interpolymer polymer 2, the macromolecules polyelectrolyte complexes made of complexes formed via hydrogen coexist in the common solvent. High two oppositely charged bonding (Iliopoulos and Audebert probability that in certain conditions polyelectrolytes (Rogacheva et al. 1985; Nikolaeva et al. 2000) the components may phase separate 1984) or of a large sequence of as soon as there is a competition for hydrogen bonded polymer1solvent. polymer2 (Pinteala et al. 2005). (2) components are strongly incompatible (Chaudemanche and Budtova 2008). Two options: (1) The same as in dilute The same as in dilute state state (2) Micro-phase separated mixture: effect of excluded volume with one component making a concentrated continuous phase and gelling (Tecante and Doublier 1999; Tecante and Doublier 2002; Chaudemanche and Budtova 2008) The same as in dilute state

Semi-dilute solutions

are in dilute or semi-dilute state; they are summarized in Table 1. For the mixtures where the concentration of at least one of the components is above the overlap concentration, the additive viscosity gadd must be calculated according to the logarithmic rule: ln gadd /1 ln g1 /2 ln g2 1

where /1 and /2 are the weight fractions at which the components are present in the mixture (/1 ? /2 = 1); g1 and g2 are the components viscosities at /1 = 1 and /2 = 1, respectively. In order to plot the experimental dependence of cellulose-lignin mixture viscosity as a function of mixture composition, the ow of cellulose-lignin8%NaOH-water of various compositions with total polymer concentration of 6% was studied at 15 C. These conditions were chosen as a compromise between not too low viscosity of lignin solution and not too fast gelation of cellulose solutions (for example, 78% cellulose solutions are gelling irreversibly within a few minutes as soon as they are extracted from the thermobath where they were prepared at -6 C). As it will be shown in the following, the gelation of all 6% mixtures studied at 15 C is slow enough to be neglected for this type of measurements: the duration of one ow experiment was less than 3040 min while gelation takes more

than 10 h. Because the mixtures showed a slightly shear thinning ow, like observed previously for cellulose-(7-9)%NaOH-water solutions (Roy et al. 2003; Egal 2006; Gavillon and Budtova 2008), the viscosity values to be plotted as a mixture composition were taken at the shear rate of 10 s-1. The viscosity of 6% mixtures at various compositions is shown in Fig. 3 together with the calculated additive dependence. The experimental points coincide with the additive dependence demonstrating the evidence of the absence of any special interactions between cellulose and lignin in 8%NaOH-water that could lead to the formation of new bonds. This important nding means that not only there are no bonds built between cellulose and lignin molecules, but that these polymers coexist in the common solvent and may phase separate as soon as their total concentration becomes high enough. Electrostatic repulsive interactions between cellulose and lignin could be one of the reasons of the absence of bonds built between the components as far as both polymers should be negatively charged in 8%NaOH-water. In our previous studies we have shown that cellulose-(79)%NaOH-water solutions are irreversibly gelling with the increase of temperature, time and concentration (Roy et al. 2003; Egal 2006; Gavillon and Budtova 2008). The gelation is due to the fact that (79)%NaOH-water is thermodynamically not a good

123

1142
6% cellulose
0 0 -1 -2 -3
2

Cellulose (2010) 17:11371146


mixture composition
0.2 0.4 0.6 0.8

6% lignin
1

G, G, Pa

G
3

-4
1

-5 -6

ln, ln(Pa s)

gelation time 0 0 2000 4000 6000 8000

time, s
10000

Fig. 3 Viscosity versus mixture composition for mixtures with 6% total polymer concentration. Points are experimental data; line corresponds to the additive dependence calculated according to Eq. 1 (see details in the text)

Fig. 4 Viscous (G0 ) and elastic (G00 ) moduli evolution in time for 6%cellulose-1.5%lignin-8%NaOH-water mixture at 15 C

cellulose solvent. Solution aging (time and temperature increase) leads to the preferential cellulose-cellulose and not cellulosesolvent interactions resulting in solution micro-phase separation and gelation. It was interesting to check if and how the presence of lignin inuences gelation of cellulose-8%NaOH-water solutions. Gelation time, tgel, was chosen as one of the important parameters characterizing gelation. It was determined from the evolution of elastic G0 and viscous G00 moduli at a xed temperature, 15 C, for various mixture compositions and taken at the moment when G0 = G00 (see an example in Fig. 4). Gelation time of cellulose-lignin mixtures is shown in Fig. 5 for three mixtures containing 5, 5.5 and 6% of cellulose. The amount of lignin was varied. Despite that low molecular weight lignin solutions in 8%NaOH-water are not gelling, lignin presence in the mixture, even in low amounts, accelerates cellulose-8%NaOH gelation (see a drastic decrease in gelation time for 5%cellulose-1% lignin as compared to 5%cellulose-0%lignin or with the increase of lignin content in the mixture of 6%cellulose with 0.5, 1, 1.5 and 2% lignin, Fig. 5). Another way of looking at the same data gives a key in the understanding of what is happening in cellulose-lignin mixtures. Let us consider mixtures with the same total polymer concentration, 6%: 6%cellulose, 5.5%cellulose-0.5%lignin and 5%cellulose-1%lignin (shown with a circle in Fig. 5). On one hand, a decrease of cellulose concentration from 6 to

35 30

tgel, h
5% cellulose

25

1
20 15 10 5 0 0 2

5.5% cellulose 6% celulose

Clign, %
4 6

Fig. 5 Gelation time as a function of lignin concentration in the mixtures containing 5% (1), 5.5% (open point) and 6% (2) cellulose. Lines are given to guide the eye. The circle shows the mixtures with the same total polymer concentration, 6%

5% in solutions without lignin leads to a strong increase in gelation time, as expected. On the other hand, the gelation of all 6% mixtures, pure 6%cellulose, 5.5%cellulose-0.5%lignin and 5%cellulose1%lignin, takes practically the same time which is surprising because 0.5 and 1% lignin solution has a very low viscosity and cannot be considered as a gelling component. A most probable explanation is that due to possible electrostatic repulsion between the polymers, lignin helps the formation of cellulose-

123

Cellulose (2010) 17:11371146

1143

rich and cellulose-poor domains, thus leading to cellulose-lignin micro-phase separation. The increase of local cellulose concentration facilitates gelation; lignin can thus be considered as gelation promoter for cellulose-8%NaOH solutions. The result obtained on gelation also suggests lignin and cellulose incompatibility in 8%NaOH-water. Cellulose-lignin gels should thus be much more heterogeneous as compared with pure cellulose-8%NaOH-water gels.

(a)

(b)

1 cm 1 cm

Regeneration of cellulose from cellulose-lignin8%NaOH-water gels and ligno-Aerocellulose morphology Cellulose from cellulose-NaOH-water gels can be regenerated if placing a gel into a liquid which is a non-solvent for cellulose (water, aqueous acid solutions, alcohols). A 3D object made of cellulose network with non-solvent liquid lling the pores is formed. Because lignin is not gelling and is not bound to cellulose, the question is what will happen with lignin during cellulose regeneration. The answer depends on the type of liquid used in regenerating bath. Qualitative observations are easy as far as lignin gives a dark brown color either to the samples or to the regenerating bath. When water bath was used, it became brown when cellulose-lignin gel was placed in it. This is because lignin was washed out from cellulose as far as lignin used is water-soluble. Higher was bath acidity, lower amount of lignin was washed out from the sample during regeneration because lignin is not soluble in acidic media and it coagulates inside the cellulose network. In the latter case samples remain dark: higher is bath acidity, darker are the samples. If cellulose-lignin-8%NaOH-water gels are very heterogeneous, regenerated samples should also be of very heterogeneous porosity. In order to study the porosity of swollen-in-non-solvent regenerated cellulose, two ways of drying that preserve the pores against collapse can be used: either freeze-drying or drying under super-critical conditions. We used the same way of drying in super-critical CO2 as reported for the preparation of ultra-light and highly porous pure cellulose, Aerocellulose (Gavillon and Budtova 2008; Sescousse and Budtova 2009). It was shown that Aerocellulose can be obtained from cellulose8%NaOH-water solutions and gels after cellulose

Fig. 6 Examples of ligno-Aerocellulose (a) made from the mixture 4%cellulose-8.57%lignin-8%NaOH-water and of Aerocellulose (b) from 5%cellulose-8%NaOH-water

regeneration in water, acid or alcohol bath, washing in acetone and drying in super-critical CO2. New ligno-Aerocelluloses with various lignin contents and of various porosities were obtained using the same approach as described above for Aerocellulose preparation. An example of ligno-Aerocellulose sample versus pure Aerocellulose is shown in Fig. 6. SEM images of dry lingo-Aerocellulose regenerated in the baths of two different acid concentrations are shown in Fig. 7. All the other preparation parameters were identical. Lower is acid bath concentration (Fig. 7a), more lignin was washed out and thus larger pores are obtained. The proportion between cellulose and lignin in two dry samples was determined in vTI-Institute for Wood Chemistry, Hamburg, Germany, for the lignoAerocelluloses made from the same initial mixture with high initial lignin content but regenerated in different baths: one of 0.1 mol L-1 acetic acid and the other of 1 mol L-1. The initial mixture was 4%cellulose-8.6%lignin, giving the initial proportion lignin:cellulose = 2.15. Samples preparation was identical except acetic acid concentration in regenerating bath. The amount of lignin in the lignoAerocelluloses was determined as a Klason and as acid-soluble lignin using UV spectrometer. The amount of acid-soluble fraction was very low, below 53% of Klason lignin. The proportion between lignin and cellulose obtained was 0.14 for the dilute acid bath and 0.33 for the concentrated one. In other words, keeping in mind that the amount of cellulose in wet and dry samples does no change, 82% of lignin

123

1144

Cellulose (2010) 17:11371146

Fig. 7 SEM images of ligno-Aerocellulose from celluloselignin mixtures: regenerated in 0.1 mol L-1 (a) and in 1 mol L-1 acetic acid bath. The scale bar is 2 lm.

Concentration of cellulose and lignin in the initial mixture was 4 and 4.3%, respectively. Gelation was at 65 C for 2 h

that was mixed with cellulose was then washed out during regeneration in 0.1 mol L-1 acetic acid bath and 65% was washed out in 1 mol L-1 bath. It is a direct proof that cellulose and lignin macromolecules are not bound together. The reason could be, as mentioned above, the electrostatic repulsive interactions. Mercury porosimetry was used to qualitatively analyse the porosity of ligno-Aerocellulose. An example of the cumulated volume V as a function of pressure for two ligno-Aerocelluloses made from the mixtures containing different amount of lignin but regenerated in the baths of the same acid concentration, 0.1 mol L-1, is shown in Fig. 8a. Pores cumulative volume is higher for the initial mixture containing larger amount of lignin (6.9 cm3 g-1 for sample 2 against 9 cm3 g-1 for sample 1, curves 2 and 1 in Fig. 8a, respectively) because a signicant part of lignin is washed out during regeneration, creating pores and channels. Mercury is not penetrating in Aerocelluloses as obtained from measuring the sample weight before after the experiment; the samples are totally compressed (see compressiondecompression notations in Fig. 8a). Pore size distribution was thus calculated using buckling theory developed by Pirard et al. 1995 for hyperporous materials and tested for polyurethane foams by Pirard et al. 2003 and cellulose acetate aerogels by Fischer et al. 2006. The result is shown in Fig. 6b for the same samples as in Fig. 8a. The approach of Pirard et al. (1995) suggests that the size D of the largest pores remaining after compression at a pressure P is determined as

(a) V, cm 3/g
10

decompression 2

compression
2

0 1.00E+03

P, Pa
1.00E+05 1.00E+07 1.00E+09

(b) dV/dD
8

2
2

0 0 1 2 3

arb. units
4

Fig. 8 Cumulated volume as a function of applied pressure (a) and pores size distribution (b) for the Aerocelluloses made from the initial 4%cellulose-3.3%lignin (1) and 4%cellulose8.6%lignin mixtures (2); gelation was at 65 C for 2 h, regenerated in 0.1 mol L-1 acetic acid and dried in supercritical CO2

123

Cellulose (2010) 17:11371146

1145

D kf P0:25

concentration in the mixture, larger pores were obtained in the dry material.
Acknowledgments This work was supported by ANR (France), Carbocell project, number ANR-06-MAPR-0004 and the EC 6th framework, European Polysaccharide Network of Excellence (EPNOE) project, number NMP3-CT-2005500375. Authors are grateful to Juegen Puls and Bodo Saake (vTI-Institute for Wood Chemistry, Hamburg, Germany) for providing and characterisation of lignin, to A. Rigacci and tique et Proce de s, Mines S. Berthon-Fabry (Centre Energe ParisTech, France) for stimulating discussions, to Pierre tique et Proce de s, Mines ParisTech, Ilbizian (Centre Energe France) for the supercritical CO2 drying, to B. Simon (Saft, France) for the measurements of Aerocellulose porosity and to S. Jacomet (CEMEF, Mines ParisTech) for help with SEM experiments.

where kf is buckling strength constant which depends on the mechanical properties of pore walls. As far as kf is not known for Aerocellulose and requires special measurements which were out of the scope of this work, the pore sizes were calculated in arbitrary units as D * P-0.25. This approximation does not give the absolute values of pore diameters but allows a qualitative comparison of the size distribution dV/dD for two ligno-Aerocellulose samples, as far as their buckling constant is the same. As expected, the pore size distribution for the sample 2 with higher initial lignin content (Fig. 8b) is shifted towards higher pore size values and is wider than the one with lower initial lignin content (curve 1). This difference is due to large voids that are formed because of washed-out lignin during regeneration: higher was lignin concentration in the mixture, more there are large pores. This trend is also reected by the density: 0.135 and 0.1 g cm-3 for sample 1 and 2, respectively. However, the specic surface obtained with BET analysis for these two and other samples is practically the same within the experimental errors, around 200 20 m2 g-1. The reason is that large pores do not make any signicant contribution to pores surface. Similar values were obtained for Aerocelluloses prepared from pure cellulose solutions, 8%NaOH-water (Gavillon and Budtova 2008) and N-methylmorpholine-N-oxide (Innerlohinger et al. 2006).

References
Atalla RH, Agarwal UP (1985) Raman microprobe evidence for lignin orientation in the cell wall of native tissue. Science 227:636637 Besombes S, Mazeau K (2005) The cellulose/lignin assembly assessed by molecular modeling. Part 2: seeking for evidence of organization of lignin molecules at the interface with cellulose. Plant Physiol Biochem 43:277286 Chaudemanche C, Budtova T (2008) Mixtures of pregelatinised maize starch and k-carrageenan: compatibility, rheology and gelation. Carbohydr Polym 72:579589 Chesson A (1993) Mechanistic model of forage cell wall degradation. In: Jung HG, Buxton DR, Hateld RD, Ralph J (eds) Forage cell wall structure and digestibility. UAS Wisconsin, Madison, pp 358 Dave V, Glasser WG (1997) Cellulose-based bers from liquid crystalline solutions: 5. Processing and morphology of CAB blends with lignin. Polymer 38:21212126 Dave V, Glasser WG, Wilkies GL (1992) Evidence of cholesteric morphology in lms of cellulose acetate butyrate by transmission electron microscopy. Polym Bull 29: 565570 Egal M (2006) Structure and properties of cellulose/NaOH aqueous solutions, gels and regenerated objects. PhD thesis, Ecole des Mines de Paris/Cemef, Sophia-Antipolis, France Fischer F, Rigacci A, Pirard R, Berthon-Fabry S, Achard P (2006) Cellulose-based aerogels. Polymer 47:76367645 Gavillon R, Budtova T (2008) Aerocellulose: new highly porous cellulose prepared from cellulose-NaOH aqueous solutions. Biomacromolecules 9:269277 Glasser WG, Rials TG, Kelley SS, Dave V (1998) Studies of the molecular interaction between cellulose and lignin as a model for the hierarchical structure of wood. In: Heinze TJ, Glasser WG (eds) Cellulose derivatives. Modication, characterization and nanostructures. ACS Symposium series 688, Chapter 19. Orlando, pp 265282 Houtman CJ, Atalla RH (1995) Cellulose-lignin interactions. A computational study. Plant Physiol 107:984997

Conclusions Mixtures of cellulose and organosolv lignin in a common solvent, 8%NaOH-water, were prepared. Mixture viscosity coincides with the one calculated according to the mixing rule thus suggesting the absence of any bonds made between cellulose and lignin due to electrostatic repulsion. The presence of lignin accelerates cellulose gelation which was hypothised to be due to the formation of celluloserich domains. Highly heterogeneous porous samples were obtained via cellulose regeneration and drying under CO2 super-critical conditions. A signicant part of lignin was washed out during regeneration conrming the repulsion between cellulose and lignin in NaOH-water solvent. Higher was lignin

123

1146 Iliopoulos I, Audebert R (1985) Inuence of concentration, molecular weight and degree of neutralization of polyacrylic acid on interpolymer complexes with polyoxyethylene. Polym Bull 13:171178 Innerlohinger J, Weber HK, Kraft G (2006) Aerocellulose: aerogels and aerogel-like materials made from cellulose. Macromol Symp 244:126138 Lindner A, Wegener G (1989) Characterisation of lignins from organosolv pulping according to the organocell process, Part 2 Residual Lignins. J Wood Chem Technol 9: 443465 Lindner A, Wegener G (1990) Characterisation of lignins from organosolv pulping according to the organocell process, Part 3 Permanganate oxidation and thioacidolysis. J Wood Chem Technol 10:331350 Maximova N, Osterberg M, Koljonen K, Stenius P (2001) Lignin adsorption on cellulose bre surfaces: effect on surface chemistry, surface morphology and paper strength. Cellulose 8:113125 Maximova N, Stenius P, Salmi J (2004) Lignin uptake by cellulose bers from aqueous solutions. Nordic Pulp Paper Res J 19:135145 Micic M, Radotic K, Jeremic M, Leblanc RM (2003) Study of self-assembly of the lignin model compound on cellulose model substrate. Macromol Biosci 3:100106 Nikolaeva O, Budtova T, Alexeev V, Frenkel S (2000) Interpolymer association between polyacrylic acid and cellulose ethers: formation and properties. J Polym Sci Part B Polym Phys 38:13231330 Perez DDS, Ruggiero R, Morais LC, Machado AEH, Mazeau K (2004) Theoretical and experimental studies on the adsorption of aromatic compounds onto cellulose. Langmuir 20:31513158 Pinteala M, Budtova T, Epure V, Belnikevich N, Harabagiu V, Simionescu BC (2005) Interpolymer complexes between

Cellulose (2010) 17:11371146 hydrophobically modied poly(methacrylic acid) and poly(N-vinylpyrrolidone). Polymer 46:70477054 Pirard R, Blacher S, Brouers F, Pirard JP (1995) Interpretation of mercury porosimetry applied to aerogels. J Mater Res 10:21142119 Pirard R, Rigacci A, Marechal JC, Quenard D, Chevalier B, Achard P, Pirard JP (2003) Characterization of hyperporous polyurethane-based gels by non-intrusive mercury porosimetry. Polymer 44:48814887 Rials TG, Glasser WG (1989) Multiphase materials with lignin VI. Effect of cellulose derivative structure on blend morphology with lignin. Wood Fiber Sci 21:8090 Rials TG, Glasser W (1990) Multiphase materials with lignin: 5. Effect of lignin structure on hydroxypropyl cellulose blend morphology. Polymer 31:13331338 Rogacheva VB, Ryzhikov SV, Shchors TV, Zezin AB, Kabanov VA (1984) Structural chemical transformations of non-stoichiometric polyelectrolyte complexes in watersalt solutions. Polym Sci USSR 26:27482756 Roy C, Budtova T, Navard P (2003) Rheological properties and gelation of aqueous cellulose-NaOH solutions. Biomacromolecules 4:259264 Sescousse R, Budtova T (2009) Inuence of processing parameters on regeneration kinetics and morphology of porous cellulose from cellulose-NaOH-water solutions. Cellulose 6:417426 Tecante A, Doublier JL (1999) Steady ow and viscoelastic behavior of crosslinked waxy corn starch-j-carrageenan pastes and gels. Carbohydr Polym 40:221231 Tecante A, Doublier JL (2002) Rheological investigation of the interaction between amylose and j-carrageenan. Carbohydr Polym 49:177183 Terashima N, Seguchi Y (1988) Heterogeneity in formation of lignin. IX. Factors affecting the formation of condensed structures in lignin. Cell Chem Technol 22:147

123

You might also like