Plan
Comptes Rendus

Plant biology and pathology / Biologie et pathologie végétales
Micromechanical understanding of the cell-wall structure
[Micromécanique de la structure de la paroi cellulaire.]
Comptes Rendus. Biologies, Volume 327 (2004) no. 9-10, pp. 873-880.

Résumés

For improving properties of pulp fibres, a better understanding of the relationships between its macroscopic mechanical properties, fibre ultrastructure, and properties of the wood polymers is important. This paper discusses such relations between elastic properties of fibres, their matrix structure and the wood polymer elastic constants. It is argued that an orientation of all of the wood polymers in the direction of the cellulose microfibrils is most likely. The elastic longitudinal modulus of cellulose is well described by the value of 134 GPa dominating the longitudinal fibre properties. In the transverse direction the amorphous polymers play a more important role.

Pour améliorer les propriétés des fibres de la pâte à papier, une meilleure compréhension des relations entre les propriétés mécaniques macroscopiques, l'ultrastructure des fibres et les propriétés des polymères constitutifs du bois est nécessaire. Cet article discute des relations entre les propriétés élastiques des fibres, la structure de leur matrice pariétale et les constantes d'élasticité des polymères du bois. Il est proposé qu'une orientation de tous les polymères du bois dans la direction des microfibrilles de la cellulose est très vraisemblable. Le module d'élasticité longitudinal de la cellulose, égal à 134 GPa, est prépondérant pour les propriétés longitudinales des fibres. Dans la direction transversale, les polymères amorphes jouent un rôle plus important.

Métadonnées
Reçu le :
Accepté le :
Publié le :
DOI : 10.1016/j.crvi.2004.03.010
Keywords: hemicelluloses, cellulose, lignin, modelling, elastic properties, wood fibres, softening
Mot clés : hémicelluloses, cellulose, lignine, modélisation, propriétés élastiques, fibres du bois, amollissement
Lennart Salmén 1

1 STFI-Packforsk AB, Box 5604, SE-114 86 Stockholm, Sweden
@article{CRBIOL_2004__327_9-10_873_0,
     author = {Lennart Salm\'en},
     title = {Micromechanical understanding of the cell-wall structure},
     journal = {Comptes Rendus. Biologies},
     pages = {873--880},
     publisher = {Elsevier},
     volume = {327},
     number = {9-10},
     year = {2004},
     doi = {10.1016/j.crvi.2004.03.010},
     language = {en},
}
TY  - JOUR
AU  - Lennart Salmén
TI  - Micromechanical understanding of the cell-wall structure
JO  - Comptes Rendus. Biologies
PY  - 2004
SP  - 873
EP  - 880
VL  - 327
IS  - 9-10
PB  - Elsevier
DO  - 10.1016/j.crvi.2004.03.010
LA  - en
ID  - CRBIOL_2004__327_9-10_873_0
ER  - 
%0 Journal Article
%A Lennart Salmén
%T Micromechanical understanding of the cell-wall structure
%J Comptes Rendus. Biologies
%D 2004
%P 873-880
%V 327
%N 9-10
%I Elsevier
%R 10.1016/j.crvi.2004.03.010
%G en
%F CRBIOL_2004__327_9-10_873_0
Lennart Salmén. Micromechanical understanding of the cell-wall structure. Comptes Rendus. Biologies, Volume 327 (2004) no. 9-10, pp. 873-880. doi : 10.1016/j.crvi.2004.03.010. https://comptes-rendus.academie-sciences.fr/biologies/articles/10.1016/j.crvi.2004.03.010/

Version originale du texte intégral

1 Introduction

The properties of wood fibre materials are highly dependent on the properties of its constituents as well as on the arrangement and properties of the wood polymers. In general, there is a rather good knowledge of how various wood products relate to the fibre properties on a macroscopic level. However, when it comes to the microscopic level in understanding how the different wood polymers interact and contribute to the fibre properties and how these interactions are affected by processing, information is still lacking.

Cellulose is of course the main contributor to the strength properties of the wood fibre with its highly arranged crystalline structure. However, changes occurring in this crystalline and aggregate structure during processing have lately been shown to be of large importance for the properties of the fibres [1]. In this rearrangement also, the hemicelluloses has been claimed to play an important role [2]. For transverse properties of fibres, the hemicelluloses and lignin have also been shown to play a much more important role than the cellulose [3]; it is the reason why a more precise description of the arrangement of these constituents is necessary.

In this article, an effort is made to analyse various factors of importance for the understanding of the relation between the properties of the wood constituents, its ultrastructure and the macroscopic properties of the fibre material.

2 Ultrastructural arrangement

The structural arrangement of the wood polymers within the cell wall could from a mechanistic standpoint be viewed by analogy as a fibre-reinforced composite system, with cellulose fibrils as reinforcing materials in a matrix of lignin and hemicelluloses. The present conception of the arrangement could be illustrated as in Fig. 1.

Fig. 1

Schematic illustration of the arrangement of the wood polymers within the secondary wall of a softwood fibre (adopted after the original picture of Åkerholm and Salmén [4]).

The cellulose microfibrils in a typical softwood fibre are 3 to 4 nm in cross section [5] and aggregated into larger entities, cellulose aggregates, of an average cross section of 16 to 20 nm [6–8]. Most probably, some of the glucomannans are incorporated into these aggregates [7,9,10]. Adjacent to these aggregates mostly glucomannan is arranged in parallel with the fibrils having a high degree of coupling to them [11]. In the space of 3 to 4 nm in-between these cellulose aggregates [12], it is likely that the lignin matrix is interspaced with xylan entities [13]. Based on spectroscopic data, indications are also that the lignin has an orientation of the phenylpropane units along with the direction of the cellulose aggregates [4]. An orientation of the plane of the aromatic rings of the lignin in the tangential direction of the cell-wall surface has also been demonstrated by Raman studies [14,15]. This lignin although highly ordered in the structure in-between the cellulose aggregates seems to behave more freely; it is shown to exhibit a high degree of viscoelasticity at normal room temperature and humidity conditions in contrast to the more elastic properties of the carbohydrates [16]. For the xylan within the matrix, a preferred orientation has been indicated, i.e. the xylan is not arranged in a random fashion [16]. Presumably the xylan has an arrangement parallel to the cellulose fibrils in similarity with the other wood polymers.

The arrangement of the matrix on a larger scale is that of a tangential lamellation [17] as evident when observing the structure of the cross section of the secondary wall by AFM (Fig. 2). Here it is evident that also the cellulose aggregates line up in rows in the tangential direction [8]. However one must remember that on the three-dimensional level the structure of the cellulose aggregates are probably more resembling the lenticular structure proposed by Boyd [18] schematically illustrated in Fig. 3. Such an undulating behaviour of the cellulose aggregates and microfibrils are what has been observed when viewing the surface structure of the delignified secondary wall by electron microscopic techniques [19]. Thus the matrix material arranged in the tangential structure will be placed as successive disks both along the circumferential direction of the fibre as well as in the length direction of it. The size of these matrix material elements is probably to a high degree reflecting the size of the dissolved lignin particles, i.e. comparable to a spherical diameter of 5 to 10 nm [20,21]. Pore size distribution measurements also indicate that the holes created in the fibre structure from the dissolved lignin and hemicelluloses are in this range; i.e. with a mean diameter of 10 nm [22–24].

Fig. 2

Atomic Force Microscopy (AFM) image of a transverse cross section of the secondary wall of a spruce wood fibre. Cellulose aggregates (lighter parts, averaging 20 nm, arranged in a pattern of tangential lamellation) are evident [8].

Fig. 3

Schematic drawing of the cellulose aggregate structure adopted after the ideas of Boyd [18].

The matrix polymers thus exist in a very homogeneous blend in-between the ‘web’ of the cellulose aggregates. It is also apparent that it is only at the high temperatures in wet conditions above 100 °C, when the lignin softens, that an aggregation of the cellulose takes place, presumably an enlargement of the contract zones in the ‘web’ of the aggregates in Fig. 3 [8]. Based on rheological properties of wood fibres, it is however concluded that lignin and maybe also hemicelluloses exhibit their own separate softening temperatures [13], thus pointing to a mechanically heterogeneous mixture. Such behaviour requires material entities of around 5 to 15 nm in diameter [25,26], a situation that is not at all unrealistic considering the actual wood-fibre wall structure. However, it has from a mechanical point been mathematically shown that the matrix arrangement, i.e. that of a heterogeneous or homogeneous mixture, has very little influence on the elastic properties of the cell wall both along the fibre axis and in the transverse direction [3].

3 Rheological properties of wood polymers

Of the wood polymers, cellulose, hemicelluloses, and lignin, most attention has been given to the cellulose when it comes to determine mechanical properties of the wood constituents. The rationale is of course the cellulose being the stiff reinforcing material in the composite and therefore the component that to a high degree determines the properties of the fibre wall.

3.1 Cellulose

For cellulose there are both theoretical estimates as well as experimental determinations of the elastic properties. Most recent analysis tends to favour values in the longitudinal direction for the cellulose crystalline region between 120 to 170 GPa [27], although there are some older theoretical estimates of such high values as 246 and 319 GPa [28,29]. When comparing theoretical estimates with experiments both for cellulose I as well as the experimentally easier material cellulose II, it seems reasonable to come to a conclusion of a modulus value of ca 134–136 GPa for the stiffness of cellulose I [29].

3.2 Hemicelluloses

For hemicelluloses data are extremely scarce. Cousins' data on extracted xylan fractions [30] indicate an elastic modulus for the dry material of 8 GPa. However, it can be questioned how representative this value is. It is known that hemicelluloses will crystallize when isolated. Recent measurements performed on amorphous hemicellulose films made of Konjac glucomannan (only containing mannose and glucose units) indicate a glassy modulus in the range of 2 GPa and a softened modulus of 20 MPa [31]. It is also known that the hemicelluloses at least the glucomannan is oriented along the cellulose fibrils [11,32,33]. Considering an orthotropic arrangement in the cell wall, a set of reasonable moduli is depicted in Table 1.

Table 1

Wood polymer data used for model calculations

Polymer Dry conditions, i.e. 12% moisture, 20 °C Moist conditions, 20 °C
Cellulose
Ex 134 GPa 134 GPa
Ey 27.2 GPa 27.2 GPa
G 4.4 GPa 4.4 GPa
ν 0.1 0.1
l/d 50 000 50 000
Hemicellulose
Ex 2.0 GPa  20 MPa
Ey 0.8 GPa  8 MPa
G 1.0 GPa  10 MPa
ν 0.2 0.2
Lignin
Ex 2.0 GPa 2.0 GPa
Ey 1.0 GPa 1.0 GPa
G 0.6 GPa 0.6 GPa
ν 0.3 0.3

3.3 Lignin

The lignin structure within the softwood fibre varies from that of a more branched nature in the middle lamella and outer cell walls to that of the normal less branched guaiacyl type in the secondary wall of spruce wood fibres, a fact that also has a bearing on the physical properties [34]. Accurate elastic properties of the lignin in question may therefore be difficult to assess. Under dry conditions, values for an extracted isotropic lignin have been reported to between 4 to 7 GPa [35,36]. However, for the lignin also, an oriented structure along the cellulose fibrils has to be accounted for [4,15]. Reasonable values for lignin at 20 °C and 12% moisture content and wet are given in Table 1. Even under wet conditions, lignin retains the high stiffness up to its softening, above ca 80 °C [13].

4 Properties along fibres

In the fibre direction, a lot of measurements on elastic properties for different cellulosic fibre materials exist. It is quite clear that an important factor determining the elasticity is the fibril angle of the S2-layer. It is also clear that the properties of the cellulose almost solely determine the value of the elasticity, while both the elasticity of the lignin and the hemicelluloses have only a marginal influence [3]. Thus, the factors that govern the longitudinal fibre elasticity are the stiffness of the cellulose, the content of cellulose and the fibril angle of these cellulose microfibrils. This also means that the relative thickness of the S2-layer is of importance [37], as it quantifies the proportion of cellulose fibrils in the longitudinal direction. Variations between species of the relative proportion of the S2-layer in the range from 50 to 80% on a volumetric basis have been found [38]. The thickness proportion of the S2-layer for spruce fibres is though in the range of 80 to 90% [39].

When it comes to comparing theoretical estimates with different experimental measurements the difficulties mostly arises due to the lack of data on the factors of importance such as chemical composition, proportion of different cell wall layers, etc. There are of course also difficulties in the values given for the elastic measurements themselves when it comes to fibres, for which it is often clear that a lot of the fibres have been damaged in the process of isolation [40].

In Fig. 4, the influences of the factors of importance in the longitudinal direction, i.e. the proportion of S2-wall (varied from a typical value of 79% to a more extreme of 84%), and the content of cellulose in the S2-wall (varied from a typical value of 50% to a more extreme of 55%) were estimated together with measurements for wood by Cave [41]. The calculations are based on the data given in Table 1.

Fig. 4

The elastic modulus of the fibre wall of wood in the longitudinal direction as a function of fibril angle in the S2-wall as affected by different factors; cellulose content in the S2-wall and proportion of S2-wall compared to measurements by Cave [41]. Reference values: cellulose content in S2-wall 50%, proportion of S2-wall 79%.

It is clear that most of the upper boundary of the experimental data (undamaged samples) lies within the regions predicted. Thus it is reasonable to assume that the variations seen in the experimental data are due to such natural variations in cell wall structure. The few experimental values that are higher are, as seen above, considerably higher. It could be questioned if these few outliers in modulus value are correct or if most likely they represent measurements for a wood material with a different fibril angle.

5 Properties in the transverse direction

When it comes to the properties in the cross direction to the fibre length data are extremely scarce and problems with sample preparation and measurement are exceedingly difficult to overcome. Data in the transverse direction of a double fibre wall indicate a behaviour in which earlywood fibres have a higher modulus than latewood ones, the former reaching values around 3 GPa [42,43]. As seen in Fig. 5, the variation in cell-wall properties affects the transverse properties a bit more than those in the longitudinal direction. Still all of the calculations fall above that of the transverse fibre wall measurements. It could here be discussed to what extent the fibre sections have been damaged in the process of isolation or to what extent there is still a lack in precision in the data for the cell wall components. It seems however clear that there are no indications that a higher elastic constant for the cellulose could fit in with these data.

Fig. 5

Transverse fibre-wall elastic modulus as a function of the fibril angle in the S2-wall as affected by different factors; cellulose content in the S2-wall, proportion of S2-wall and the fibril angle of the S1-wall. Measured values from Bergander and Salmén [43]. Reference values: dashed line; cellulose content in S2-wall 50%, proportion of S2-wall 79%, fibril angle in S1-wall 70°.

6 Softening effects

Moisture is well known to act as a softener for the wood polymers. In the range of normal relative humidities at room temperature and around it, only the hemicelluloses are affected by moisture to an appreciable extent. Modelling of this softening considering the measured stiffness of moist glucomannan reported in Table 1 is given in Fig. 6. It is interesting to note that the degree of softening is excellently modelled both in the fibre direction as well as in the transverse direction. In the latter case, the actual value of the modulus is overestimated, as previously mentioned, but as the modulus is given on a logarithmic scale, it is clearly seen that shifting the data up to the calculated ones gives a perfect match with the softening magnitude. It is also apparent from the magnitude in the modulus drop that the amorphous wood polymers play a more dominating role for the fibre properties in the transverse direction than they do in the longitudinal one.

Fig. 6

Modulus of the fibre wall of wood in longitudinal and transverse direction at dry (rigid) and moist (soft) conditions of the hemicelluloses. Measurements refer to data for softwood [44].

One also has to consider what happens with the parts of the cellulose fibrils that are accessible to moisture. At least 20% of the volume of the cellulose in a wood fibre will be absorbing moisture. If this means that the volume of material that softens when the relative humidity increases should be considered to be higher or if these parts of the cellulose still are restricted to such an extent that they will not contribute to the softening may be questioned. As seen in Fig. 7, a softening of the amorphous parts of the cellulose, modelled as if 20% of the cellulose had the same properties as the hemicelluloses, would imply a substantial decrease in the degree of softening of the wood fibre for small fibril angles in the longitudinal direction. In the transverse direction, the effect is much smaller. Compared to the measurements on wood (Fig. 7), it is however uncertain if the amorphous cellulose will behave as a softened amorphous component in this humidity range, as a larger drop in elasticity would have been expected for the measurements. However this should be checked on fibres or wood material having a fibril angle close to zero, where the softening effect is estimated to be quite large.

Fig. 7

Influence of a softening due to a relative humidity increase (20% of the cellulose is considered here to be amorphous and to behave in the same way as the hemicelluloses).

7 Conclusions

Based on the comparisons of experimental data on elastic fibre and wood properties with such data calculated from the properties of the wood polymers, it seems evident that the stiffness of the cellulose fibrils is well described by the value of 134 GPa. It must also be concluded that it is highly likely that all of the wood polymers are orthotropic in nature, in relation to the low elastic properties exhibited for the fibre in its transverse direction.

It is also most likely that the water-absorbing disordered regions of the cellulose does not fully soften it the range of relative humidity rates at room temperature, thus the property changes only reflects the softening of the hemicelluloses.

It may also be concluded that the amorphous wood polymers play a more dominating role for the transverse fibre properties than they do for longitudinal fibre properties.


Bibliographie

[1] U. Molin Pulp strength properties – Influence of carbohydrate composition, molar mass and crystalline structure, PhD thesis, Pulp and Paper chemistry and technology, KTH, Stockholm, 2002

[2] E.-L. Hult; P.T. Larsson; T. Iversen Cellulose fibril aggregation – an inherent property of kraft pulps, Polymer, Volume 42 (2001), pp. 3309-3314

[3] A. Bergander; L. Salmén Cell wall properties and their effects on the mechanical properties of fibers, J. Mater. Sci., Volume 37 (2002) no. 1, pp. 151-156

[4] M. Åkerholm; L. Salmén The mechanical role of lignin in the cell wall of spruce tracheids, Holzforschung, Volume 57 (2003) no. 5, pp. 459-465

[5] P.T. Larsson; K. Wickholm; T. Iversen Structural elements in cellulose I, 10th Int. Symp. on Wood and Pulping Chemistry, Yokohama, Japan, Japan TAPPI, 1999, pp. 602-606

[6] I. Duchesne; G. Daniel Changes in surface ultrastructure of Norway spruce fibres during kraft pulping – visualisation by field emission–SEM, Nord. Pulp Pap. Res. J., Volume 15 (2000) no. 1, pp. 54-61

[7] D. Fengel Ultrastructural behavior of cell wall polysaccharides, Tappi, Volume 53 (1970) no. 3, pp. 497-503

[8] J. Fahlén; L. Salmén Cross-sectional structure of the secondary wall of wood fibers as affected by processing, J. Mater. Sci., Volume 38 (2003) no. 1, pp. 119-126

[9] R.H. Atalla; J.M. Hackney; I. Uhlin; N.S. Thompson Hemicelluloses as structure regulators in the aggregation of native cellulose, Int. J. Biol. Macromol., Volume 15 (1993), pp. 109-112

[10] L. Salmén Structure–property relations for wood; from cell-wall polymeric arrangement to the macroscopic behavior (H.-C. Spatz; T. Speck, eds.), Plant Biomechanics 2000, George Thieme Verlag, Stuttgart, Germany, 2000, pp. 452-462

[11] M. Åkerholm; L. Salmén Interactions between wood polymers studied by dynamic FT-IR spectroscopy, Polymer, Volume 42 (2001) no. 3, pp. 963-969

[12] N. Terashima; K. Fukushima; L.-F. He; K. Takabe Comprehensive model of the lignified plant cell wall (H.G. Jung; D.H. Buxton; R.D. Hatfield; J. Ralph, eds.), Forage cell wall structure and digestibility, ASA-CSSA-SSSA, Madison, WI, USA, 1993, pp. 247-270

[13] L. Salmén; A.-M. Olsson Interaction between hemicelluloses, lignin and cellulose: structure-property relationships, J. Pulp Pap. Sci., Volume 24 (1998) no. 3, pp. 99-103

[14] U.P. Agarwal; R.H. Atalla In-situ Raman microprobe studies of plant cell walls: macromolecular organization and compositional variability in the secondary wall of Picea mariana (Mill.) B.S.P., Planta, Volume 169 (1986), pp. 325-332

[15] R.H. Atalla; U.P. Agarwal Raman microprobe evidence for lignin orientation in the cell walls of native woody tissue, Science, Volume 227 (1985), pp. 636-639

[16] M. Åkerholm Ultrastructural aspects of pulp fibers as studied by dynamic FT-IR spectroscopy, PhD thesis, Fibre and Polymer Technology, KTH, Stockholm, 2003

[17] A.J. Kerr; D.A.I. Goring The ultrastructural arrangements of the wood cell wall, Cellul. Chem. Technol., Volume 9 (1975), pp. 536-573

[18] J.D. Boyd An anatomical explanation for viscoelastic and mechanosorptive creep in wood, and effects of loading rate on strength (P. Baas, ed.), New perspective in wood anatomy, Martinus Nijhoff/Dr W. Junk Pub., La Hague, 1982, pp. 171-222

[19] I. Duchesne; K. Takabe; G. Daniel Ultrastructural localisation of glucomannan in kraft pulp fibres, Holzforschung, Volume 57 (2003) no. 1, pp. 62-68

[20] J.M. Willis; W.Q. Yean; D.A.I. Goring Molecular weights of lignosulphonate and carbohydrate leached from sulphite chemimechanical pulp, J. Wood Chem. Technol., Volume 7 (1987) no. 2, pp. 259-268

[21] B.D. Favis; W.Q. Yean; D.A.I. Goring Molecular weight of lignin fractions leached from unbleached kraft pulp fibers, J. Wood Chem. Technol., Volume 4 (1984) no. 3, pp. 313-320

[22] M. Häggkvist; D. Solberg; L. Wågberg; L. Ödberg The influence of two wet strength agents on pore size swelling of pulp fibers and on tensile strength properties, Nord. Pulp Pap. Res. J., Volume 13 (1998) no. 4, pp. 292-298

[23] J. Berthold; L. Salmén Effects of mechanical and chemical treatments on the pore-size distribution in wood pulps examined by Inverse Size Exclusion Chromatography (ISEC), J. Pulp Pap. Sci., Volume 23 (1997) no. 6, p. J245-J253

[24] J. E Stone; A.M. Scallan A structural model for the cell wall of water swollen wood pulp fibres based on their accessibility to macromolecules, Cellul. Chem. Technol., Volume 2 (1968), pp. 343-358

[25] D.S. Kaplan Structure–property relationships in copolymers to composites: molecular interpretation of the glass-transition phenomenon, J. Appl. Polym. Sci., Volume 20 (1976), pp. 2615-2629

[26] C. Schick; E. Donth Characteristic length of glass transition; experimental evidence, Physica Scripta, Volume 43 (1981), pp. 423-429

[27] K. Tashiro; M. Kobayashi Theoretical evaluation of three-dimensional elastic constants of native and regenerated celluloses: role of hydrogen bonds, Polymer, Volume 32 (1991) no. 8, pp. 1516-1526

[28] P.P. Gillis Effect of hydrogen bonds on the axial stiffness of crystalline native cellulose, J. Polym. Sci. A, Volume 2 (1969) no. 7, pp. 783-794

[29] L.M. Kroon-Batenburg; J. Kroon; M.G. Norholt Chain modulus and intramolecular hydrogen bonding in native and regenerated cellulose fibers, Polym. Commun., Volume 27 (1986), pp. 290-292

[30] W.J. Cousins Young's modulus of hemicellulose as related to moisture content, Wood Sci. Technol., Volume 12 (1978), pp. 161-167

[31] A.-M. Olsson, STFI, unpublished results, 2003

[32] C.Y. Liang; K.H. Bassett; E.A. McGinnes; R.H. Marchessault Infrared spectra of crystalline polysaccharides. VII. Thin wood sections, Tappi, Volume 43 (1960) no. 12, pp. 1017-1024

[33] K. Keegstra; K.W. Talmadge; W.D. Bauer; P. Albersheim The structure of plant cell walls. III. A model of the walls of suspension-cultivated sycamore cells based on the interconnections of the macromolecular components, Plant Physiol., Volume 51 (1973), pp. 188-196

[34] B. Monties Biological variability of lignification and mechanical behaviour of wood plants, 2nd Int. Conf. of the European Society for Wood Mechanics, STFI, Stockholm, Sweden, 2003, pp. 1-13

[35] W.J. Cousins Elastic modulus of lignin as related to moisture content, Wood Sci. Technol., Volume 10 (1976), pp. 9-17

[36] W.J. Cousins; R.W. Armstrong; W.H. Robinson Young's modulus of lignin from a continuous indentation test, J. Mater. Sci., Volume 10 (1975), pp. 1655-1658

[37] L. Salmén; A. de Ruvo A model for the prediction of fiber elasticity, Wood Fiber Sci., Volume 17 (1985) no. 3, pp. 336-350

[38] H. Saiki Proportion of component layers in tracheid wall of early wood and late wood of some conifers, J. Jpn Wood Res. Soc., Volume 16 (1970) no. 5, pp. 244-249

[39] S.E. Anagnost; R.E. Mark; R.B. Hanna Utilization of soft rot cavity formation as a tool for understanding the relation between microfibril angle and mechanical properties of cellulose fibers (R. Perkins, ed.), Mechanics of cellulosic materials 1999, ASME, Blacksburg, Virginia, 1999, pp. 43-63

[40] D.H. Page; F. El-Hosseiny; K. Winkler; A.P.S. Lancaster Elastic modulus of single wood pulp fibers, Tappi J., Volume 60 (1977) no. 4, pp. 114-117

[41] I.D. Cave The longitudinal Young's modulus of Pinus radiata, Wood Sci. Technol., Volume 3 (1969) no. 1, pp. 40-48

[42] A. Bergander; L. Salmén Variations in transverse fibre wall properties: relations between elastic properties and structure, Holzforschung, Volume 54 (2000) no. 6, pp. 655-661

[43] A. Bergander; L. Salmén The transverse elastic modulus of the native wood fibre wall, J. Pulp Pap. Sci., Volume 26 (2000) no. 6, pp. 234-238

[44] F.P. Kollman; W. Côté Solid Wood, Principles of Wood Science and Technology, vol. 1, Springer, New York, 1968


To Bernard Monties in recognition of his enthusiastic and inspiring work leading us to a much-improved concept of the lignin structure and role of lignin in plant cell walls enhancing our understanding of the function of the different cell wall constituents in plant fibres.

Commentaires - Politique


Ces articles pourraient vous intéresser

Formation of macromolecular lignin in ginkgo xylem cell walls as observed by field emission scanning electron microscopy

Noritsugu Terashima; Tatsuya Awano; Keiji Takabe; ...

C. R. Biol (2004)


Lignification and tension wood

Gilles Pilate; Brigitte Chabbert; Bernard Cathala; ...

C. R. Biol (2004)


Oxalic acid: a microbial metabolite of interest for the pulping industry

Valérie Meyer-Pinson; Katia Ruel; François Gaudard; ...

C. R. Biol (2004)