You are on page 1of 94

Cellulose degradation in pulp

fibers studied as changes in molar


mass distributions

Rickard Berggren

Doctoral Thesis

Royal Institute of Technology


Department of Fibre and Polymer Technology
Division of Wood Chemistry and Pulp Technology
Stockholm 2003

Cellulose degradation in pulp fibers


studied as changes in molar mass
distributions

Rickard Berggren
2003
Doctoral Thesis
Supervisor: Assoc. Prof. Mikael Lindstrm, Swedish Pulp
and Paper Research Institute
This work has been performed at the Swedish Pulp and
Paper Research Institute

Royal Institute of Technology


Department of Fibre and Polymer
Technology
Division of Wood Chemistry
and Pulp Technology

Swedish Pulp and Paper Research Institute

Akademisk avhandling som med tillstnd av Kungliga Tekniska Hgskolan i Stockholm


framlgges till offentlig granskning fr avlggande av teknologie doktorsexamen fredagen
den 25 april 2003 klockan 14.00 i STFI-salen, Drottning Kristinas vg 61, Stockholm.
Avhandlingen frsvaras p svenska.

Rickard Berggren
Stockholm 2003

En sjman ber inte om medvind,


han lr sig segla.
G. Lindborg
If you're out there all alone
and you don't know where to go to,
come and take a trip with me to Future World
K. Hansen

Cellulose degradation in pulp fibers studied as


changes in molar mass distributions
Rickard Berggren, Royal Institute of Technology, Department of Fibre and Polymer
Technology, Stockholm, Sweden
Abstract
In this thesis, size-exclusion chromatography (SEC) of wood polymers dissolved in lithium
chloride/N,N-dimethylacetamide (LiCl/DMAc) has been used to characterize the molar
mass distributions (MMD) of wood polymers in pulp fibers after chemical degradation.
Characterization of birch kraft pulps subjected to ozone degradation and acid hydrolysis,
respectively, rendered different changes in the MMD. Ozone degradation resulted in large
redistributions of the original MMD, observed as the development of a distinct fraction of
cellulose with intermediate molar mass. Acid hydrolysis resulted in minor changes of the
original MMD compared to ozonation. Fibers subjected to acid hydrolysis were
considerably weaker than ozonated fibers. These results indicated that there are differences
in how the two chemicals degrade the fiber.
The solubility of softwood kraft pulp fibers was enhanced by derivatization of the fiber
polymers with ethyl-isocyanate during simultaneous dissolution in LiCl/DMAc. The
derivatization made it possible to achieve reliable estimations of the MMD, and hence
molar masses, of softwood kraft pulps. The derivatization procedure made it possible to
dissolve 90 % of softwood kraft pulps with kappa numbers over 50.
Severe alkaline degradation of birch and Norway spruce wood chips was studied both by
varying the pulping time and by varying the initial alkali concentration. Differences were
found in the MMD of the two fiber types, and the alkaline degradation was found to affect
polymers in the entire MMD.
Multi-angular laser light scattering (MALLS) was used as a detection technique with SEC
on cellulosic samples. The MMD and average molar masses obtained through directstandard calibration with commercial standards were compared with MMD and molar
masses as obtained by MALLS-detection. Large discrepancies were found, and two
methods of correcting for these discrepancies were developed.
Theoretical simulations of polymer degradation were performed. Random, or
homogeneous degradation was used as a model for alkaline cellulose chain scission, and a
resemblance with experimental data was observed. End-wise depolymerization of cellulose
was also simulated and the results are discussed in the light of experimentally observed
MMD.
Keywords: cellulose, kraft pulp, birch, spruce, ozonation, acid hydrolysis, degradation, MMD, sizeexclusion chromatography, light scattering, molar mass, chain scission

Sammanfattning
I denna avhandling har storlekskromatografi, size-exclusion chromatography (SEC), av
vedpolymerer upplsta i litiumklorid/N,N-dimetylacetamid (LiCl/DMAc) anvnts fr att
karakterisera molekylviktsfrdelningarna av vedpolymerer efter kemisk nedbrytning.
Karakterisering av bjrksulfatmassa nedbruten genom ozonbehandling eller sur hydrolys
resulterade i olika frndringar i molekylviktsfrdelningar. Ozonbehandling resulterade i
stora frndringar av den ursprungliga molekylviktsfrdelningen, observerat som bildandet
av en tydlig cellulosafrdelning med intermedir molekylvikt. Sur hydrolys resulterade i
mindre frndringar av den ursprungliga frdelningen, jmfrt med ozonbehandlingen.
Surt hydrolyserade fibrer var mycket svagare n de ozonbehandlade fibrerna och resultaten
visar att det finns skillnader i hur de tv kemikalierna bryter ner fibern.
Lsligheten av barrsulfatmassa kades genom derivatisering av polymererna med etylisocyanat under samtidig upplsning i 8 % LiCl/DMAc. Derivatiseringen gjorde det
mjligt att gra tillfrlitliga uppskattningar p molekylviktsfrdelningarna, och drmed
molekylvikterna, av polymererna i barrsulfatmassa. Derivatiseringsfrfarandet gjorde det
ocks mjligt att lsa upp 90 % av barrsulfatmassa med kappa-tal ver 50.
Omfattande alkalisk nedbrytning av bjrkflis och granflis studerades genom en variation
av antingen koktid eller initial alkalihalt. Skillnader observerades i molekylviktsfrdelningarna hos de tv fibertyperna, och den alkaliska nedbrytningen pverkade hela
molekylviktsfrdelningarna.
Ljusspridning (MALLS) anvndes som en detektionsteknik tillsammans med SEC av
vedpolymerer. Molekylviktsfrdelningarna och medelmolekylvikterna, mtta antingen
med hjlp av kalibrering med standarder eller med hjlp av MALLS, jmfrdes. Stora
skillnader i dessa resultat observerades, och tv metoder att korrigera fr dessa skillnader
utvecklades.
Teoretiska simuleringar av nedbrytning av vedpolymererna genomfrdes. Slumpmssig
eller homogen nedbrytning anvndes som modell fr alkalisk hydrolys av
polymerkedjorna, och likheter med experimentella data observerades. Nedbrytning av
enbart ndgrupperna (peeling) simulerades ocks, och resultaten diskuteras i frhllande
till experimentella molekylviktsfrdelningar.

List of publications

This thesis is a summary of the following papers, which are referred to in the text by their
Roman numerals:
I.

Fiber strength in relation to molecular mass distribution of hardwood


kraft pulp. Degradation by ozone and acid hydrolysis.
Rickard Berggren, Fredrik Berthold, Elisabeth Sjholm, Mikael Lindstrm
Nord. Pulp. Pap. Res. J., 16, 4, 333-338, 2001.

II.

Dissolution of softwood kraft pulps by direct derivatisation in


LiCl/DMAc
Fredrik Berthold, Kristina Gustafsson, Rickard Berggren, Elisabeth Sjholm,
Mikael Lindstrm
Submitted to Journal of Applied Polymer Science.

III.

Alkaline degradation of birch and spruce: influence of degradation


conditions on molecular mass distributions and fibre strength
Rickard Berggren, Ulrika Molin, Fredrik Berthold, Helena Lennholm, Mikael
Lindstrm
Carbohydr. Polym., 51, 3, 255-264, 2003.

IV.

Improved methods to evaluate the molar mass distributions of cellulose


in kraft pulp
Rickard Berggren, Fredrik Berthold, Elisabeth Sjholm, Mikael Lindstrm
J. Appl. Polym. Sci., 88, 5, 1170-1179, 2003.

V.

Comparison between theoretical simulations and observed degradation


patterns of cellulose in fiber cell walls:
Alkaline degradation
Rickard Berggren, Fredrik Berthold, Mikael Lindstrm
Manuscript.

Table of contents
1.

INTRODUCTION.......................................................................................................1

2.

WOOD: STRUCTURE AND CONSTITUENTS .....................................................3


2.1
TRACHEID FIBERS ..................................................................................................3
2.2
ULTRASTRUCTURAL ARRANGEMENT OF WOOD POLYMERS ....................................3
2.3
WOOD FIBER CONSTITUENTS ..................................................................................4
2.3.1.
Cellulose........................................................................................................4
2.3.2.
Hemicellulose................................................................................................5
2.3.3.
Lignin ............................................................................................................6

3.

PULPING AND WOOD POLYMER DEGRADATION.........................................7


3.1
3.2
3.3

4.

MOLAR MASSES OF POLYMERS ......................................................................11


4.1
4.2
4.3

5.

DELIGNIFICATION ..................................................................................................7
ALKALINE DEGRADATION OF POLYSACCHARIDES ..................................................7
ACID HYDROLYSIS .................................................................................................8

INTRINSIC VISCOSITY ...........................................................................................12


LIGHT SCATTERING ..............................................................................................12
SIZE-EXCLUSION CHROMATOGRAPHY ..................................................................13

SIZE-EXCLUSION CHROMATOGRAPHY OF CELLULOSE ........................15


5.1
CELLULOSE SOLVENTS .........................................................................................15
5.1.1.
The LiCl/DMAc solvent system ...................................................................16
5.1.2.
Interpretation of a molar mass distribution ................................................16
5.2
CALIBRATION PRINCIPLES ....................................................................................19

6.

OBJECTIVE AND APPROACHES........................................................................21

7.

MATERIALS AND METHODS..............................................................................23


7.1
PRODUCTION OF WOOD POLYMER BEADS .............................................................23
7.2
DEGRADATION .....................................................................................................23
7.2.1.
Ozonation (paper I).....................................................................................23
7.2.2.
Acid hydrolysis (papers I and IV)................................................................23
7.2.3.
Alkaline degradation (paper III) .................................................................24
7.2.4.
Soda anthraquinone degradation of WPB (paper IV) .................................24
7.2.5.
Alkaline degradation of cotton linters (paper V).........................................25

Alkaline pulping (paper V) ..........................................................................25


7.2.6.
7.3
SIZE-EXCLUSION CHROMATOGRAPHY ..................................................................25
7.3.1.
Dissolution of underivatized samples..........................................................25
7.3.2.
Dissolution of derivatized samples..............................................................26
7.3.3.
Chromatography .........................................................................................26
7.3.4.
Multi-Angular Laser Light Scattering .........................................................27
7.3.5.
Additional analyses .....................................................................................27
7.4
THEORETICAL SIMULATION OF POLYMER DEGRADATION .....................................28
7.4.1.
Description of the simulation model............................................................28
8.

RESULTS AND DISCUSSION................................................................................35


8.1
DEGRADATION PATTERNS OBSERVED DURING ACID HYDROLYSIS AND OZONATION
(PAPER I)..........................................................................................................................35
8.2
IMPROVING THE SOLUBILITY OF SOFTWOOD KRAFT PULPS IN LICL/DMAC (PAPER
II)
.............................................................................................................................39
8.2.1.
Derivatization of cellulose ..........................................................................39
8.2.2.
Influence of derivatization on the MMD and molar mass ...........................41
8.2.3.
Reproducibility, degree of dissolution and substitution ..............................43
8.3
DEGRADATION PATTERNS OBSERVED DURING ALKALINE DEGRADATION OF
SOFTWOOD AND HARDWOOD (PAPER III) .........................................................................44
8.4
IMPROVING THE ACCURACY IN THE DETERMINATION OF MOLAR MASS OF
CELLULOSE IN KRAFT PULPS (PAPER IV) ..........................................................................49
8.4.1.
Method I: Correction of the molar mass of cellulose ..................................51
8.4.2.
Method II: Cellulose-equivalent molar masses of pullulan standards ........52
8.4.3.
The applicability of the improved evaluation methods................................54
8.5
SIMULATION OF CELLULOSE DEGRADATION (PAPER V)........................................56
8.5.1.
Validation of the simulation model..............................................................56
8.5.2.
Random degradation ...................................................................................57
8.5.3.
End-wise depolymerization .........................................................................58
8.5.4.
Linear combinations of chain scission cases...............................................59
8.5.5.
Peeling of cotton linters ..............................................................................62
8.5.6.
Limitations of the simulation study .............................................................64

9.

CONCLUSIONS .......................................................................................................65

10.

ACKNOWLEDGEMENTS..................................................................................67

11.

REFERENCES......................................................................................................69

APPENDIX I. ....................................................................................................................79
APPENDIX II....................................................................................................................81

1. Introduction
Kraft pulping and subsequent bleaching is one of the most versatile processes for
producing pulp to be used in paper and board products. The resulting kraft pulp fibers have
superior strength properties to fibers produced by using other pulping techniques, such as
mechanical pulping or sulfite pulping.
It is central to measure the effects of the chemical reactions occurring during the pulping
and bleaching of wood fibers in trying to obtain the optimal product. It is also central for
an understanding of the underlying reasons for the deterioration of the strength properties
of the fibers.
The fiber consists mainly of polymeric molecules of varying type and size. One of the
fundamental properties of a polymer is its molar mass, which affects the strength
properties of the polymer and the fiber. The decrease in molar mass of the wood polymers
during chemical degradation is normally estimated by measuring the intrinsic viscosity of
dissolved pulp fibers. The intrinsic viscosity is however a rather crude measurement, as it
only provides one average value of the degree of polymerization of the wood polymer
mixture. It does not give any information about which molecules have been degraded and
to what extent.
A more suitable technique to better describe the degradation of the wood polymers is sizeexclusion chromatography (SEC), which provides the entire molar mass distribution of the
wood polymers: lignin, hemicellulose and cellulose. The technique also provides the molar
masses of the wood polymers.
This thesis deals with the use of SEC of wood polymers in the solvent lithium
chloride/N,N-dimethylacetamide as a tool to measure and understand the depolymerization
of cellulose in pulp fibers during chemical degradation.

2. Wood: structure and constituents


2.1 Tracheid fibers
Wood consists of different fiber types. Softwood consists mainly of tracheids and ray cells,
while hardwoods consist of vessel elements, fiber tracheids or libriform cells and ray cells
(Core et al. 1976; Sjstrm 1993d). Softwood tracheids are normally the subject of interest
in any model of the fiber cell wall. The length of Scandinavian softwood tracheids is about
3 mm and the width is about 30 m. Hardwood (birch) tracheids are shorter, about 1 mm,
and with similar widths as the softwood tracheids. The width of the tracheids and the
thickness of the cell walls vary with growth season.
A number of theoretical models have been proposed for these fiber types, and several
models have been presented to describe the detailed structure of the wood fiber (Fengel
1970; Scallan 1974; Fengel and Wegener 1984c; Sell and Zimmermann 1993; Brndstrm
2002). The different models agree that the cell wall consists of different layers, all with
different chemical compositions, thicknesses and functions in the wood, see figure 1. The
layers of the tracheids are called, going from the outside to the inside: the primary wall (P),
the secondary wall (S) and finally the warty layer, which is located on the inside of the
cell. The middle lamella surrounds the tracheids and is enriched in lignin. The primary cell
wall contains cellulose, hemicellulose, pectin, protein and lignin. The secondary wall is
divided into three sub-layers numbered from the outside, S1, S2 and S3, all with different
structural arrangements of cellulose in the layers. The S2-layer is the largest part of the cell
wall, and it is suggested to have the greatest impact on the chemical and physical
properties of the fiber. The main components of the S2-layer are cellulose, hemicellulose
and lignin. The composition of the warty layer is complex and varies with species
(Sjstrm 1993e).

2.2 Ultrastructural arrangement of wood polymers


The wood polymers form complex structures in the living tree, and this affects the
reactivity of the polymers during delignification. The matrix polymers, lignin and
hemicellulose, surround the cellulose (Fengel 1971), which is in turn arranged in more or
less crystalline regions, called fibrils, elementary fibrils, or microfibrils. The arrangement
of these fibrils in the S2-layer of the wood has been proposed to be either concentric (Kerr
and Goring 1975) or radial (Sell and Zimmermann 1993), and the topic is still under
debate.

Figure 1. Cell wall model of a Norway spruce tracheid from mature latewood.
Reproduced with the kind permission from Brndstrm (2002).
The cellulose fibrils are assumed to be largely crystalline, and a decreasing crystallinity is
proposed as a result of various types of surface imperfection (Rowland and Roberts 1972).
Recent studies using solid-state NMR have increased the knowledge regarding the
crystalline character of cellulose, and it has been shown that the cellulose fibrils form
aggregates in the native wood (Wickholm 2001). These aggregates have been shown to
increase in size with increasing temperature during kraft pulping (Hult 2001; Fahln 2002),
with a possible effect on the strength properties of the fibers (Molin 2002).

2.3 Wood fiber constituents


2.3.1.

Cellulose

Cellulose is the most important constituent of the wood cell wall, as it is the main
component in both the native wood fiber and the processed pulp. The cellulose fraction of
native wood is about 40 % (on dry weight), depending on the species. Cellulose is a
homopolysaccharide on the molecular level, composed solely of -D-glucopyranose units,
joined together by (14)- glycosidic bonds (Sjstrm 1993a) to form cellobiose units, the
smallest repeating unit in cellulose (figure 2). The degree of polymerization (DP) of
cellulose present in the living tree is unknown, but the size of the native molecule is often
stated to be 5000-10000 glucopyranose units (Fengel and Wegener 1984d).

OH
HO
HO

OH

O HO

OH

OH

OH

Figure 2. Chemical structure of cellobiose.


Cellulose is the load-bearing element of the pulp fiber, and chemical degradation of the
cellulose results in a fiber with inferior strength properties (Gurnagul et al. 1992; Page
1994).

2.3.2.

Hemicellulose

A number of different molecules present in wood are called hemicelluloses. The word
hemicellulose first appeared in 1891 (Schulze, 1891), and its origin is based on the
assumption that these polysaccharides were precursors to cellulose (Fengel and Wegener
1984a), but the assumption has proven to be wrong (Sjstrm 1993b). The compositions of
the hemicelluloses in hardwoods and softwoods differ slightly. In hardwood, such as birch,
the main hemicellulose is O-acetyl-4-O-methylglucuronoxylan, called glucuronoxylan or
simply xylan. The content of xylan in hardwood wood fibers is about 15-30%, depending
on the species (Sjstrm 1993c).
In softwood, the main hemicelluloses are O-acetyl-galactoglucomannan and arabino-(4-Omethylglucurono)xylan, simply called glucomannan and xylan, respectively.

O
HO

OH

O HO

O
O
O O
OH

O
HO
MeO
HO2C

OH
O

HO

O
OH

HO
O

OH
O

OH

Figure 3. Principal composition of arabinoglucuronoxylan. R denotes extension of


polymeric chain.

The content of glucomannan and xylan in softwood is 20% and 10%, respectively. The
hemicellulose polymers in wood are much smaller in size than the cellulose, as the
hemicelluloses have DPs of 100-200 (Sjstrm 1993f).

2.3.3.

Lignin

Lignin is a complex three-dimensional molecule that gives the wood fiber rigidity. It
enhances the trees resistance towards microorganisms while it acts as a chemical adhesive
joining the fibers together in the stem.
Softwood lignin is composed of the precursor trans-coniferyl alcohol, hardwood lignin is
composed of both trans-coniferyl and trans-sinapyl alcohol, and grass lignin is composed
of trans-p-coumaryl alcohol. These precursors are joined together to form a polymeric
macromolecule (Sarkanen and Ludwig 1971), which is assumed to have an infinite molar
mass in the living tree. An example of a proposed softwood lignin structure is shown in
figure 4.
OH

HO
Lignin-O

OH

Lignin
OMe

OMe

O
HO

OMe
HO

MeO
OH

HO

HO

OH

OH HO

OH
O

OMe

MeO
O

HO
OMe
OH
O
OH

HO

OH

O
OMe

OH

OMe
O

CHO

Figure 4. Model of softwood lignin, adopted from Brunow et al. (1998).

3. Pulping and wood polymer degradation


3.1 Delignification
The main objective of the delignification of wood chips is to separate the pulp fibers. This
separation may be performed by several different methods, either mechanically or
chemically. Chemical delignification may be performed by adding a number of reagents
such as alkali, sulfite, anthraquinone, polysulfide and hydrogen sulfide to the aqueous
pulping liquor. The results presented in this thesis are mainly related to the properties of
pulp fibers produced by the kraft pulping process and subsequent chemical treatments.
Kraft pulping is the common name for the delignification of wood chips by the action of a
solution of hydroxide and hydrogen sulfide on the wood chips at an elevated temperature.
The process leads to pulps with better strength properties than the sulfite process
(Rydholm 1965) and it is the most common process to produce chemical pulp in Sweden
today (Skogsindustrierna, 2003). The chemical reactions occurring in the wood matrix
during delignification are complicated: hemicellulose is partly degraded and dissolved,
extractives are removed to a large extent, and the lignin is extensively degraded and
dissolved (Gierer 1980).
The kinetics of the delignification is normally divided into three different phases, all
having different effects on the carbohydrates and the lignin (Gierer 1980). The process
suffers from selectivity problems due to the high temperature, but the selectivity has been
improved by implementation of the principles of modified kraft pulping (Teder and Olm
1981).

3.2 Alkaline degradation of polysaccharides


The alkaline conditions under kraft pulping degrade the carbohydrates in the fibers. The
most important reactions of the carbohydrates are end-wise depolymerization (peeling) and
alkaline hydrolysis of the wood polymer chains.
The peeling reaction removes one unit at a time from the reducing end of the cellulose
chain, while a competitive stopping reaction, via a -hydroxycarbonyl elimination that
transforms the end-group of the cellulose chain to a stable D-glucometasaccharinic acid,
hinders further peeling to take place (Fengel and Wegener 1984b). The activation energies
for the peeling and stopping reactions have been estimated to be 103 kJ/mol and 135
kJ/mol, respectively (Haas et al. 1967), which implies that the importance of the stopping
reaction increases with temperature. Franzon and Samuelsson studied un-mercerized
cotton and concluded that in average 65 glucosidic units were peeled off before a stopping

reaction occurred (Franzon and Samuelsson 1957), and this was later confirmed (Lai and
Sarkanen 1967). The corresponding number in mercerized cotton was 40 units. The
peeling reaction is thus assumed to affect the yield of the alkaline pulping process.
The effect of the peeling reactions on the molar mass of cellulose is however small,
compared to the effect of alkaline hydrolysis that decreases the molar mass of the cellulose
by chain scission along the entire chain (Samuelsson et al. 1953; Corbett and Richards
1957). Nevertheless, although the molar mass of cellulose is rapidly reduced, the effect on
the yield of the pure alkaline chain scission is low. The alkaline hydrolysis is more
important at higher temperatures; the activation energy for the reaction has been
determined to be 150 kJ/mol for cotton cellulose (Lai and Sarkanen 1967) and 179 kJ/mol
in the soda-anthraquinone pulping of spruce (Kubes et al. 1983).
The ultrastructure of the cellulosic fibers may also affect the rate of alkaline degradation.
Peeling, stopping and random chain scission by alkaline hydrolysis of fibrous and
amorphous cellulose substrates were studied by Gentile et al. (1987). Both the peeling and
stopping reactions were found to occur more rapidly in the amorphous substrate than in the
fibrous material. It was suggested that alkaline hydrolysis occurs only in the amorphous
substrate at the temperatures studied (60C and 80C).
The hemicelluloses are partly degraded during kraft pulping, and the loss of hemicelluloses
is responsible for most of the total yield loss observed during kraft pulping. The
hemicelluloses are involved in precipitation (Yllner and Enstrm 1956; Yllner and
Enstrm 1957), the formation of hexenuronic acid residues (Teleman et al. 1995),
saponification, swelling, dissolution, peeling and alkaline cleavage reactions (Lai 2001).

3.3 Acid hydrolysis


Acid hydrolysis of cellulose is an interesting reaction as it provides information about the
structure and morphological factors that affect the rate and extent of chemical reactions in
cellulosic samples (Krssig 1993a). Dissolving pulp grades are usually produced under
acidic conditions (Wilkes 2001).
Cellulose may be degraded by acid hydrolysis either heterogeneously or homogeneously.
In this case, homogeneity implies that the entire structure is susceptible for degradation.
Heterogeneous acid hydrolysis of cellulose is performed at low concentration of acids, and
leaves a residue of cellulose not susceptible for degradation. The degradation reaction
proceeds according to at least two kinetically different phases. The first phase is rapid and
is associated with the degradation of easily accessible regions in the cellulose structure.
The reaction rate declines during the second stage of the reaction, until the degree of

polymerization (DP) of the cellulosic residue approaches the leveling-off-degree of


polymerization (LODP). The LODP is defined as the DP at which no further acid
hydrolysis takes place, and is around of 100-300 units (Battista et al. 1956). The rates of
degradation in the different phases have been found to differ by two orders of magnitude
(Durawalla and Shet 1962).
It has been suggested that further degradation of the cellulosic residue with a size close to
the LODP occurs according to an end-attack model (Sharples 1957; Wood et al. 1989),
which has been expanded to apply only to cellulose of the crystalline form cellulose II (Lin
et al. 1993).
Homogenous degradation is performed at a high acid concentration, which is required to
disrupt and degrade the crystalline part of cellulose. Celluloses from different sources have
been showed to react under homogeneous conditions at different rates depending on the
source; wood cellulose degrades twice as quickly as cotton cellulose. This was explained
as being due to the existence of oxidized groups in the cellulose chain of the wood
cellulose (Marchessault and Rnby 1959). The study of the homogeneous degradation of
cellulose in phosphoric acid has led to the development of the Ekenstam-expression, which
is commonly used in kinetic studies of cellulose depolymerization (Ekenstam 1936).
Much work has been devoted to the acid hydrolysis of cellulose and reviews discuss the
details of the reactions (Nevell 1985; Krssig 1993a; Lai 2001).

4. Molar masses of polymers


Molecules of various chain lengths build up polymeric materials, i.e. the material has a
molar mass distribution (MMD). This is especially true for natural polymers such as
cellulose. The MMD of a polymer may be illustrated by calculating molar mass averages
of the distribution. The commonly used averages are defined as
Number average molar mass

n M
n
i

Mn =

(1)

Weight average molar mass

n M
n M

Mw =

2
i

(2)

z-average molar mass

Mz =

n M
n M
i

3
i
2
i

(3)

where ni is the number of molecules having the molar mass Mi.


The width of the MMD is called the polydispersity (PD), and is defined as
PD =

Mw
Mn

(4)

The values of the molar mass averages are always in the order Mz>Mw>Mn, except in the
case of a monodisperse distribution, where PD=1. Pure monodisperse distributions are
seldom observed for polymers other than proteins.
Several methods may be applied to determine the molar mass of a polymer, and brief
descriptions of some of the techniques applicable to cellulose are given below.

11

4.1 Intrinsic viscosity


The intrinsic viscosity of a dilute solution of a polymer is related to the molar mass of the
polymer (Staudinger and Heuer 1930). The relation is described by the empirical MarkHouwink equation:

= KM a

(5)

where is the intrinsic viscosity of a sample dissolved in a specified solvent and M is the
molar mass. K and a are empirical parameters, where a is a measure of the extension of the
molecule. The value of a lies between 0.5 and 1.0 for random coils, and is over 1 for rigid
rod molecules. The parameters are determined by fractionating the MMD into narrow
fractions, followed by measurements of the molar mass by an absolute method and
determinations of the intrinsic viscosity of the fractions. If the Mark-Houwink parameters
are known for the polymer in the specified solvent, it is possible to calculate the viscosity
average molar mass, Mv, from the molar mass distribution. Mv is defined as
ni M i1+ a
Mv =

ni M i

1/ a

(6)

Normally, viscosity measurements of a pulp sample dissolved in cupriethylenediamine


(CED) are used to estimate the carbohydrate degradation occurring during pulping and
bleaching. The relationship between the intrinsic viscosity () and the degree of
polymerization (DP) of cellulose samples has been formulated as (Evans and Wallis
1989):
DPv

0.9

(7)

= 1.65

4.2 Light scattering


Light scattering is an absolute detection technique that provides the weight average
molecular mass (Mw) of the sample without calibration with external standards according
to the following equation

K c
1
=
+ 2 A2 c
R( ) M w P( )

12

(8)

where the constant K* is defined as


2

dn
4 2 n02
dc
K* =
N A 40

(9)

c is the polymer concentration, R() describes the excess of scattered light from the
polymer at angle , P() is a form factor, A2 is the second virial coefficient (a measurement
of solvent-solute interaction), dn/dc the specific refractive index increment of the polymer
in solution, no the refractive index of the solvent, 0 the wavelength of the incident light in
vacuum and NA the Avogadro number (Wyatt 1993). The measurements can be performed
in batch mode, i.e. by direct measurement of the scattered light at a number of specified
concentrations, from which a Zimm-plot is constructed (Zimm 1948). Measurements can
also be performed by attaching the light scattering detector to a SEC-column, followed by
a concentration-sensitive detector, such as a refractive index or an ultraviolet detector. The
batch mode provides the Mw, the second virial coefficient and radius (root-mean-square
radius) of the sample in solution, while the online mode provides mass and radius
distributions, from which mass averages (Mn, Mw and Mz) are calculated.

4.3 Size-exclusion chromatography


Size-exclusion chromatography (SEC) is a separation technology used to separate
molecules according to their size in solution, and the origin of the technology is attributed
to Porath and Flodin (1959). The sample is dissolved in a solvent and introduced into a
separating column, which is filled with a porous packing material. The size of the pores in
the packing material determines the molecular range over which the column is active. A
constant flow of solvent is applied to the column, and separation takes place by trapping of
the sample molecules within the pores. Smaller molecules are retained for a longer time in
the column, as they are able to penetrate the porous packing material to a greater extent
than larger molecules. Molecules larger than the pores are not retained in the column at all,
and they are eluted first. They are followed by the molecules that are small enough to
penetrate the pores, and finally the smaller molecules are eluted (Yau et al. 1979; Skoog
and Leary 1992).
The technique is used to monitor the molar mass distribution of polymeric samples, and
the molar mass averages Mn, Mw and Mz can be calculated from MMD provided that the
system has been correctly calibrated. The Mv-value may also be calculated from the MMD,
provided that the correct Mark-Houwink parameters of the system are known.

13

5. Size-exclusion chromatography of cellulose


5.1 Cellulose solvents
Size-exclusion chromatography (SEC) of cellulose requires that the sample is dissolved in
a solvent. Unfortunately, high molecular mass cellulose is insoluble in most common
solvents. This is due to the crystallinity, the crystallite size, the crystallite size distribution
and the hydrogen bonds holding the cellulose structure together (Krssig 1993b).
Attempts have therefore been made to dissolve derivatized cellulose and the MMD of
derivatized cellulose has hitherto commonly been determined on cellulose carbamates
obtained through derivatization with phenyl isocyanate (Schroeder and Haigh 1979;
Sundquist and Rantanen 1983; Evans et al. 1989). The sample is derivatized in pyridine or
dimethylsulfoxide (DMSO), precipitated and re-dissolved in tetrahydrofuran (THF), which
is used as the mobile phase in the subsequent SEC-characterization. The precipitation step
can lead to a loss of low molecular mass cellulose (Wood et al. 1986) and it is necessary to
remove the lignin using e.g. chlorite treatment prior to the dissolution. Carbohydrate
degradation is assumed not to occur during this treatment, but it cannot be excluded. The
removal of lignin in combination with the strong UV-absorption of the phenyl isocyanate
group also leads to a loss of information regarding possible lignin-carbohydrate complexes
(LCC) present in the sample.
A solvent, which was earlier used for the measurement of the intrinsic viscosity of underivatized cellulose, is cupraammonium hydroxide (Cuam). The solvent decomposes
during long storage times, and it may, due to its alkalinity, degrade the dissolved cellulose.
Cupriethylenediamine (CED) has now replaced Cuam as a solvent for instrinsic viscosity
measurements, and the CED-solution is more stable than the Cuam-solution (Johnson
1985). Cadmium tris(ethylenediamine) (Cadoxen) may also be used to dissolve cellulose.
Cadoxen is colorless, which facilitates SEC, and cellulose solutions are stable. However,
the solution is time-consuming to prepare, and it contains the toxic compound cadmium
(Brown 1967; Johnson 1985).
A cellulose solvent used on an industrial scale is N-methyl morpholine oxide (NMMO)
(Johnson 1969), which is used to dissolve fibers during the production of Lyocell-fibers.
The solvent has a complex ternary phase diagram, and this means that careful control of
the water content is required to achieve economical solutions of cellulose (Woodings
2001).

15

5.1.1.

The LiCl/DMAc solvent system

A solvent system frequently used to dissolve cellulosic samples is a mixture of lithium


chloride/N,N-dimethylacetamide (LiCl/DMAc). The solvent/sample-system was first
described two decades ago (McCormick and Lichatowich 1979; Turbak et al. 1981). The
solvent/sample-system is colorless and it is thus also suitable for liquid chromatography
coupled to light absorbing detectors. The UV-cutoff is 270 nm, which makes it possible to
study the lignin distribution in kraft pulps. The system is fairly stable, as little or no
degradation of the samples has been reported (Strlic et al. 1998). No simple explanation of
the mechanism of dissolution of cellulose in LiCl/DMAc has been proposed, although
several studies have been performed (Morgenstern and Kammer 1996).
Cotton, sulfite and birch kraft pulps can be dissolved at high degrees of dissolution without
derivatization (Kennedy et al. 1990; Westermark and Gustafsson 1994; Sjholm et al.
1997; Striegel 1997). However, problems have been reported with underivatized softwood
kraft pulps, probably due to gelation of the glucomannan fraction (Hortling et al. 1994;
Sjholm et al. 1997). LiCl/DMAc has also been found to be a good solvent for a number of
cellulose derivatives (McCormick and Lichatowich 1979; McCormick and Callais 1987).
The solvent/sample-system has other drawbacks. One example, in addition to the poor
solubility of softwood kraft pulps, is the high salt concentration, which both increases the
viscosity of the solution and complicates preparative SEC. Another is the corrosive nature
of the solvent, which limits the choice of material in the chromatographic system.
Although there are some drawbacks in the solvent/sample system, LiCl/DMAc is
nevertheless one of the most suitable solvent systems available for cellulosic materials. It
is compatible with common column materials, it is stable and it dissolves birch kraft pulp
and cotton to a satisfactory extent without derivatization.

5.1.2.

Interpretation of a molar mass distribution

A SEC-characterization of unbleached birch kraft pulp generally gives a MMD having two
distinctive distributions (figure 5). A larger, high molecular mass peak represents the
MMD of the cellulose fraction and a smaller, low molecular mass peak represents the
hemicellulose and lignin fractions (Sjholm et al. 2000a)

16

1.6
RI

dw/dlogM

1.2

UV

0.8
0.4
0
3

5
log M

Figure 5. Differential molar mass distribution of a birch kraft pulp. RI and UV denote
refractive index and ultraviolet detection, respectively.
The area under a MMD is always normalized to unity in order to make it possible to
compare the MMDs from injections with different sample concentrations. The SEC
columns may be attached to an ultraviolet detector (UV) and a differential refractive index
detector (RI), connected in series. The results from the UV-detector describe the
distribution of UV-absorbing molecules in the sample, which are related to the lignin
(Westermark and Gustafsson 1994). The RI-detector describes the total mass distribution
of the cell wall polymers in the sample. The combined information from the UV and the
RI-detector can be used to evaluate possible lignin-carbohydrate complexes (LCC) in kraft
pulps (Karlsson and Westermark 1996). The present thesis is mainly concerned with the
MMD as obtained by RI-detection, which corresponds mainly to the carbohydrate fraction
of the pulp fibers.
The result of a SEC characterization of a kraft pulp is normally a plot of the differential
mass fraction plotted versus the logarithm of the molar mass. The ordinate describes the
mass fraction eluting between log M and log M+dlog M. This mass fraction describes the
mass of the polymer, i.e. the number of molecules multiplied by its molar mass.
The significance of the dw/d(logM)-term may be more easily understood if the cumulative
weight fraction of the polymer is plotted versus the molar mass, figure 6. The dw/d(logM)term describes the absolute value of the derivate of the cumulative weight fraction with
respect to the logarithmic molar mass at each point of the cumulative distribution. If that
value is plotted versus log M, figure 5 is obtained which is the normal way of representing
a MMD of a cellulosic material.
17

Weight fraction (%)

100
RI

80

UV

60
40
20
0
3

log M

Figure 6. Cumulative plot of the MMD in figure 5. The arrow denotes order of elution
(larger molecules elutes before smaller molecules).
The MMD may be represented in different ways, depending on the purpose of the
characterization, and it is necessary to be familiar with the relations between these
representations (see section 7.4). One way is to describe the differential mass fraction
eluting between M and M+dM as shown in figure 7. However, this plot is difficult to
interpret in the case of birch kraft pulp polymers.

0.003

dw/dM

RI
UV

0.002

0.001

0
0

250000

500000

750000

1000000

Figure 7. Mass fraction plot of the MMD in figure 5.


Another way is to plot the number fraction of the polymer eluting between M and M+dM
(data not shown). Thus, it is possible to represent the same sample in different ways, all
with very different appearances, but with the same content. The molar masses may be
calculated from any of the distributions using the equations 1-4.

18

5.2 Calibration principles


The signal from a detector in a SEC-system has to be calibrated. The output from a
detector is only a record of the sample concentration versus elution time, and calibration is
required so that the elution time can be related to the molar mass of the eluting sample.
The calibration of the system may be achieved by universal calibration (Grubisic et al.
1967; Yau et al. 1979; Striegel and Timpa 1996), provided that a viscosity detector is
available. The principle behind the calibration is that for any polymer the product of the
limiting viscosity, [], and the molar mass, M, is proportional to the hydrodynamic volume
of an equivalent sphere, Vh, i.e.
[ ]M Vh

(10)

In the ideal case, the molecules are separated in the columns according to their
hydrodynamic volume, which is proportional to the radius of gyration (Rg) raised to the
power of three. The same retention time will therefore be obtained for any two polymers
having the same value of []M, Vh or Rg.
Two molecules with the same hydrodynamic volume will have the same retention time.
This fact is used when the system is calibrated with standards of known molar mass,
known as direct-standard calibration. In this case, a set of standards, not necessarily of the
same polymer as the sample, is used to relate the elution time to the molar mass. It is
assumed that the samples and the standards have the same relation between molar mass
and hydrodynamic volume, and thus between molar mass and elution time. It is therefore
important to choose standards that resemble the samples as much as possible with respect
to chemical structure and conformation, in order to achieve a correct calibration. For
cellulose, the commercially available pullulan standard is often used (Westermark and
Gustafsson 1994; Strlic et al. 1998; Sjholm et al. 2000a; Sjholm et al. 2000c), but the
applicability of the pullulan has recently been questioned (Bikova and Treimanis 2002).
If a light scattering detector is available, the molecular mass of the eluting sample is
obtained in every slice of the chromatogram, and no external calibration is needed,
provided that the dn/dc-value of the solvent/sample system is known (Wyatt 1993). Recent
studies have illustrated the capability of characterizing dissolving pulps in LiCl/DMAc
using light scattering detection (Schelosky et al. 1999; Schult et al. 2002).

19

6. Objective and approaches


The objective of this thesis is to study the effect of chemical treatments on strength
properties of pulp fibers and how these effects relate to changes in molar mass
distributions of the pulp polymers. In order to do this, a new and improved method of
dissolving the samples prior to the characterization by size-exclusion chromatography was
developed. The use of light scattering as detection technique was studied, and a simulation
model was developed in order to elucidate the experimentally observed degradation.

21

7. Materials and Methods


A number of different pulps, both industrial and laboratory-made, were used. Cotton
linters were also used as raw material. See appendix I for descriptive data of all these
samples.

7.1 Production of wood polymer beads


Regenerated cellulose spheres, as described by Lindstrm et al (1999), were produced as a
model system for degradation studies. For a detailed description of the production of these
wood polymer beads (WPB), see stlund (2001). The principle behind the production of
the beads is: fibers are dissolved in 8% LiCl/DMAc, followed by re-precipitation of the
solute by dripping the solution into a mixture of ethanol and water, which results in sphereshaped cellulosic beads. The degree of crystallinity is low in the beads, and the chemical
composition is similar to that of the original pulp (Lindstrm et al. 1999).

7.2 Degradation
Cellulosic samples were degraded by several different chemicals under different conditions
during the course of this work.

7.2.1.

Ozonation (paper I)

Pulp samples were degraded by ozonation at high consistency (40%) at ambient


temperature in a rotating glass vessel connected to a Fischer 5000 ozone generator. Prior to
ozonation, the pH of the samples was adjusted to 3 with sulfuric acid followed by fluffing.
Eight pulp samples were subjected to ozone dosages ranging from 0.1 to 3% (w/w). After
washing with water, the samples were buffered to pH 7.3 with a 0.02 M phosphate buffer.

7.2.2.

Acid hydrolysis (papers I and IV)

Acid degradation was performed by adding 25 ml 2 M HCl per gram pulp in sealed
polyethylene bags, which gave a pulp consistency of 3%. The temperature during
hydrolysis was 41C and the treatment was performed for 3, 6, 10, 14, 17, 25 and 40 h.
After washing, the acid-treated samples were buffered to pH 7.3 with a 0.02 M phosphate
buffer.

23

Wood polymer beads made from an unbleached pulp were also degraded by acid
hydrolysis (paper IV). A sample of 15 mg was degraded with 5 ml 2M hydrochloric acid at
ambient temperature for degradation times of 1h, 3h, and 23h. Degraded beads were
washed repeatedly with deionized water to neutral pH before dissolution.

7.2.3.

Alkaline degradation (paper III)

Hand-sawn spruce chips with minimal damage and hand-sorted laboratory-chipped birch
chips were used to perform extended alkaline degradation under conditions similar to those
prevailing during kraft pulping. The pulps were digested in autoclaves which were heated
in a glycol bath. The liquid:wood ratio was of 8:1 during the experiments. After the cooks,
the pulps were washed with de-ionized water overnight and defibrated.
The degradation conditions in the different cooks were varied by changing the pulping
time or the initial alkali concentration. Chemical charges and reaction times in these
experiments are given in appendix II.
All the spruce chips were impregnated in the cooking liquor by increasing the temperature
from 70C to 170C at a rate of 1oC/min. The temperature was kept at 170C for the rest of
the cooking time. All the birch chips were impregnated at 120C for 30 minutes and then
heated at maximum speed (10 minutes) to 170C and kept there for the rest of the cooking
time. The initial hydrogen sulfide ion concentration was 0.3 M for all the cooks.
The residual alkali in the black liquor was determined by titration with HCl to pH 10.7. To
achieve a correct value, the black liquor was first diluted eight times and the dissolved
lignin and carbonate ions were precipitated with BaCl2. The hydrogen sulfide ion
concentration was determined using a potentiometric titration with AgNO3 modified from
the method of Chiu and Paszner (1975).

7.2.4.

Soda anthraquinone degradation of WPB (paper IV)

Wood polymer beads from a fully bleached pulp were degraded under soda-anthraquinone
(soda-AQ) pulping conditions. For full details of the chemical conditions, see stlund
(2001). Two temperatures were used (130C and 170), the alkali concentrations were
varied between 0.1 and 0.7 M and the AQ concentrations were varied between 0.1% and
9%. The liquid:wood ratio was kept constant at 200:1.

24

7.2.5.

Alkaline degradation of cotton linters (paper V)

Cotton linters were subjected to degradation in alkali. 5 grams of fibers were put in a steel
vessel, and 80 ml 1 M NaOH solution was added. The vessel was heated in a glycol bath,
thermostated to 130C for 1, 2, 4, 5.5 and 7.5 hours. The samples were thereafter washed
repeatedly with water until a washing solution with a neutral pH was obtained.

7.2.6.

Alkaline pulping (paper V)

Hand-sorted spruce chips (4g) were pulped in steel autoclaves. Initial alkali, hydrogen
sulfide and sodium concentrations were 0.4 M, 0.22 M, and 1.7 M respectively. The
liquor:wood ratio was 20:1. The pulping included a heating stage from 70C to 145C,
after which the pulping liquor was removed and replaced with fresh liquor. Thereafter, the
autoclaves were heated to the pulping temperature 170C at which pulping took place for
70, 100, 130 and 200 minutes. Subsequent defibration and washing of the pulps were
carried out according to normal laboratory procedures.

7.3 Size-exclusion chromatography


7.3.1.

Dissolution of underivatized samples

The underivatized birch kraft pulps and cotton linters were dissolved according to a
procedure developed earlier at STFI (Sjholm et al. 1997; Sjholm et al. 2000a; Sjholm et
al. 2000b). Washed pulp samples (15 mg oven dry weight pulp) were activated in 15 ml
deionized water at 4C for one hour. The water was removed by vacuum filtration and 15
ml of methanol was added and removed by vacuum filtration after 30 minutes. The
procedure was repeated twice with methanol and three times with degassed DMAc. A
solution of 8% (w/v) LiCl in DMAc was added and gently stirred under a nitrogen
atmosphere at 4C for five days. The samples were then equilibrated at room temperature
for 30 minutes and diluted with degassed DMAc to sample and LiCl concentrations of 0.05
and 0.5%, respectively. After 2 hours, the samples were de-aggregated (Sjholm et al.
2000b) in a Retsch vibratory ball mill type MM-2 (intensity 70) for 30 minutes and filtered
through a 0.45 m PTFE filter before the chromatographic characterization.
The dissolution procedure for WPB was similar, but performed with a lower amount of
solvents (water, methanol and DMAc) since only 2-5 mg sample were used.

25

7.3.2.

Dissolution of derivatized samples

Kraft pulp samples were simultaneously derivatized and dissolved as follows: 15 mg fibers
were activated for one hour in 15 ml deionized water at 4C. Thereafter, the fibers were
solvent-exchanged 3 times with 15 ml dry DMAc at ambient temperature and placed in a
flat-bottomed glass cylinder. 1.9 ml 8% LiCl/DMAc and 3 mmol of the derivatizing
reagent then added. The samples were left under argon at ambient temperature for 5 days
with mild magnetic stirring. Finally, the samples were diluted to 0.5% LiCl by the addition
of 27.4 ml DMAc and the excess reagent was quenched by adding 500 l dry methanol.
The samples were de-aggregated and filtered in same manner as the underivatized samples
before the chromatographic separation.
After dilution with DMAc, the samples for the determination of the degree of substitution
were precipitated in methanol, washed twice with water:methanol (7:3) and twice with
water. The samples were freeze-dried and the nitrogen content was determined at
Mikrokemi AB, Uppsala, Sweden. Non-dissolved residues were determined
gravimetrically after ultra-centrifugation and washing of the residue with water to remove
salt and traces of DMAc.

7.3.3.

Chromatography

The chromatographic system consisted of a 2690 Separation Module (Waters Corp,


Milford, MA, USA.) equipped with a guard column (Mixed-A 20 m 7.5x50 mm, Polymer
Laboratories, Shropshire, UK) followed by four columns (Mixed-A 20 m, 7.5 x 300 mm,
Polymer Laboratories) connected in series. The mobile phase was 0.5% (w/v) LiCl/DMAc,
the flow rate was 1 ml/min and the injection volume was 200 l. The mobile phase was
filtered through a 0.2 m PTFE inline filter. The separations were performed at 80C. Four
mixtures of narrow pullulan standards with nominal masses of 738 Da, 5.8 kDa, 12.2 kDa ,
23.7 kDa, 48 kDa, 100 kDa, 186 kDa, 380 kDa, 853 kDa, and 1 660 kDa (Polymer
Laboratories) were used for calibration in papers I, III and IV while narrow pullulan
standards with molecular masses of 1660, 380, 48, 5.8, and 0.738 kDa (Polymer
Laboratories) were used to calibrate the chromatographic system in papers II and V.
Detection was performed online by a 2487 dual wavelength absorbance (UV) detector
(Waters Corp.) at 295 nm, followed by a 410 differential refractive index (RI) detector
(Waters Corp.) thermostated at 40 C..
Data analysis were performed using the Millenium 3.05.01 software (Waters Corp.).

26

7.3.4.

Multi-Angular Laser Light Scattering

The configuration of the detectors used for the light scattering study (paper IV) was
slightly different from the configuration used during the other studies. SECcharacterization was performed relative to standards using one detector configuration
(SEC/RI, System I), and by light scattering (SEC/MALLS/RI, System II) using another
detector configuration. Relative and absolute detection were also carried out
simultaneously in series. The detectors for System I were a 2487 dual wavelength (UV)
detector (Waters Corp.) operating at a wavelength of 295 nm followed by a 410
differential refractive index (RI) detector (Waters Corp.) thermostated at 40 C. For the
light scattering detection in System II, the configuration was a multi-angular laser light
scattering detector (MALLS) (DAWN DSP), followed by an Optilab DSP RI detector
(both from Wyatt Technology Corp, Santa Barbara, CA, USA). Both the MALLS and the
Optilab DSP RI detectors were operating at 488 nm. The order of the detectors was
MALLS-UV-Optilab DSP RIWaters RI.
Two separate column sets were used (both obtained from Polymer Laboratories), one for
calibration and one to verify the calibration. The column sets consisted of a guard column
(Mixed-A, 20 m, 7.5 x 50 mm) and four Mixed-A (20 m, 7.5 x 300 mm) columns
connected in series, all thermostated to 80C
The Optilab RI used in system II was thermostated at 40C. The output voltage from the
detector was calibrated to known refractive indexes by injecting six known concentrations
of sodium chloride dissolved in deionized water, and this gave an instrument-specific RIcalibration constant. The light scattering detector was calibrated with toluene and
normalized by injecting 200 l of a 30 kDa narrow polystyrene standard solution having a
concentration of 1 mg/ml in 0.5% LiCl/DMAc.
The detectors used in the DAWN instrument were numbers 4, 5, 6, 7, 8, 10, 12, 14 and 16.
Narrow interference filters to eliminate the fluorescence from any lignin present in the
samples were placed in front of all the detectors. Detectors at higher and lower angles
were omitted due to the low signal-to-noise (S/N) ratio.
Light scattering data were evaluated using the software ASTRA 4.73.04 (Wyatt
Technology Corp).

7.3.5.

Additional analyses

The intrinsic viscosity and kappa number of the cellulosic samples were determined
according to standard methods SCAN-C 15:99 and SCAN-C 1:77, respectively. The

27

carbohydrate compositions were determined using a gas chromatographic method


(Theander and Westerlund 1986). The fiber length and fiber shape of the samples were
determined by image analysis with the STFI-Fibermaster (Karlsson et al. 1999).
Laboratory sheets for the measurement of strength properties were made according to
SCAN-CM 26:99. The fiber strength was measured as dry and rewetted zero-span tensile
index according to ISO 15361. Nitrogen contents of derivatized softwood pulp samples
dissolved in LiCl/DMAc were determined at Mikrokemi AB, Uppsala, Sweden.

7.4 Theoretical simulation of polymer degradation


A common way to monitor polymer degradation is by measuring the average molar mass,
which is then used to calculate the rate constants for the degrading reactions. This was
done in the 1930:s by Ekenstam, who proposed the following relation when studying the
homogeneous degradation of cellulose in phosphoric acid (Ekenstam 1936).
1
1

kt = ln

DP
DP
0
t

(11)

The k is the reaction rate constant, t the reaction time and DP0 and DPt the degrees of
polymerization of the cellulose initially and at time t, respectively.
Tanford (1961) describes other methods of using the average molar masses to estimate the
kinetics of polymer degradation. It is stated by Tanford that cellulose is a special case since
the structure of the reacting material also is affecting the rate of scission of the molecules.
The use of molar mass averages does not reveal any information about the molar mass
distribution of the actual polymer subjected to degradation. The purpose of the theoretical
work in this study was to develop a model that could predict and relate the molar mass
distribution of the wood polymers during the degradation to any type of scission pattern
present in the fiber. The model is described below and the first attempts to use the model
on wood polymers are discussed in the results section.

7.4.1.

Description of the simulation model

The development of the MMD during polymer degradation has been the subject of several
studies (Montroll and Simha 1940; Guiata et al. 1990; Viebke et al. 1996). The two main
approaches in the recent literature are Monte Carlo simulations and a deterministic
equation solving technique. A Monte-Carlo approach has been thoroughly investigated and
developed by Tobita for a variety of polymers and scission cases (Tobita 1995; Tobita

28

2001). An example using a Monte-Carlo approach has also been published modeling
degradation of linear polymers (Emsley and Heywood 1995). In this study the
deterministic equation solving technique developed by Ballauff and Wolf (1981) was
adopted. The same approach has recently been used to study the degradation of nucleic
acids (Tanigawa et al. 1996) and guar galactomannan (Tayal and Khan 2000). The main
reason for choosing this approach was the relative ease with which the system of equations
could be handled and solved.
The following is a description of the approach that was used in this thesis. Normally, a
MMD of a cellulosic polymer is depicted graphically as a plot of the differential mass
fraction dw/d(logM) versus the logarithm of the molar mass, log M (figure 5). The
dw/d(logM)-term describes the mass fraction of the polymer eluting between molar masses
log M and log M+dlogM. This description is achieved by measuring the detector response
W(t) at the elution time t and normalizing the response with respect to the sum of the
detector response over the entire elution time of the distribution according to the equation:
WN (t ) =

W (t )

(12)

W (t )dt
0

The normalized detector response, WN(t), is then calibrated to the elution time via the
calibration constant of the separation system according to:

dw / d (log M ) = WN (t )

dt
d (log M )

(13)

where dt/d(logM) is the inverse of the calibration constant for the system at the elution
time t.
The model considers the rate constant, ki, for the degradation of a molecule of chain length
i, and the rate constant of degradation into two new fragments of lengths j and i-j,
described as ki,j and ki,i-j, respectively. Hence, summarizing all ki,x over all x for a molecule
of length i gives the value of ki which describes the degradation rate for a molecule of
length i.

29

The model assumes that the degradation is first order with respect to the number fraction
of i-mers and the following mass balance for a polymer with degree of polymerization i
holds in the entire system of polymers:
i 1

df i
= k i , j f i + ( k i +1,1 + k i +1,i ) f i +1 + ... + ( k r ,i + k r ,r 1 ) f r
dt
j =1

(14)

where dfi/dt is the change in concentration of the molecule of length i with respect to
simulation time, ki,j are the individual rate constants, fi is the number fraction of polymer of
length i, and r is the highest degree of polymerization of the polymers in the system.
The change in concentration of the molecules in the system with respect to simulation time
is given by the equation 15 where the matrix A is described by equation 16.
dF (t )
= AF (t )
dt

0 k 2,1 + k 2,1

k 2,1
0

0
A = 0

0
0

0
0

(15)

k3,1 + k3, 2 ... k r ,1 + k r ,r 1

...
k3, 2 + k3,1 ...

2
(16)

...
k3, j ...

j =1

0
..
...
r 1

0
0 kr , j
j =1

This matrix contains the rate constants for chain scission satisfying equation 15 over the
entire distribution of chain lengths and it is triangular since it is stipulated that no
recombination of original or degraded fragments can occur.
The vector F(t) in equation 15, is described by:
f (1, t )

f ( 2, t )
F (t ) =
..

f (r , t )

(17)

The F(t)- term thus describes the number fraction of the molecules within the DP-range
from 1 to r in the time-dependent system of degrading molecules, where f(i,t) is the
number fraction of polymers with a DP of i at the simulation or degradation time t .

30

The model handles the scissions of individual linkages between the subunits constituting
the polymer, and a transformation of the differential mass fraction to provide the number
fraction of the polymers is thus necessary. The mass fraction dw/d(logM) that is normally
used to depict the MMD of a polymer, is used to express df/dM, which is the mass fraction
of the polymer eluting between masses M and M+dM.
dw
df
d log M =
dM
d (log M )
dM

(18)

df
d (log M )
dw
1
dw
=
=
dM
dM
d (log M ) M ln 10 d (log M )

(19)

which yields

where M is the mass of the polymer of length i.


The number ratio, f(i,t), of the polymer eluting between the degrees of polymerization, DP,
and DP+dDP at a simulation time t is then obtained by:

f (i, t ) =

1
dw
moi 2 ln10 d (log M )

(20)

where m0 is the mass of the monomer/subunit constituting the polymer. Equation 15 can
now be solved. This can be done by linear algebra, but in our case it is solved using the
built-in ordinary differential equation solver in the software Matlab (The Mathworks Inc.,
Natick, MA, USA) running on a desktop computer. The resulting vector F(t) from the
simulation is transformed to the dw/d(log M) distribution according to equation 20. The
dw/d(log M) distribution is finally plotted versus log M and the molar masses of the
simulated distribution are calculated from this plot.

Chain scission cases


Two basic cases of chain scission, both studied alone and as linear combinations of each
other, have been considered. The purposes of the cases are to simulate either alkaline
hydrolysis or peeling, and both cases describe the scission rate of a glucosidic linkage as a
function of the position of the linkage within a polymer chain of length i. Several other
cases may be studied (Ballauff and Wolf 1981; Tanigawa et al. 1996; Tayal and Khan
2000) but the following two are the most interesting for alkaline pulping.

31

Random scission
In this mode, the rate constant for chain scission is equal over the entire chain length,
irrespective of molar mass of the polymer, i.e. for any i or j,
(21)

ki, j = k

In this case, the diagonal element in the matrix A is given by:


j 1

A j , j =
i =1

Ai , j

(22)

An interpretation of equation 22 is that the polymeric fragments of length i and j-i are
retained within the system after scission, and that the number of molecules susceptible for
degradation in the system increases with simulation time. The modeling of random
scission was intended to simulate alkaline hydrolysis of cellulose.

Peeling or end-wise depolymerization


The rate constants for end-wise depolymerization are given by:
(23)
k i ,i 1 = k , i > 2
k = 0 otherwise

where i is the length of the polymer.


Thus, the simulated end-wise depolymerization reaction removes only one monomeric
residue at a time from the reducing end of the polymer, as a first approximation to the
experimental case. The diagonal element in matrix A is given by:
A1,1 = 0
j 1

A j , j = Ai , j = k

(24)

i=2

since the units peeled off cannot undergo further chain scission. Hence, the number of
molecules in the system susceptible to degradation does not increase during simulation.

32

The turnover of linkages, X, was used as a measure of the number of degraded linkages
during a simulation. The value of X was determined after each time interval of simulation
according to:
r

X =

[ f (i, t ) (i 1)]
i =2

[ f (i, t = 0) (i 1)]

(25)

i =2

since the number of glycosidic linkages in a polymer of length i is equal to i-1.


The theoretical yield of the simulations, Y, was determined according to:

Y=

[ f (i , t ) i m ]
0

i =1

(26)

[ f (i , t = 0 ) i m ]
0

i =1

During the simulations, no consideration was given to the kind of polysaccharide in which
the actual bond was located. All the linkages between the chain units are treated equally in
terms of reactivity even though there are degradation rate differences between cellulose
and different hemicelluloses (Lai 2001).

33

8. Results and discussion


8.1 Degradation patterns observed during acid hydrolysis and
ozonation (paper I)
Ozone has been considered to be one of the bleaching chemicals to be used as and
alternative to chlorine dioxide (van Lierop et al. 1996). Unfortunately, the delignification
reactions taking place during bleaching are accompanied by a concomitant degradation of
cellulose and hemicelluloses, which leads to both a decrease in yield and a deterioration in
fiber properties, e.g. strength. This is also the case for ozone-delignification that also has
been claimed to suffer from low selectivity towards lignin, compared to chlorine dioxide
(van Lierop et al. 1996).
The cellulose degradation during ozonation and acid hydrolysis was studied in paper I on
an unbleached birch kraft pulp and on cotton linters, both subjected to aqueous-phase acid
hydrolysis or ozonation at high consistency. However, this section deals only with the
results from the degradation of the birch pulps.

Zero-span tensile index


(Nm/g)

The viscosity and fiber strength of the samples decreased in different manners with
increasing ozone dosage and with increasing hydrolysis time, as shown in figure 8.
200
Z-pulp

150

A-pulp

100
50
0
400

600

800

1000

1200

Viscosity (ml/g)

Figure 8. Rewetted zero-span tensile index of pulp fibers subjected to acid hydrolysis
(A-pulp, ) or ozonation (Z-pulp, ) plotted versus intrinsic viscosity.
The initial zero-span tensile index of the pulp was 183 Nm/g. Up to an ozone dosage of
1.6% (w/w), the loss in fiber strength was only 5% of the initial strength, while the
intrinsic viscosity dropped from 1160 ml/g to 880 ml/g. Even at the highest ozone dosage,
at an intrinsic viscosity of 510 ml/g, the loss in fiber strength was only 25%. Acid
hydrolysis of the pulp was more detrimental to fiber strength than ozone treatment. The

35

most severely acid-degraded pulp fibers had an intrinsic viscosity of 810 ml/g and 70% of
the initial fiber strength was lost.
The intrinsic viscosity is clearly insufficient for evaluating different cellulose
depolymerization reactions, and similar results have been reported from studies of different
kinds of alkaline pulping (Kubes et al. 1981). The viscosity losses were similar in the aciddegraded and ozone-degraded pulp samples, but the fiber strength of the acid-treated fibers
decreased more than that of the ozone-treated fibers.
The assumption that acid degradation occurs at local deformations, weak points, in the
fiber cell wall (Gurnagul et al. 1992) could explain our observations. The strengths and
viscosity levels shown in figure 8 indicate that there are differences in the homogeneity of
the degrading reactions. The interpretation of the results is that ozone degradation occurs
more evenly over the entire cellulose fiber than acid degradation, resulting in retention of
fiber strength during ozone treatment. If the ozone had preferentially attacked the weak
points in the fiber, the fiber strength would immediately have decreased as was seen in the
case acid-treated pulp samples.
A shape factor of 100% corresponds to a straight fiber, and it has been suggested that a
deformed fiber in general leads to a decrease in zero-span tensile index (Mohlin et al.
1996). Fiber deformation, determined as shape factor, is presented together with the zerospan tensile index in figure 9 for the degraded pulp samples. The shape factor changed
irregularly with increasing ozone dosage. During acid hydrolysis, a continuously
decreasing trend was more evident. Although the strength loss was greater in the aciddegraded pulp than in the ozone-degraded pulp, there were almost no differences in shape
factor between the two treatments. Thus, the reason to the strength differences is not
related to differences in shape factor.

36

Zero-span tensile index


(Nm/g)

200
150
100

Z-pulp

A-pulp

50
0
94.0

94.4

94.8

95.2

95.6

96.0

Shape factor (%)

Figure 9. Rewetted zero-span of pulp fibers subjected to acid hydrolysis (A-pulp, ) or


ozonation (Z-pulp, ) plotted versus shape factor.
Figure 10 shows the MMDs for the ozone-degraded birch pulp samples. All the MMDs
reported in this section are based on refractive index (RI)-detection, unless otherwise are
stated. The larger, high molecular mass peak represents the MMD of the cellulose fraction
and the smaller, low molecular mass peak represents the hemicellulose and lignin fractions
(Sjholm et al. 2000a). The cellulose fraction of the pulp was considerably degraded in the
presence of ozone. From the original peak, with a log peak molecular mass (log Mp) = 6.2,
a part of the pulp cellulose fraction was extensively degraded and a second cellulose peak
having a value of log Mp 5 was observed. Part of the cellulose seems however to be only
slightly degraded, as there is still a peak with log Mp 6 after the treatment with 3%
ozone. The polydispersity, i.e. the ratio of the weight average to the number average
molecular mass (Mw/Mn), of the cellulose fraction increased from 2.2 in the untreated pulp
to 3.9 in the most severely degraded pulp.

37

dw/dlogM
3.0

4.0

5.0

6.0

7.0

log M (relative to pullulan)

dw/dlogM

Figure 10. MMD of birch kraft pulps subjected to ozonation. The arrow denotes
increasing ozone dosage.

3.0

4.0

5.0

6.0

7.0

log M (relative to pullulan)

Figure 11. MMD of birch kraft pulps subjected to acid hydrolysis. The arrow denotes
increasing time of degradation.
Figure 11 shows the MMDs of birch fibers subjected to acid hydrolysis. The degradation
pattern is quite different from that of ozonation (figure 10). The shortest acid treatment
time, i.e. three hours, shifted both the cellulose and hemicellulose distributions towards
lower molecular mass. However, the rate of degradation decreased after this initial
degradation, and only a slight drop in log Mp of the hemicellulose distribution and almost
no shift in the log Mp of the cellulose distribution were observed upon prolonged treatment.
A minor broadening of the cellulose peak, which may indicate the development of an
intermediate peak, was detected as the polydispersity of the cellulose fraction increased
from 2.2 to 2.8.

38

One of the reasons for the differences in degradation patterns may be the lignin content of
the pulp, as lignin can promote the formation of secondary radicals in the course of ozone
delignification (Eriksson and Reitberger 1995; Lind et al. 1997). Under kraft cooking
conditions, lignin is enriched at the surface of fibers (Laine 1996), where it has been
suggested that the ozone reactions take place (Secrist and Singh 1971). It has been
suggested that radicals formed in ozone-lignin reactions contribute to the cellulose
degradation only if the supply of ozone is large enough to propagate radical chain reactions
(Johansson and Lind 1999). The dosage of ozone in the present study was sufficiently high
to achieve significant cellulose degradation, which also is observed as a shift in the MMD
in figure 10. It is thus suggested that the intermediate cellulose distribution in the MMD
represents cellulose degraded by radicals formed in lignin-ozone reactions. As the reaction
front proceeds inwards through the fiber wall according to a shrinking core model (Zhang
et al. 1998), it is plausible to suggest that the degraded cellulose is located in the external
regions of the pulp fiber. The formation of two distinct cellulose distributions during
ozonation of a unbleached birch kraft pulp implies two modes of degradation, either at the
ozone-front in the fiber or in the already ozonated region of the fiber, as discussed
previously (Zhang et al. 1999).
The same phenomenon has also been discussed in terms of the MMD of kraft hemlock
pulp in the development of the above-mentioned shrinking core model (Griffin et al.
1998). The cellulose degradation and the bimodality observed in the MMD in the present
study are more distinct than has previously been reported.

8.2 Improving the solubility of softwood kraft pulps in


LiCl/DMAc (paper II)
8.2.1.

Derivatization of cellulose

A study was carried out in order to increase the solubility of softwood kraft pulps in the
solvent LiCl/DMAc. The aim was to increase the solubility by simultaneous derivatization
and dissolution of the cellulose, thus allowing characterizations of the MMD of softwood
kraft pulps by SEC in LiCl/DMAc.
Figure 12 shows the MMD of an underivatized softwood kraft pulp with an intrinsic
viscosity of 1080 ml/g and a kappa number of 18 dissolved to an extent of 92% in 0.5%
LiCl/DMAc.

39

RI

dw/dlogM

UV

5
6
log M (relative to pullulan)

Figure 12. MMD of an unbleached softwood kraft pulp dissolved without derivatization in
LiCl/DMAc. RI and UV denote refractive index and ultraviolet detection, respectively.
MMD of derivatized sample are shown in figure 13.
The MMD in figure 12 differs substantially from those of birch kraft pulps shown in
figures 10 and 11. The MMD of bleached and unbleached softwood kraft pulps has been
thoroughly studied (Sjholm et al. 2000a) and it has been shown to be distributed
differently in comparison with that of a hardwood kraft pulp. The carbohydrate
composition in the eluting MMD changed with decreasing molar mass of the sample. This
result, in combination with studies of the non-dissolvable part of softwood kraft pulp
(Sjholm et al. 1997), suggested that glucomannan is associated with cellulose during the
dissolution and characterization of underivatized softwood samples in the LiCl/DMAc
solvent system. This association makes it difficult to elucidate the correct molar mass of
the cellulose and hemicellulose fractions. The same situation seems to be valid in the
sample shown in figure 12.
Seven derivatization reagents were tested with regard to increase the solubility of an
unbleached softwood kraft pulp. The derivatives tested were: acetyl chloride, propionyl
chloride, butionyl chloride, methyl isocyanate, ethyl isocyanate, propyl isocyanate, phenyl
isocyanate. Acylations increased the solubility of the softwood kraft fibers in LiCl/DMAc,
but the treatments did not always lead to full dissolution. The reason for the poor
reproducibility may be the lack of acid acceptor, or the fact that the samples in the present
study had a considerably larger molar mass than those used by McCormick et al.
(McCormick and Lichatowich 1979; McCormick and Callais 1987), resulting in a lower
accessibility towards the reagents.
The formations of carbamates increased the solubility of the softwood pulp sample. Phenyl
isocyanate worked well as derivatizing reagent, but the strong UV absorption of the phenyl

40

carbamate made the detection of lignin impossible and the use of this reagent was because
of this not further studied. Methyl isocyanate was omitted due to its high vapor pressure
and to handling restrictions because of its toxicity. Ethyl isocyanate (EIC) and propyl
isocyanate (PIC) worked equally well. EIC is a smaller molecule than PIC and was
assumed to affect the hydrodynamic volume of the samples less than PIC, and
derivatization of pulps with EIC was therefore chosen for further study.

8.2.2.

Influence of derivatization on the MMD and molar mass

The derivatization of the softwood kraft pulp introduced major changes in the MMD
(figure 13) compared with that of the underivatized softwood kraft pulp (figure 12).
Figure 13. MMD of a softwood kraft pulp derivatized with ethyl isocyanate. RI and UV
RI

dw/dlogM

UV

log M (relative to pullulan)

denote refractive index and ultraviolet detection, respectively. MMD of underivatized


sample are shown in figure 12.
The typical bimodal MMD of hardwood pulps was also observed for the derivatized
softwood pulp, although the resolution between the low molar mass (LM) distribution and
the high molar mass (HM) distribution was somewhat lower than in the case of the
hardwood kraft pulp. The Mw of the HM distribution in underivatized softwood kraft pulp
samples were surprisingly high, especially considering the low intrinsic viscosity of the
unbleached pulp (table 1).

41

Table 1. Solubility, weight average molar masses (Mw, relative to pullulan) and relative amounts of low
molecular mass (LM) and high molecular mass (HM) distributions as calculated from the chromatograms (UV
and RI) of derivatized and underivatized softwood and hardwood kraft pulps. The Mw was calculated from the RIdistributions.
Mw
Sample
1

Softwood

Softwood , derivatized
Hardwood

Area% (RI)

Area% (UV)

Solubility (%)

LM

HM

LM

HM

LM

HM

92

32000

2338000

96

19

81

100

35400

1765000

20

80

75

25

100

25000

2081000

28

72

84

16

100

29000

2226000

28

72

84

16

Hardwood , derivatized

1) intrinsic viscosity 1080 ml/g and kappa number 18 (underivatized sample)


2) intrinsic viscosity 1310 ml/g and kappa number 17 (underivatized sample)

Derivatization decreased the average molar mass to a more expected value. The relative
areas of the LM and HM distributions in the RI-chromatograms of derivatized softwood
kraft pulps were also different from those of their underivatized counterparts (table 1). The
relative areas of the LM and HM distributions of the derivatized softwood pulps (20%:
80%) reflected the ratio between hemicellulose and cellulose in the sample better than the
MMD of the underivatized sample.
A hardwood kraft pulp was derivatized with EIC in the same manner as the softwood kraft
pulp. The SEC-characterization revealed that the derivatization resulted in no
redistribution of the sample from the HM distribution to the LM distribution. In addition,
the increase in Mw of the LM and HM distribution was small, implying that the introduced
EIC-group did not influence the hydrodynamic volume of the sample to any great extent
(table 1).
It has been suggested that glucomannan is associated with cellulose in spruce fibers
(Salmn and Olsson 1998). Adsorption of glucomannan was observed when cotton linters
were treated with softwood glucomannan fractions isolated from holocellulose (Hansson
1970). MMD measurements showed that high molecular mass material formed when
hardwood pulp was treated with softwood glucomannan (Karlsson and Westermark 1997),
which was suggested as evidence of strong interactions between cellulose and
glucomannan. Further, it has been reported that the carbohydrates in the HM distribution
isolated from underivatized softwood pulp contained mannose (Sjholm et al. 2000a). All
these observations strongly suggest that there are associations between cellulose and
glucomannan in softwood kraft pulp, as is also indicated in figure 11 and table 1. The
redistribution of material from the HM range to the LM range of the distribution seen after
EIC-derivatization of softwood kraft pulps (figures 12 and 13) clearly indicates that strong

42

associations, mainly between glucomannan/cellulose, are ruptured by the derivatization of


the hydroxyl groups of the polysaccharides. This redistribution means that it is possible to
make more correct estimations of the MMD and molar masses of softwood kraft pulps
after derivatization, as the low molecular hemicellulose elutes at the expected elution time,
and does not seem to contaminate the cellulose fraction of the MMD. However, the
carbohydrate composition of the eluting MMD has not been determined within the scope
of this study.
The concomitant shift of the lignin from the HM to the LM part of the distribution suggests
that lignin is associated with the glucomannan rather than directly with the cellulose.
However, the UV-absorbance in the high molar mass fraction after derivatization indicates
that considerable amounts of lignin remain associated with the HM distribution, both in
hardwood and softwood kraft pulps (table 1).
8.2.3.

Reproducibility, degree of dissolution and substitution

It is desirable to dissolve softwood pulp samples, which provide information about the
earlier stages of a kraft cook, but unfortunately the solubility of underivatized kraft pulp
fibers in LiCl/DMAc is reduced at higher kappa numbers. In order to study the effect of
increasing lignin content during EIC-derivatization, five kraft pulps with kappa numbers
ranging from 13 to 53 (corresponding to lignin contents of 2 % to 8 %) were dissolved
using the EIC/LiCl/DMAc method. A gravimetric determination of the non-dissolved
residue showed that 90% dissolution is possible for kraft pulps with kappa numbers around
50 (figure 14). Thus, kraft pulps containing fairly large amounts of lignin have a high
solubility when derivatized with EIC.

% dissolved

100
95
90
85
80
13

18

31

41

53

Kappa number

Figure 14. Percentage dissolved for unbleached kraft pulps with different lignin content.
Samples dissolved through derivatization according to the EIC/LiCl/DMAc-method.

43

The maximum theoretical degree of substitution is three, since the glucosidic units in
cellulose contains three hydroxyl groups. The degree of substitution of derivatized
softwood pulp samples was determined by nitrogen analysis to be 1.9, with a standard
deviation of 0.13 (n=4). This is slightly lower than the previously reported values of 2.1
(McCormick and Callais 1987), but it was nevertheless considered to be high enough to
provide complete of dissolution of the sample.
The reproducibility of the EIC/LiCl/DMAc-method was evaluated in five separate MMD
determinations of the same softwood kraft pulp. The level of reproducibility of the
determination of the molar mass of the HM distribution was high, with a relative standard
deviation (RSD) of 2.7 % (n=5). These numbers are similar to the reproducibility for
hardwood pulps where a RSD of 2.2 % (n=3) has been reported for the HM distribution
(Sjholm et al. 2000b).

8.3 Degradation patterns observed during alkaline degradation


of softwood and hardwood (paper III)
The purpose of chemical pulping is to degrade and dissolve the lignin in wood. In kraft
pulping processes, lignin is degraded in a solution containing hydrogen sulfide and
hydroxide ions at an elevated temperature. Unfortunately, the cellulose and hemicelluloses
are also partly degraded and dissolved under these conditions. This decreases the yield and
the molecular mass of both cellulose and hemicelluloses. A decrease in molecular mass is
undesirable, since the mechanical strength of wood pulps, cotton and cellulose derivates
decreases with decreasing molecular mass (Staudinger and Reinecke 1938; Sookne and
Harris 1945; Ifju 1964; Gurnagul et al. 1992).
The aim of the work reported in paper III was to study the pattern of the alkaline
degradation of cellulose and hemicelluloses during kraft pulping and how the MMD and
fiber strength were influenced by different alkaline conditions. The degradation of Norway
spruce (Picea abies) and birch (Betula sp.) was studied in two series: by varying the
pulping time using the same initial alkali concentration, or by varying the initial alkali
concentration and maintaining a constant cooking time. Experimental conditions are given
in appendix II and the sample designations are listed in table 2.

44

Table 2. Designations of the samples in paper III.


Sample abbreviation

Wood Specie

Degradation strategy

BA

Birch

Increased initial Alkali level

BT

Birch

Increased pulping Time

SA

Spruce

Increased initial Alkali level

ST

Spruce

Increased pulping Time

Figures 15 and 16 show the MMD of birch pulps produced at different alkali
concentrations and at different degradation times, respectively. The typical bimodal MMD
is clearly seen for pulps with high molecular masses. The bimodality of the MMD for the
more severely degraded pulps is less clear, since the degraded cellulose fraction in these
cases overlaps the hemicellulose fraction.

dw/dlogM

Increasing alkali concentration

5
6
log M (relative to pullulan)

Figure 15. MMD of birch pulps produced with different alkali concentrations (BA-series).

dw/dlogM

Increased pulping time

5
6
log M (relative to pulllan)

Figure 16. MMD of birch pulps produced with different pulping times (BT-series).

45

Differences in MMD of the fibers were observed between the BT and BA series. In the BT
series, a fraction of higher molecular mass cellulose distributed at log M>6, was seen
throughout the entire series. At high alkali concentrations, all the high molecular mass
cellulose was degraded and no such high molecular mass fraction was evident (figure 15).
The difference in degradation pattern is also observed when the polydispersity of the
MMD of the pulps is plotted versus the intrinsic viscosity of the pulps, figure 17.

Polydispersity (Mw/Mn)

25
20

BA

15

BT
SA

10

ST

5
0
0

500
1000
Intrinsic viscosity (ml/g)

1500

Figure 17. Polydipersity (Mw/Mn) for the birch (B*) and spruce (S*) pulps produced with
a) different pulping times (*T) and b) different alkali concentrations (*A). Molar masses
calculated from the entire MMD.
In addition to the fraction of high molar mass cellulose, the fibers in the BT series
contained a larger fraction of xylose than those in the BA series. It is known that more
xylan is dissolved in cooks with higher alkali concentration (Aurell and Hartler 1965;
Gustavsson and Al-Dajani 2000). In the case of the BT series, a distinct distribution of
xylan was observed in the MMD of all the pulps, regardless of degradation level (figure
16). However, in the MMD of the BA series, no distinct xylan distribution could be
discerned in the more degraded pulps produced with high initial alkali concentration
(figure 15).
During this work, it was assumed that if the wood polymers had been degraded
homogeneously, the entire MMD would have been shifted towards lower molar mass. This
was the case for the MMD observed in the BA-series. The remaining high molar mass
cellulose fraction in the MMD of the BT-series resulted in the suggestion that the cellulose
degradation in the BT series was interpreted to occur less homogeneously than the
degradation in the BA series.
The pulps in the BT-series had higher strengths compared to the BA-series pulps,
evaluated as dry zero-span tensile index (figure 18) at comparable viscosity levels. It is

46

suggested that the reason for the higher strength is related to the residue of the high
molecular mass cellulose found in the MMD of these fibers, despite severe alkaline
degradation. The plot of dry zero-span tensile indexes versus Mz-values versus exemplifies
this. A single relation between fiber strength and Mz is observed, as Mz is more affected by
larger molecules than Mw or than the intrinsic viscosity. Plotting zero-span tensile index
versus the other estimates of the molar mass revealed two different relations for the two
birch series.
However, when the fiber strength was measured as rewetted zero-span tensile index, the
same relationships were found for the birch pulps, irrespective of degradation strategy
(data not shown). Rewetted zero-span tensile index is more sensitive to local defects than
dry zero-span tensile index (Mohlin and Alfredsson 1990). It seems that these defects were
more important for the rewetted zero-span tensile index than the presence of higher
molecular mass material.

47

220
Dry zero-span tensile
index (Nm/g)

BA

180

BT
140

SA
ST

100
60
0

1500

220
Dry zero-span tensile
index (Nm/g)

500
1000
Intrinsic viscosity (m l/g)

180

BA
BT

140

SA

100

ST

60

500

1000

1500

2000

M w (kg/m ol)
220
Dry zero-span tensile
index (Nm/g)

BA

180

BT
140

SA
ST

100
60
0

1000

2000

3000

4000

Mz (kg/mol)

Figure 18. Dry zero-span tensile index plotted versus: A) intrinsic viscosity, B) Mw and C)
Mz for birch and spruce pulps.

48

The development of the derivatization method (paper II) made it possible to study the
degradation pattern of lignin-containing softwood kraft pulps. Figure 19 shows the MMD
of the softwood series subjected to different alkali concentrations.

dw/dlogM

Increasing alkali concentration

5
6
log M (relative to pullulan)

Figure 19. MMD of spruce pulps produced with different alkali concentrations (SAseries).
Degradation of spruce chips, independent of degradation conditions, resulted in
degradation patterns similar to those observed in the BA-series: a high molecular mass
cellulose fraction was not evident in the MMD of the wood polymers in any of the spruceseries. The same results was also suggested when the polydispersity was plotted versus
intrinsic viscosity (figure 17). Thus, it was concluded that there were no differences in
homogeneity of the cellulose degradation pattern of the spruce chips. The relations
between fiber strength and molar masses in the spruce series were also the same,
irrespective of degradation strategy (figure 18).
Alkaline degradation of birch and spruce seems to proceed differently, depending both on
the wood specie and degradation type, when degradation is evaluated as changes in MMD
and decrease in fiber strength.

8.4 Improving the accuracy in the determination of molar mass


of cellulose in kraft pulps (paper IV)
The objective of paper IV was to implement a multi-angular laser light scattering
(MALLS) detector in the SEC-characterizations of kraft pulps. This was performed
because the separation system used during SEC needs a calibration. The direct-standard
calibration normally performed is based on a number of assumptions, where the
assumption of chemical similarity between standards and samples is the most important

49

and probably erroneous. The advantage of using a MALLS-detector is that calibration with
external standards is avoided.
The most common method of calibrating the chromatographic system used for the
characterization of pulps is by using a series of well-characterized pullulan standards
(Westermark and Gustafsson 1994; Strlic et al. 1998; Sjholm et al. 2000a; Sjholm et al.
2000c). Pullulan is a polymer consisting of (14) linked -D-glucopyranose units with
every third (14)-linkage replaced by a (16)-linkage (Popa and Spiridon 1998), while
cellulose consists solely of (14)-linked -D-glucopyranose units. This difference in
linkage structure between standards and samples has been found to cause an
overestimation of the molecular mass and its distribution in water-soluble cellulose
derivatives (Poch et al. 1998) as well as in chitosans (Otty et al. 1996), and there is a
suspicion that a similar situation may exist in the case of cellulose and pullulan (Bikova
and Treimanis 2002).
The MMDs of the same sample of unbleached birch kraft pulp determined by two different
sets of detectors, SEC/RI and SEC/MALLS/RI, are shown in figure 20.

System I

dw/dlogM

System II

log M

Figure 20. MMD of a birch kraft pulp characterized by size-exclusion chromatography


calibrated with I) pullulan standards (SEC/RI), and II) light scattering detection
(SEC/MALLS/RI).
The results from system I were obtained by calibration with nominal peak molar masses of
pullulan standards, and the results from system II were obtained by MALLS-detection in
combination with RI-detection (dn/dc of a fully bleached kraft pulp dissolved in 0.5%
LiCl/DMAc determined at 488 nm and 40C to 0.108). Clearly, there are differences in the
reported MMD, and consequently, in the calculated molar masses of the samples,
depending on how the detection is performed and calibrated.

50

The difference in the MMDs in figure 20 is probably due to structural differences between
standards and samples. The (16)-linkages in pullulan lead to conformational freedom of
the pullulan (Brant and Burton 1981) compared to structures containing only (14)linkages. At a given hydrodynamic volume, or elution time, the calculated molecular mass
of the cellulose is smaller than the molecular mass of pullulan. This leads to an
overestimation of the molecular mass of cellulose. A similar situation was found when a
water-soluble cellulose derivative was studied (Poch et al. 1998), and two methods of
obtaining more correct molar mass were presented. Similar methods, based on the findings
by Poch et al. (1998) were thus derived for the cellulose/LiCl/DMAc system.

8.4.1.

Method I: Correction of the molar mass of cellulose

The differences in molar mass determined by the two detector techniques are shown for
cellulose samples with a large range of molar masses in figure 21.

log M x MALLS

6.2
Mw

5.8

Mn

R = 0.963

R = 0.995

5.4
5
4.6
4.6

5.1

5.6

6.1

6.6

log Mx relative pullulan

Figure 21. Relation between average molar masses (Mw and Mn) determined relative to
pullulan standards (SEC/RI) and by light scattering detection (SEC/MALLS/RI).
The resulting correlations for the Mn and Mw values were found to be
85 0.02)
M w,Correlated = (4.00 1.08) M w(0,.Pull
.690.04)
M n,Correlated = (43.8 25.3) M n(0,Pull

(27)
(28)

where Mw,Pull and Mn,Pull is the weight average and number average molar mass,
respectively of the cellulose fraction estimated relative to pullulan standards. The
confidence limits were calculated according to a t-test at the 95% confidence level. (Miller
and Miller 1993).

51

The relations shown can be used to relate erroneous molar masses of the cellulose fraction
estimated relative to pullulan standards to more correct molar masses, measured by light
scattering detection. These relationships are useful when determining the average molar
masses of cellulose by RI-detection when no light scattering detector is available.

8.4.2.

Method II: Cellulose-equivalent molar masses of pullulan


standards

A second method was developed in order to calculate cellulose equivalent molar masses of
the standards. The advantage of this method is that not only the average molar masses but
the entire MMD is taken into account.
The purpose of calibrating a SEC-system with a set of standards is to obtain a relationship
between the logarithm of the molar mass of the standards and the elution time (equation
29). A similar calibration line is normally obtained from the SEC/MALLS/RI
measurements (equation 30). The equations hold the calibration constants a and b from the
corresponding detection system.
= b pull + a

log M

pull

log M

MALLS

= b MALLS

pull

+ a MALLS t

(29)
(30)

Figure 22 shows that the regression lines from the two systems differ, but this difference
may be used to express new, cellulose-equivalent molar masses of the standards used.

52

log M

5
Pullulan, log M according to supplier

Kraft pulps, log M according to system II

3
20

22

24

26
28
Elution time (min)

30

32

34

Figure 22. Calibration curve obtained by direct-standard calibration with pullulan


standards (system I) and calibration curves for SEC/MALLS/RI-detection (system II). The
ellipse indicates the region corresponding to elution time of cellulose.
At the time t where the calibration curve for the pullulan standards and the elution curve
for the pulp samples intersect, the following expression may be derived.
M

celleq

= M

MALLS

= qM

p
pull

(31)

where q and p are functions of the slopes (apull and aMALLS) and intercepts on the ordinate
(bMALLS and bpull) of the pullulan and MALLS calibration lines, respectively.
If the molar mass of an arbitrary pullulan standard is inserted in equation 31, the equivalent
molar mass of the cellulose eluting at any time t from the MALLS detection is obtained. It
is thus possible to calculate the cellulose-equivalent molar mass of the narrow standards,
Mcelleq, by using equation 31.
The calibration constants bMALLS and aMALLS differ slightly from sample to sample and also
from injection to injection, which is indicated in figure 22. In order to minimize the errors
in the functions q and p in equation 31, thirty injections were performed with high molar
mass birch kraft pulps and wood polymer beads (WPB) with lower molar mass. Only the
cellulose fraction was taken into account in the estimation of the functions q and p, as the
low S/N ratio of the hemicellulose fraction would increase the error in the functions. The
resulting cellulose-equivalent molar masses of the pullulan standards are shown in table 3,
with standard deviations

53

Table 3. Nominal and cellulose-equivalent molar masses of pullulan standards based on method II, together with
standard deviation of the cellulose-equivalent molar masses (n=30).
Nominal molecular
mass

Cellulose-equivalent

Standard

molecular mass

deviation

738

2300

980

5800

11200

3600

12200

19800

5500

23700

33100

8400

48000

57300

12700

100000

101000

19100

186000

164000

26800

380000

287000

39300

853000

540000

62300

1660000

912000

98000

*As stated by the supplier.

The standard deviations in the low molar mass range are large, and this is due to the low
S/N-ratio of the light scattering in that range. The standards of low molar masses are
however located well outside the elution range of cellulose. The deviations in the high
molar mass range are considered low as they only have small effects on the calibration
curve (data not shown).

8.4.3.

The applicability of the improved evaluation methods

The applicability of the two methods was evaluated by the characterization of the molar
mass and the MMDs of WPBs subjected to acid hydrolysis on a second column set.
Figure 23 shows the MMD of the undegraded WPB as obtained by three different
methods:
a) SEC/RI relative to nominal molar masses of standards (MMDrel.pull.) by directstandard calibration
b) SEC/RI relative cellulose-equivalent molar masses of standards (MMDrel.cell.eq.) by
using method II
c) SEC/MALLS/RI (MMDMALLS)

54

MMD rel.cell.eq.

dw/dlogM

MMDMALLS

MMDrel.pull.

log M

Figure 23. MMD of a birch kraft pulp determined by three calibration techniques using:
a) the nominal molar mass of the pullulan standard (MMDrel.pull.), b) the celluloseequivalent molar masses of the pullan standards (MMDrel.cell.eq.), c) light scattering
detection (MMDMALLS).
The SEC/RI method relative to nominal molar masses of standards provided a distribution
of cellulose with higher molar mass than the SEC/MALLS/RI technique, as was expected.
The method of calibrating the chromatographic system with cellulose-equivalent molar
masses of pullulan standards resulted in an MMD (MMDrel.cell.eq) that agreed much better
with the MMD characterized by the MALLS detector (MMDMALLS). Consequently, method
II is preferred to characterize the MMD of cellulosic samples rather than by using the
direct-standard calibration method.
The relative errors of the weight average molar mass from the MMDs of the degraded
WPBs are shown in table 4, assuming that the SEC/MALLS/RI method provides the
correct value.
Table 4. Relative error of the estimated Mw values of acid-hydrolyzed WPB shown in figure 23. Mw estimated: a)
relative to the nominal molar masses of standards, b) by method I, and c) by method II. All the values are
compared to values from SEC/MALLS/RI. The signs denote overestimation (+) and underestimation (-).
Hydrolysis t (h)

a) Nominal

b) Method I

c) Method II

+55 %

-20 %

-16 %

+70 %

-12 %

-6 %

+70 %

-9 %

-1 %

23

+90 %

+6 %

+20 %

55

The values were estimated relative to the nominal molar masses of standards, by correcting
the molar mass according to method I, and by using cellulose equivalent molar masses of
the standards (method II). The errors introduced by using the nominal molar masses of
standards was reduced by using either method I or II. It is however recommended that
method II should be used when a light scattering detector is not available, as an estimation
of the entire distribution is obtained. By using method I, only the average molar masses of
the distribution is obtained.

8.5 Simulation of cellulose degradation (paper V)


A study was undertaken to simulate the way in which the shape of the MMD of wood
polymers would develop during degradation, provided the degradation occurred according
to pre-defined patterns of cellulose chain scission in the fiber. The study focused on
random degradation and end-wise degradation, so-called peeling, of the molecules within
the pulp fiber.

8.5.1.

Validation of the simulation model

In order to validate the simulation model, both fictive and real distributions were subjected
to simulated random degradation. The results were found to be consistent with those of
earlier studies (Ballauff and Wolf 1981; Guiata et al. 1990) indicating that the
polydispersity (Mw/Mn) of a single distribution, irrespective of the initial polydispersity,
approaches the value of 2 during the simulation of random degradation (data not shown).
Figure 24 shows an example of the simulation of random degradation. The MMD
measured by RI-detection while performing SEC of an unbleached birch kraft pulp, was
used as initial data.

56

dw/dlogM

5
log M

Figure 24. MMD obtained by simulating homogeneous random degradation of the


molecules in a birch kraft pulp.
The distribution was divided into subunits having a size corresponding to 20 glucosidic
units. The smaller the subunits used, the more exact will the calculated results be.
However, as the subunits became smaller, more and more computer-memory was required,
and longer simulation times were necessary. Supplementary calculations made with larger
and smaller subunits indicated that the results did not depend to any considerable extent
upon the size of the subunits.

8.5.2.

Random degradation

The assumption made to simulate the MMD (figure 24) was that each linkage in all
polymer chains had the same degradation rate, and that all the linkages were equally
susceptible. The degradation pattern shown in figure 24 is thus not only interpreted as
representing degradation according to a random pattern, it is also interpreted as
representing degradation according to a homogeneous pattern, since all the material was
subjected to random degradation.
The simulated MMD in figure 24 reveals several interesting features. In the case of the
high molar mass distribution, it is observed that the entire distribution is shifted towards
lower mass range. The entire distribution is degraded, as no cellulose initially found to
have a molar mass higher than logM = 6 retained its initial molar mass. This has
previously been observed during the exaggerated kraft pulping of birch and pine chips in
paper II, where only empirical conclusions could be drawn regarding the homogeneity of
the degradation. It can thus be confirmed that some of the assumptions made in that study
were correct; homogeneous degradation does result in a shift of the entire MMD, including
the high molar mass region, towards lower molar masses, as exemplified in figure 24.

57

Although there are differences in reactivity between linkages in hemicellulose and


cellulose, no consideration was here given to the kind of molecule existing at a given chain
length (Lai 2001). The low molar mass distribution, representing the hemicellulose
fraction, was clearly visible as a low molecular shoulder throughout the simulation. Some
degradation of the hemicellulose distribution is indicated in figure 24, but the distribution
is still visible throughout the simulation. The reason for this is found in equation 14. The
rate of degradation of a given hemicellulose molecule is lower than that of a cellulose
molecule, as there are fewer hydrolysable linkages in a hemicellulose molecule.
The simulated MMDs differ from the MMDs reported for random degradation of linear
polymers (Emsley and Heywood 1995). In that study it was found that random degradation
did not change the position of the MMD. This is contradicted by the results in figure 24.
The explanation of the differences may be related to how random degradation is defined
and how the calculations are performed. Although our model is based on an algorithm, the
approach used by Emsley and Heywood (1995) was a Monte Carlo simulation in which
random degradation was defined differently. Random degradation may be defined in terms
of the reactivity of different glucosidic linkages within a cellulose. This definition is the
same in both studies. It may also be expressed in terms of the size of the fraction of the
material that is subject to any kind of degradation, and it is this parameter that probably
explains the differences in the MMD obtained by simulating random degradation.
It is here proposed that if the polymers in pulp fibers are assumed to be subjected to
homogeneous degradation, it implies that the entire material, i.e. 100% of the material
constituting the pulp fiber, has the same reactivity according to a random pattern within the
polymer. The resulting MMD of birch kraft pulps is exemplified in figure 24.

8.5.3.

End-wise depolymerization

Figure 25 shows the effect of pure end-wise degradation with rate constants according to
equations 23 and 24 applied to the MMD of a birch kraft pulp. The position of the high
molar mass cellulose distribution and, consequently, the molar mass is only slightly
affected by peeling. This is expected, as extensive peeling is required to have an effect on
these large molecules. The effect on the low molar mass distribution is however more
obvious. Here, the material is degraded during the course of simulation, and finally
removed from the system after it reaches a DP of 2 (which translates to 40 glucosidic units
since the smallest unit in the simulation has the size of 20 units).

58

dw/dlogM
3

5
log M

Figure 25. MMD obtained after theoretical simulation of end-wise depolymerization of the
molecules in a birch kraft pulp.

8.5.4.

Linear combinations of chain scission cases

Several modes of degradation are active simultaneously under process-like conditions. It is


possible to assign different rates of reaction and different chain scission cases to different
relative amounts of the MMD in the simulation model. By doing this, it was possible to
simulate experimental results by linear combinations of peeling and random chain
scissions.
Figure 26 shows experimental MMDs from the RI-detection during SEC-characterization
of softwood, kraft pulped to different lignin contents. The shift in the position and the
change in shape of the MMD during pulping was not the same as in the simulation in
figure 24, and this is interpreted to indicate that the cellulose degradation during kraft
pulping is of a more heterogeneous character than is depicted in figure 24.

59

dw/dlogM

log M

Figure 26. MMD of spruce kraft pulps, digested to different lignin contents.
Efforts were made to simulate the indicated heterogeneity of the degradation in the
experimental MMD in figure 26, by using the MMD of the pulps with highest lignin
content as initial data when simulating the degradation. The assumptions made prior to
these efforts were:
The wood fibers contained two different fractions of polymers; one more accessible
fraction giving a higher rate of degradation and one less accessible giving a lower
rate of degradation,
The two fractions contained polymers from the entire MMD,
Both fractions were susceptible to random degradation with different rate constants
depending on to which fraction the material belonged, and end-wise
depolymerization with the same rate constant, regardless of fraction.
The molecules are still degraded into chains of random lengths but with different rates, i.e.
the degradation is no longer homogeneous, but still random.
In order to find a minimum in the difference between the experimental and simulated
MMDs, the relationship between the two random rate constants and the composition of the
system was varied. The ratio between the reaction rates was varied from 0.05 to 0.5 and
the composition of the system was varied in 10 %-increments, from 0-100%. The resulting
simulated MMDs with the best fit to the experimental MMDs are presented in figure 27.
The conditions found to best describe the experimental MMDs were:
20 % of the sample was more accessible, while 80% was less accessible.
The relation between the rate constants was 0.1, i.e. the rate constant of the less
accessible fraction was an order of magnitude lower than the rate constant of the
more accessible fraction.

60

dw/dlogM
3

5
log M

Figure 27. MMD obtained by simulation of heterogeneous degradation of wood polymers


during softwood kraft pulping. The material was assumed to be composed of two fractions
of different relative amounts with different reactivities: 20% with higher reactivity and
80% with a lower reactivity.
These parameters resulted in simulated MMDs very similar to the experimental MMDs.
However, one reservation is necessary. The molar masses of the cellulose in figures 24 and
27 all greatly exceed the molar masses of hydrocelluloses previously studied (Machell and
Richards 1957; Haas et al. 1967; Johansson and Samuelsson 1975). It is not possible to
extend the interpretation of the degradation pattern to cellulose having a molar mass lower
than that depicted in figure 24, as the rate of degradation is assumed to be affected as the
size of the molecules approaches the size of the molecules in cellulose crystallites (Battista
et al. 1956; Gentile et al. 1987). It nevertheless seems probable that the alkaline
degradation depicted in figure 26 is due to a random degradation of the linkages in the
cellulose chain.
A more careful analysis using simulation in combination with experimental data can be
used to better understand and describe the molecular degradation of the cell wall during
conditions similar to those prevailing in industrial processes. The analysis provides
information that is neither observed by intrinsic viscosity measurement nor by the study of
the development of the MMDs per se.
This difference in reactivity is very interesting. Although it is not difficult to envisage e.g.
that cellulose on the surface of fibrils or fibrillar aggregates is more susceptible to
degradation than chains in the interior, this is seldom discussed in the literature. Instead,
most studies suggest that the fiber cell wall should be considered to be a swollen gel in
which active species can move quite easily (Malcolm and Grace 1989). The experimental

61

results given here actually suggest that the alkaline degradation process proceeds
heterogeneously.

8.5.5.

Peeling of cotton linters

In order to study the effect of the peeling reaction on cellulosic material and the utilization
of the simulation model, cotton linters were subjected to alkali at rather low temperatures
(130C) compared to normal kraft pulping temperatures. At this temperature, the
degradation is dominated by peeling and only minor alkaline hydrolysis occurs (Haas et al.
1967; Lai and Sarkanen 1967). In addition, the ratio between the rate constants of peeling
and stopping reaction is approximately 250 (Haas et al. 1967). The stopping reaction thus
has only a minor influence on the results. The resulting weight average molar masses are
plotted vs. treatment time in figure 28.
The changes in weight average molar mass show that the initial degradation rate is high.
This rapid degradation is interpreted as being a degradation of the more accessible parts of
the structure, or more probably, as alkaline cleavage of oxidized cellulose structures that
are more sensitive towards alkali (Richards 1971). After the initial depolymerization stage,
the rate of degradation decreases, and the weight average molar mass even increases
during the intermediate stages of degradation.
Attempts were also made to simulate the experimental MMDs. The original MMD from
the cotton linters was used as raw data, and simulated MMDs were obtained by assuming
that there were two different degradation phases. During the first phase, it was assumed
that only random, homogeneous, chain scission took place. The initial stage of hydrolysis
was assumed to last until the Mw-value of the simulated MMD reached the same Mw-value
as the cotton linters subjected to 60 min of alkaline degradation. Thereafter, only end-wise
depolymerization of the cellulose was simulated. The resulting weight average molar
masses from the experimental MMDs and after such simulation attempts are also showed
in figure 28.

62

1.2E+06
Simulated
Experimental

Mw

1.0E+06

8.0E+05

6.0E+05
0

100

200
300
t (min)

400

500

Figure 28. Weight average molar masses (Mw) of experimental MMD and Mw of MMD
obtained while simulating peeling of cotton linters.
The development of the Mw-values of the simulated MMD is similar to that of the
experimental Mw-values. After the initial stage of hydrolysis an increase is observed, that
passes over to a decrease in the final stages of hydrolysis. The simulation model may thus
be used to study the development of the molar mass during end-wise depolymerization.
However, the theoretical yield from the simulated peeling of cotton linters in figure 28 is
low (Johansson and Samuelsson 1975), especially after longer degradation times, implying
that the model may be too simple to describe the peeling reaction in terms of yield with a
high level of accuracy (table 5).
Table 5. Weight average molar mass (Mw), turnover (equation 25) and yield (equation 26) from the simulation of
alkaline peeling of cotton linters.
Simulation time

t0

t1

t2

t3

t4

t5

t6

t7

t8

Mw (kDa)

1148

963

829

827

786

786

789

784

771

Turnover, X

0.0

0.3

0.6

3.0

60.6

85.0

94.3

97.8

99.2

Yield, Y

100.0

100.0

100.0

97.6

39.8

15.1

5.7

2.2

0.8

A tentative explanation of this deficiency is that not only one unit at a time is peeled off
from the reducing end during experimental peeling. Instead, a number of units are removed
simultaneously before a stopping reaction occurs (Franzon and Samuelsson 1957; Lai and
Sarkanen 1967). Hence, further development of the simulation model is needed in order to
describe the development of the yield and simultaneous development of the MMD during
the end-wise depolymerization of cellulose.

63

8.5.6.

Limitations of the simulation study

The construction of simulation models always involves simplifications that introduce


errors and drawbacks when studying complex materials, and paper V is no exception. The
model used in this study is a bulk model, meaning that no consideration is given to
differences in reactivity caused by topochemical differences. The model could nevertheless
simulate gross differences in reactivity by assuming different reactivities to mass fractions
of an MMD. For more complex modes of degradation, however, such as acid hydrolysis,
this approach is not feasible. One of the objectives of this study was to simulate acid
hydrolysis, which has been proposed to occur in disordered regions of cellulose (Sharples
1957; Durawalla and Shet 1962; Gurnagul et al. 1992). That objective was not fulfilled,
mainly because of the rather complex character of the acid hydrolysis, which means that
several parameters of the cellulose in the cell wall need to be taken into account to
simulate acid hydrolysis properly. Some examples of these parameters are; the rates of the
two phases of degradation (Durawalla and Shet 1962), the leveling off degree of
polymerization (Battista et al. 1956) and the fibrillar structure of cellulose (Wickholm et
al. 2001).
Instead of using the model of Ballauff and Wolf, which is a versatile, but nevertheless a
rather simplified model, it would be useful to develop a model based on a Monte Carlotechnique, where it would be possible to achieve an assessment of cellulose with different
reaction rates and different crystallinities.
There are other limitations in the work presented in paper V. The peeling reaction is
accompanied by a stopping reaction (Franzon and Samuelsson 1957; Machell and Richards
1957), and a correct implementation of the stopping and peeling reaction has not yet been
achieved. The different polysaccharides present in the pulp fiber all have different
reactivities during alkaline degradation (Lai 2001), and a more exact treatment of these
reactivities is preferred.

64

9. Conclusions
This thesis has shown some of the benefits achieved by using size-exclusion
chromatography in combination with more traditional analytical techniques to understand
the effect of cellulose degradation during pulping and bleaching processes. The thesis also
describes the development of dissolution procedures and detection techniques to be used
for characterizing the molar mass distribution of chemically produced wood pulp fibers. It
was found that:
SEC revealed that ozone degradation and acid hydrolysis resulted in two
completely different degradation patterns, not observed in intrinsic viscosity
measurements. The differences in zero-span tensile strength between ozonedegraded and acid-hydrolyzed pulp could not be explained by differences in shape
factor. The type of degradation influenced the fiber strength more than the change
in shape factor.
The solubility of the polymers in softwood kraft pulp fibers was enhanced by
derivatization of the samples during simultaneous dissolution in 8 % LiCl/DMAc.
This makes it possible to characterize the MMD by SEC using 0.5% LiCl/DMAc as
solvent system, allowing a more correct determination of the mass averages of the
pulp polymers than by the characterization of underivatized softwood samples.
Alkaline degradation of cellulose in birch and spruce wood showed significant
differences between different degradation conditions and different wood species.
Differences in fiber strength observed during alkaline degradation were related to
changes in molar mass distributions.
The comparison between the molar masses of the samples estimated by directstandard calibration and by MALLS showed that the direct-standard calibration
resulted in an overestimation of the molar masses of the samples. Two methods to
correct for this overestimation were developed.
The usefulness of applying theoretical considerations and models to better
understand complex degradation phenomena occurring during kraft pulping of
wood fibers was demonstrated. Alkaline hydrolysis has been interpreted to occur as
a random, not necessarily homogeneous, degradation of the polymer chains. A
hypothesis for the development of the MMD of wood fibers under kraft pulping
conditions was proposed.

65

10. Acknowledgements
First, I would like to thank my supervisor Mikael Lindstrm. This work would not have
been accomplished without his support, belief and mentorship from the beginning to the
end. Special thanks to Fredrik Berthold and Elisabeth Sjholm for all their inspiration,
encouragement and ideas. Many thanks are also due to my other co-authors (Kristina
Gustafsson, Ulrika Molin and Helena Lennholm) for a very enjoyable time.
Parts of this work were carried out within the framework of the Wood Ultrastructure
Research Centre (WURC), and financial support from WURC and STFI is gratefully
acknowledged.
Many people have been involved, directly or indirectly, during the course of this work, and
I thank
Eva-Lisa, Rickard, Anders, Maria, Anna, Sofia, Jessica, John, Sara and Lina, for all
the laughter,
Per Persson for all the discussions of the important and not so very important
aspects of life and for valuable criticism of this manuscript,
Maria Nordstrm for valuable criticism of this manuscript,
Anthony Bristow for the linguistic revision,
Therese Sundquist, Erica Wedin and Manne Kallin for the exceptional support of
books, journal articles and computers,
the guys at KTH: Ragnar, Patrik, Martin L, Mikael and Gunnar
all my colleagues at STFI and KTH for the past five years.
Last, but certainly not least: a thought to my family, particularly Maria who has been very
tolerant and loving to the absent-minded person hanging around the house during the last
months.

67

11. References
Aurell, R. and N. Hartler (1965). Kraft pulping of pine part 2. Sv. Papperstidn., 4, 97102.
Ballauff, M. and B. A. Wolf (1981). Degradation of chain molecules. 1. Exact solution of
the kinetic equations. Macromolecules, 14, 3, 654-658.
Battista, O. A., S. Coppick, J. A. Howsmon, F. F. Morehead and S. W. A. (1956). Leveloff degree of polymerization. Ind. Eng. Chem., 48, 2, 333-335.
Bikova, T. and A. Treimanis (2002). Problems of the MMD analysis of cellulose by SEC
using DMA/LiCl: a review. Carbohyd. Polym., 48, 1, 23-28.
Brant, D. A. and B. A. Burton (1981). The configurational statistics of pullulan and some
related glucans. In: Solution Properties of Polysacharides. D. A. Brant,
Washington, DC, ACS Symposium Series. 150, 81-99.
Brown, W. (1967). The cellulose solvent cadoxen. Sv. Papperstidn., 70, 15, 458-461.
Brunow, G., I. Kilpelinen, J. Sipil, K. Syrjnen, P. Karhunen, H. Setla and P. Rummako
(1998). Oxidative coupling of phenols and the biosynthesis of lignin. In: Lignin
and lignan biosynthesis. N. G. Lewis and K. V. Sarkanen, Washington, DC,
American Chemical Society, 146.
Brndstrm, J. (2002). Morphology of Norway spruce tracheids with emphasis on cell wall
organisation. PhD-thesis. Department of Wood Science, Swedish University of
Agricultural Sciences, Uppsala.
Chiu, S. and L. Paszner (1975). Potentiometric titration behavior of sodium sulfide,
methyl mercaptan, dimethyl sulfide, dimethyl disulfide and polysulfides in mixed
alkaline solutions and sulfate pulping black liquors. Anal. Chem., 47, 12, 1910.
Corbett, W. M. and G. N. Richards (1957). The degradation of cellulose by aqueos
sodium hydroxide at 170C. Sv. Papperstidn., 60, 21, 791-794.
Core, H. A., W. A. Ct and A. C. Day (1976). Wood- structure and identification.
Syracuse, Syracuse University Press, 31.
Durawalla, E. H. and R. T. Shet (1962). Heterogeneous acid hydrolysis of alpha-cellulose
from Sudanese cotton. Text. Res. J, 32, 11, 942-954.

69

Ekenstam, A. (1936). ber das Verhalten der Cellulose in Mineralsure-Lsungen, II.


Mitteil: Kinetisches Studium des Abbus der Cellulose in Sure-Lsungen. Ber.
Dtsch. Chem. Ges., 69, 553-559.
Emsley, A. M. and R. J. Heywood (1995). Computer modelling of the degradation of
linear polymers. Polym. Degr. Stab., 49, 145-149.
Eriksson, T. and T. Reitberger (1995). Formation of hydroxyl radicals from direct ozone
reactions with pulp constituents. 8th International Symposium on Wood and
Pulping Chemistry, Helsinki, Vol. 2, 349-354.
Evans, R. and A. F. A. Wallis (1989). Cellulose molecular weights-determined by
viscometry. J. Appl. Polym. Sci., 37, 8, 2331-2340.
Evans, R., R. H. Wearne and A. F. A. Wallis (1989). Molecular Weight Distribution of
Cellulose as Its Tricarbanilate by High-Performance Size-exclusion
Chromatography. J. Appl. Polym. Sci., 37, 12, 3291-3303.
Fahln, J. (2002). Ultrastructural arrangement of the polymers in wood fibers. Techn.
Lic.-thesis. Department of Polymer Technology, Royal Institute of Technology,
Stockholm.
Fengel, D. (1970). Ultrastructural behaviour of cell wall polysaccharides. Tappi, 53, 3,
497-503.
Fengel, D. (1971). Ideas on the ultrastructural organisation of the cell wall components.
J. Polym. Sci. Part C, 36, 383-392.
Fengel, D. and G. Wegener (1984a). Wood chemistry, ultrastructure, reactions.
Berlin/New York, Walter de Gruyter, 106.
Fengel, D. and G. Wegener (1984b). Wood chemistry, ultrastructure, reactions.
Berlin/New York, Walter de Gruyter, 296-300.
Fengel, D. and G. Wegener (1984c). Wood chemistry, ultrastructure, reactions.
Berlin/New York, Walter de Gruyter, 13-17.
Fengel, D. and G. Wegener (1984d). Wood chemistry, ultrastructure, reactions.
Berlin/New York, Walter de Gruyter, 75.
Franzon, O. and O. Samuelsson (1957). Degradation of cellulose by alkali cooking. Sv.
Papperstidn., 60, 23, 872-877.

70

Gentile, V. M., L. R. Schroeder and R. H. Atalla (1987). Physical structure and alkaline
degradation of hydrocellulose. In: The structures of cellulose. R. H. Atalla,
Washington, DC, ACS Symposium Series. 340, 273-291.
Gierer, J. (1980). Chemical aspect of kraft pulping. Wood Sci. Techn., 14, 241-266.
Griffin, R., Y. Ni and A. R. P. Van Heningen (1998). The development of delignification
and lignin-cellulose selectivity during ozone bleaching. J. Pulp Pap. Sci., 24, 4,
111-115.
Grubisic, A., P. Rempp and H. A. Benoit (1967). A universal calibration for gel
permeation chromatography. Polym. Lett., 5, 753-759.
Guiata, M., O. Chiantore and M. P. Luda (1990). Monte Carlo simulations of polymer
degradations 1. Degradations without volatilization. Macromolecules, 23, 7, 20872095.
Gurnagul, N., D. H. Page and M. G. Paice (1992). The effect of cellulose degradation on
the strength of wood pulp fibres. Nord. Pulp Pap. Res. J., 7, 3, 152-154.
Gustavsson, C. A. S. and W. W. Al-Dajani (2000). The influence of cooking conditions
on the degradation of hexenuronic acid, xylan, glucomannan, and cellulose during
kraft pulping of softwood. Nord. Pulp Pap. Res. J., 15, 2, 160-167.
Haas, D. W., B. F. Hrutfiord and K. V. Sarkanen (1967). Kinetic study of the alkaline
degradation of cotton hydrocellulose. J. Appl. Polym. Sci., 11, 587-600.
Hansson, J.-. (1970). Sorption of hemicelluloses on cellulose fibres. Part 2. Sorption of
glucomannan. Holzforschung, 24, 3, 77-83.
Hortling, B., P. Frm and J. Sundquist (1994). Investigations of pulp components
(polysacharides, residual lignins) using an HP/SEC system with viscosimetric, RI
and UV detectors. Third European Workshop on Lignocellulosics and Pulp,
Stockholm, 256-259.
Hult, E.-L. (2001). CP/MAS 13C-NMR spectroscopy applied to structure and interaction
studies on wood and pulp fibres. PhD-thesis. Department of Pulp and Paper
Chemistry and Technology, Royal Institute of Technology, Stockholm.
Ifju, G. (1964). Tensile strength behaviour as a function of cellulose in wood. Forest
Products Journal, 14, 8, 366-372.
Johansson, E. and J. Lind (1999). Free radical degradation of fibers during HC pulp
ozonation conditions. 10th International Symposium on Wood and Pulping
Chemistry, Yokohama, Vol. 1, 308-313.

71

Johansson, M. H. and O. Samuelsson (1975). End-wise degradation of hydrocellulose


during hot alkali treatment. J. Appl. Polym. Sci., 19, 3007-30013.
Johnson, D. C. (1985). Solvents for cellulose. In: Cellulose chemistry and its
applications. T. P. Nevell and S. H. Zeronian, West Sussex, Ellis Horwood
Limited., 181-201.
Johnson, D. L. (1969). Process for strengthening swellable fibrous material with an amine
oxide and the resulting material. Patent no US 3,447,956, USA.
Karlsson, H., P.-I. Fransson and U.-B. Mohlin (1999). STFI FiberMaster. 6th
International Conference on New Available Technologies, Stockholm, 367-374.
Karlsson, O. and U. Westermark (1996). Evidence for chemical bonds between lignin and
cellulose in kraft pulps. J. Pulp Pap. Sci., 22, 10, J397-J401.
Karlsson, O. and U. Westermark (1997). The significance of glucomannan for the
condensation of cellulose and lignin under kraft pulping conditions. Nord. Pulp
Pap. Res. J., 12, 3, 203-206.
Kennedy, J. F., Z. S. Rivera, C. A. White, L. L. Lloyd and F. P. Warner (1990).
Molecular weight characterization of underivatised cellulose by GPC using
Lithium chloride-dimethylacetamide solvent system. Cell. Chem. Technol., 24, 3,
319-325.
Kerr, A. J. and D. A. Goring (1975). The ultrastructural arrangement of the wood cell
wall. Cell. Chem. Technol., 9, 536-573.
Krssig, H. A. (1993a). Cellulose structure, accessibility and
Yverdon,Switzerland, Gordon and Beach Science Publishers, 187-199.

reactivity.

Krssig, H. A. (1993b). Cellulose structure, accessibility and


Yverdon,Switzerland, Gordon and Beach Science Publishers, 167-169.

reactivity.

Kubes, G. J., B. I. Fleming, J. M. Macleod and H. I. Bolker (1983). Viscosities of


unbleached pulps. II: The G-factor. J. Wood Chem. Techn., 3, 3, 313-333.
Kubes, G. J., J. M. Macleod, B. I. Fleming and H. I. Bolker (1981). The viscosities of
unbleached alkaline pulps. J. Wood Chem. Techn., 1, 1, 1-10.
Lai, Y.-Z. (2001). Chemical degradation. In: Wood and cellulosic chemistry. D. N.-S.
Hon and N. Shiraishi, New York, Marcel Dekker., 443-512.
Lai, Y.-Z. and K. V. Sarkanen (1967). Kinetics of alkaline hydrolysis of glycosidic bonds
in cotton cellulose. Cell. Chem. Technol., 1, 517-527.

72

Laine, J. (1996). The effect of cooking and bleaching on the surface chemistry and charge
properties of kraft pulp fibres. PhD-thesis. Department of Forest Products
Technology, Helsinki University of Technology, Helsinki.
Lin, C.-H., A. H. Conner and C. G. Hill (1993). The heterogeneous, dilute-acid hydrolysis
of cellulose. In: Cellulosics: chemical and material aspects. J. F. Kennedy, G. O.
Phillips and P. A. Williams, New York, Ellis Horwood Limited., 177-182.
Lind, J., G. Mernyi and K. Wegner (1997). The chemistry of ozone during pulp
bleaching. J. Wood Chem. Technol., 17, 3, 297-326.
Lindstrm, M., F. Berthold and B. Pettersson (1999). Presentation of a novel
carbohydrate matrix system for fundamental investigations of wood cell wall
components and their chemistry. 10th International Symposium on Wood and
Pulping Chemistry, Yokohama, Vol. 1, 102-105.
Machell, G. and G. N. Richards (1957). The alkali degradation of polysacharides. Part II.
The alkali stable residue from the action of sodium hydroxide on cellulose. J.
Chem. Soc., 4500-4506.
Malcolm, E. and T. M. Grace, Eds. (1989). Alkaline pulping, Pulp and paper manufacture.
Atlanta/Montreal, TAPPI/CPPA, 45.
Marchessault, R. H. and B. G. Rnby (1959). Hydrolysis of cellulose in phosphoric acid
solution-inductive effects. Sv. Papperstidn., 62, 7, 230-240.
McCormick, C. L. and P. A. Callais (1987). Derivatisation of cellulose in lithium/chloride
and N-N-dimethylacetamide solutions. Polymer, 28, 13, 2317-2323.
McCormick, C. L. and D. K. Lichatowich (1979). Homogenous solution reactions of
cellulose, chitin, and other polysacharides to produce controlled-activity pesticide
systems. J. Polym. Sci. Pol. Lett., 17, 8, 479-484.
Miller, J. C. and J. N. Miller (1993). Statistics for Analytical Chemistry. West Sussex, Ellis
Horwood PTR Prentice Hall, 110.
Mohlin, U.-B. and C. Alfredsson (1990). Fibre deformation and its implications in pulp
characterization. Nord. Pulp Pap. Res. J., 5, 4, 172-179.
Mohlin, U.-B., J. Dahlbom and J. Hornatowska (1996). Fiber deformation and sheet
strength. Tappi J., 79, 6, 105-111.
Molin, U. (2002). Pulp strength properties-Influence of carbohydrate composition, molar
mass and crystalline structure. PhD-thesis. Department of Pulp and Paper
Chemistry and Technology, Royal Institute of Technology, Stockholm.

73

Montroll, E. W. and R. Simha (1940). Thery of depolymerization of long chain


molecules. J. Chem. Phys., 8, 721-727.
Morgenstern, B. and H.-W. Kammer (1996). Solvation of cellulose-LiCl-DMAc
solutions. Trends Polymer Sci., 4, 3, 87-92.
Nevell, T. P. (1985). Degradation of cellulose by acids, alkalis, and mechanical means.
In: Cellulose chemistry and its applications. T. P. Nevell and S. H. Zeronian, West
Sussex, Ellis Horwood Limited., 223-242.
Otty, M. H., K. M. Vrum, B. E. Christensen, M. W. Anthonson and O. Smidsrd (1996).
Preparative and analytical size-exclusion chromatography of chitosans.
Carbohydr. Polym., 31, 253-261.
Page, D. H. (1994). The structure and strength of paper. Workshop on the Effects of
Aging on Printing and Writing Papers, Philadelphia, PA, 41-49.
Poch, D. S., A. J. Ribes and D. L. Tipton (1998). Characterization of Methocel(TM):
correlation of static light-scattering data to GPC molar mass data based on pullulan
standards. J. Appl. Polym. Sci., 70, 2197-2210.
Popa, V. I. and I. Spiridon (1998). Hemicelluloses: structure and properties. In:
Polysacharides: structural diversity and functional versatility. S. Dumitru, New
York, Marcel Dekker Inc., 301.
Porath, J and P. Flodin (1959). Gel filtration: a method for desalting and group
separation. Nature, 183, 1657-1659.
Richards, G. N. (1971). Alkaline degradation. In: Cellulose and cellulose derivatives,
Part V. N. M. Bikales and L. Segal, E. Ott. New York, John Wiley & Sons. 5,
1007-1014.
Rowland, S. O. and E. J. Roberts (1972). The nature of accessible surfaces in the
microstructure of native cellulose. J. Polym. Sci. Part A-1, 10, 2447-2461.
Rydholm., S. A. (1965). Pulping Processes. New York, John Wiley & Sons Inc., 290.
Salmn, L. and A.-M. Olsson (1998). Interaction between hemicelluloses, lignin and
cellulose: structure-property relationships. J. Pulp Pap. Sci., 24, 3, 99-102.
Samuelsson, O., G. Grangrd, K. Jnsson and K. Scramm (1953). A comparison between
the degradation of cotton and pulp. Sv. Papperstidn., 56, 20, 779-784.
Sarkanen, K. V. and C. H. Ludwig (1971). Lignins-occurence, formation , structure and
reactions. New York, Wiley Interscience, 1.

74

Scallan, A. M. (1974). The structure of the cell wall of wood-A consequence of


anisotropic inter-microfibrillar bonding? Wood Sci., 6, 266-271.
Schelosky, N., T. Rder and T. Baldinger (1999). Molmassenverteilung cellulosischer
Produkte mittels Grssenausschlusschromatographie in DMAc/LiCl. Papier, 53,
12, 728-738.
Schroeder, R. L. and F. C. Haigh (1979). Cellulose and wood pulp polysacharides. Gel
Permeation chromatography analysis. Tappi J., 62, 10, 103-105.
Schult, T., T. Hjerde, O. I. Optun, P. J. Kleppe and S. Moe (2002). Characterization of
cellulose by SEC-MALLS. Cellulose, 9, 2, 149-158.
Schulze, E. (1891). Zur Kentniss der chemischen Zusammensetzung der pflanzlichen
Zellmembranen. Ber. Dtsch. Chem. Ges., 24, 2277-2287.
Secrist, R. B. and R. P. Singh (1971). Kraft pulp bleaching II. Studies on the ozonation of
chemical pulps. Tappi J., 54, 4, 581-584.
Sell, J. and T. Zimmermann (1993). Radial fibril agglomeration of the S2 on transverse
fracture surfaces of tracheids of tension-loaded spruce and white fir. Holz Roh
Werkst., 51, 384.
Sharples, A. (1957). The hydrolysis of cellulose and its relation to structure. Trans.
Faraday Soc., 53, 1003-1013.
Sjholm, E., K. Gustafsson, F. Berthold and A. Colmsj (2000a). Influence of the
carbohydrate composition on the molecular weight distributions of kraft pulps.
Carbohyd. Polym., 41, 1, 1-7.
Sjholm, E., K. Gustafsson, B. Eriksson, W. Brown and A. Colmsj (2000b).
Aggregation of cellulose in lithium chloride/N,N-dimethylacetamide. Carbohyd.
Polym., 41, 2, 153-161.
Sjholm, E., K. Gustafsson, E. Norman, T. Reitberger and A. Colmsj (2000c). Fibre
strength in relation to molecular weight distribution of hardwood kraft pulp.
Degradation by gamma irradiation, oxygen/alkali or alkali. Nord. Pulp Pap. Res.
J., 15, 4, 326-332.
Sjholm, E., K. Gustafsson, B. Pettersson and A. Colmsj (1997). Characterization of the
cellulosic residues from lithium chloride/N,N-dimethylacetamide dissolution of
softwood kraft pulp. Carbohyd. Polym., 32, 1, 57-63.
Sjstrm, E. (1993a). Wood Chemistry. London, Academic Press, 54.

75

Sjstrm, E. (1993b). Wood Chemistry. London, Academic Press, 63.


Sjstrm, E. (1993c). Wood Chemistry. London, Academic Press, 67.
Sjstrm, E. (1993d). Wood Chemistry. London, Academic Press, 7-11.
Sjstrm, E. (1993e). Wood Chemistry. London, Academic Press, 13-17.
Sjstrm, E. (1993f). Wood Chemistry. London, Academic Press, 64.
Skogsindustrierna (2003). Freningen Sveriges Skogsindustrier,
http://www.skogsindustrierna.org/oh-bilder/. Access date: 2003-03-10.
Skoog, D. A. and J. L. Leary (1992). Principles of instrumental analysis. Orlando,
Saunders College Publishing, 659.
Sookne, A. M. and M. Harris (1945). Polymolecularity and Mechanical Properties of
Cellulose Acetate. Ind. Eng. Chem., 37, 5, 478-482.
Staudinger, H. and W. Heuer (1930). ber hochpolymerere Verbindungen: Beziehungen
zwishen Viscositt und Molekulargewicht bei Poly-styrolen. Ber. Dtsch. Chem.
Ges., 63, 11, 222-234.
Staudinger, H. and F. Reinecke (1938). Ueber makromolekulare Verbindungen. Der
Papierfabrikant, 36, 489-495.
Striegel, A. M. (1997). Theory and applications of LiCl/DMAc in the analysis of
polysacharides. Carbohydr. Polym., 34, 267-264.
Striegel, A. M. and J. D. Timpa (1996). Gel permeation chromatography of
polysacharides using universal caibration. Int. J. Polym. Anal. Char., 2, 213-220.
Strlic, M., J. Kolar, M. Zigon and B. Pihlar (1998). Evaluation of size-exclusion
chromatography and viscometry for the determination of molecular masses of
oxidized cellulose. J. Chromatogr. A, 805, 93-99.
Sundquist, J. and T. Rantanen (1983). Characterisation of technical pulps by carbanilation
of cellulose. Pap. Puu, 65, 11, 733-737.
Tanford, G. (1961). Physical chemistry of macromolecules. New York, John Wiley & Sons
Inc., 612-618.

76

Tanigawa, M., M. Suzuto, K. Fukudome and K. Yamaoka (1996). Changes in molecular


weights and molecular weight distributions of differently stranded nucleic acids
after sonication: gel permeation chromatography/low angle laser light scattering
evaluation and computer simulation. Macromolecules, 29, 23, 7418-7425.
Tayal, A. and S. A. Khan (2000). Degradation of a water-soluble polymer: molecular
weight changes and chain scission characteristics. Macromolecules, 33, 26, 94889493.
Teder, A. and L. Olm (1981). Extended delignification by combination of modified kraft
pulping and oxygen bleaching. Pap. Puu, 63, 4a, 317-322.
Teleman, A., V. Harjunp, M. Tenkanen, J. Buchert, T. Hausalo, T. Drakenberg and T.
Vourinen
(1995).
Characterisation
of
4-deoxy-beta-L-threo-hex-4enopyranosyluronic acid attached to xylan in pine kraft pulp and pulping liquor by
1H and 13C NMR spectroscopy. Carbohydr. Res, 272, 1, 55-71.
Theander, O. and E. A. Westerlund (1986). Studies on dietary fiber. 3. Improved
procedures for analysis of dietary fiber. J. Agr. Food. Chem., 34, 2, 330-336.
Tobita, H. (1995). Simulation model for the modification of polymers via crosslinkning
and degradation. Polymer, 36, 13, 2585-2596.
Tobita, H. (2001). Simultaneous long-chain branching and random scission: I. Monte
Carlo simulation. J. Polym. Sci. Pol. Phys., 39, 4, 391-403.
Turbak, A. F., A. El-Kafrawy, F. W. Snyder and A. B. Auerbauch (1981). Solvent systems
for cellulose. Patent no US 4,302,252, USA.
van Lierop, B., A. Skothos and N. Liebergott (1996). Ozone delignification. In: Pulp
bleaching-principles and practice. C. W. Dence and D. W. Reeve, Atlanta, TAPPI
Press., 323.
Westermark, U. and K. Gustafsson (1994). Molecular size distribution of wood polymers
in birch kraft pulp. Holzforschung, 48, 146-150.
Wickholm, K. (2001). Structural elements in native cellulose. PhD-thesis. Department of
Pulp and Paper Chemistry and Technology, Royal Institute of Technology,
Stockholm.
Wickholm, K., E.-L. Hult, P. T. Larsson, T. Iversen and H. Lennholm (2001).
Quantification of cellulose forms in complex cellulose materials: a chemometric
model. Cellulose, 8, 2, 139-148.

77

Viebke, J., E. Elble and U. W. Gedde (1996). Degradation of polyolefin pipes in hot
water applications: simulation of the degradation process. Polym. Eng. Sci., 36, 4,
458-466.
Wilkes, A. G. (2001). The viscose process. In: Regenerated cellulose fibres. C.
Woodings, Cambridge, Woodhead Publishing Limited., 39.
Wood, B. F., A. H. Conner and C. G. Hill (1986). The effect of precipitation on the
molecular weight distribution of cellulose tricarbanilate. J. Appl. Polym. Sci, 32,
3703-3303.
Wood, B. F., A. H. Conner and C. G. Hill (1989). The heterogeneous character of the
dilute acid hydrolysis of cellulose. J. Appl. Polym. Sci, 37, 1373-1394.
Woodings, C. (2001). A brief history of regenerated cellulosic fibers. In: Regenerated
cellulose fibres. C. Woodings, Cambridge, Woodhead Publishing Limited., 14-19.
Wyatt, P. J. (1993). Light scattering and the absolute characterization of
macromolecules. Anal. Chim. Acta.,, 272, 1-40.
Yau, W. W., J. J. Kirkland and D. D. Bly (1979). Modern Size-Exclusion Liquid
Chromatography: Practice of Gel Permeation and Gel Filtration Chromatography.
New York, John Wiley & Sons Inc., 285-313.
Yllner, S. and B. Enstrm (1956). Studies of the adsorption of xylan on cellulose fibres
during the sulphate cook, Part 1. Sv. Papperstid., 59, 6, 229-232.
Yllner, S. and B. Enstrm (1957). Studies of the adsorption of xylan on cellulose fibres
during the sulphate cook Part 2. Sv. Papperstidn., 60, 15, 549-554.
Zhang, X. Z., G. J. Kang, Y. Ni, A. R. P. van Heiningen, A. Mislankar, A. Darabie and D.
Reeve (1998). Initial delignification and cellulose degradation of conventional and
ethanol-assisted ozonation. J. Wood Chem. Techn., 18, 2, 129-157.
Zhang, X. Z., Y. Ni and A. van Heiningen (1999). Kinetics of cellulose degradation
during ozone bleaching. 85th Annual meeting, PAPTAC, Vol. A Preprints, A307A313.
Zimm, B. (1948). Apparatus and methods for measurement and interpretation of the
angular variation of light scattering; preliminary results on polystyrene solutions.
J. Chem. Phys., 16, 12, 1099-1116.
stlund, I. (2001). Aspects of carbohydrate degradation in soda-AQ cook. The influence of
lignin and anthraquinone. M.Sc.-thesis., Chalmers University, Gothenburg.

78

Appendix I.
Data of pulps used in this work.

Hardwood pulps
M / L 1)

Paper
no

Bleaching
sequence

Intrinsic
viscosity (ml/g)

Kappa
no

I, IV

Unbleached

1160

15

O D EOP DD

820

NA

II

Unbleached

1310

17

IV

Unbleached

1300

17

IV

Unbleached

1100

11

IV

Unbleached

1260

13

IV

Unbleached,

1180

16

IV

Unbleached

1090

15

IV

930

10

IV

DEDED

1150

<1

Unbleached

NA

23

Softwood pulps
L

II

Unbleached

950

13

II

Unbleached

1080

18

II

Unbleached

1230

31

II

Unbleached

1270

41

II

Unbleached

1330

53

Cotton linters
M

920

NA

2080

NA

1) M and L denote pulps produced in a mill or in the laboratory, respectively.


2) NA: not analyzed

79

Appendix II
Pulping conditions for experiments in paper III.
Cooking conditions for birch
Cooking time at

[HS-]

[OH-]

Residual [OH-]

170C (hours)

(mol/l)

(mol/l)

(mol/l)

BT1

0.85

0.3

1.2

0.56

BT2

2.1

0.3

1.2

0.47

BT3

3.6

0.3

1.2

0.42

BT4

5.2

0.3

1.2

0.37

BT5

7.3

0.3

1.2

0.33

BT6

10

0.3

1.2

0.30

BA1

1.0

0.3

0.8

0.20

BA2

1.0

0.3

1.2

0.52

BA3

1.0

0.3

1.8

1.04

BA4

1.0

0.3

2.5

1.59

BA5

1.0

0.3

3.2

2.19

BA6

1.0

0.3

4.0

2.89

Cooking time at

[HS-]

[OH-]

Residual [OH-]

170C (hours)

(mol/l)

(mol/l)

(mol/l)

ST1

1.5

0.3

1.0

0.46

ST2

2.5

0.3

1.0

0.42

ST3

3.75

0.3

1.0

0.39

ST4

5.25

0.3

1.0

0.36

ST5

8.0

0.3

1.0

0.30

Pulp

Cooking conditions for spruce.


Pulp

ST6

21.5

0.3

1.0

0.18

SA1

2.0

0.3

0.7

0.18

SA2

2.0

0.3

1.0

0.44

SA3

2.0

0.3

1.4

0.75

SA4

2.0

0.3

1.9

1.12

SA5

2.0

0.3

2.6

1.69

SA6

2.0

0.3

3.5

2.21

81

You might also like