Download as pdf or txt
Download as pdf or txt
You are on page 1of 25

This article was downloaded by: [Georgia Tech Library]

On: 11 November 2014, At: 15:50


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954
Registered office: Mortimer House, 37-41 Mortimer Street, London W1T
3JH, UK

Journal of Wood Chemistry


and Technology
Publication details, including instructions for
authors and subscription information:
http://www.tandfonline.com/loi/lwct20

An Innovative Approach
Characterising the
Interactions Leading to Pitch
Deposition
a a a
David Vercoe , Karen Stack , Adrian Blackman
a b
, Brian Yates & Desmond Richardson
a
School of Chemistry , University of Tasmania ,
Private Bag 75, Hobart, Tasmania, 7001, Australia
b
Research and Development Australasia ,
Norske‐Skog Paper Mills (Australia) Limited ,
Boyer, Tasmania
Published online: 24 Aug 2007.

To cite this article: David Vercoe , Karen Stack , Adrian Blackman , Brian Yates &
Desmond Richardson (2005) An Innovative Approach Characterising the Interactions
Leading to Pitch Deposition, Journal of Wood Chemistry and Technology, 24:2,
115-137, DOI: 10.1081/WCT-200026562

To link to this article: http://dx.doi.org/10.1081/WCT-200026562

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all
the information (the “Content”) contained in the publications on our
platform. However, Taylor & Francis, our agents, and our licensors
make no representations or warranties whatsoever as to the accuracy,
completeness, or suitability for any purpose of the Content. Any opinions
and views expressed in this publication are the opinions and views of
the authors, and are not the views of or endorsed by Taylor & Francis.
The accuracy of the Content should not be relied upon and should be
independently verified with primary sources of information. Taylor and
Francis shall not be liable for any losses, actions, claims, proceedings,
demands, costs, expenses, damages, and other liabilities whatsoever
or howsoever caused arising directly or indirectly in connection with, in
relation to or arising out of the use of the Content.

This article may be used for research, teaching, and private study
purposes. Any substantial or systematic reproduction, redistribution,
reselling, loan, sub-licensing, systematic supply, or distribution in any form
to anyone is expressly forbidden. Terms & Conditions of access and use
can be found at http://www.tandfonline.com/page/terms-and-conditions
Downloaded by [Georgia Tech Library] at 15:50 11 November 2014
JOURNAL OF WOOD CHEMISTRY AND TECHNOLOGY
Vol. 24, No. 2, pp. 115–137, 2004

An Innovative Approach Characterising the


Downloaded by [Georgia Tech Library] at 15:50 11 November 2014

Interactions Leading to Pitch Deposition

David Vercoe,1, * Karen Stack,1 Adrian Blackman,1


Brian Yates,1 and Desmond Richardson2
1
School of Chemistry, University of Tasmania, Hobart, Tasmania
2
Research and Development Australasia, Norske-Skog Paper Mills
(Australia) Limited, Boyer, Tasmania

ABSTRACT

In the production of mechanical pulp for paper manufacture, resinous


material is released from the wood. This resinous material forms colloidal
dispersions which can agglomerate and deposit onto surfaces. These
deposits are known as pitch. Pitch has been cited in the literature as
early as the 1920s but there is still much more to learn about the inter-
actions that occur in its formation. We know what the major components
of pitch are and how these wood extractives can vary with type, location
and age of the tree as well as seasonal effects. It is known and cited in the
literature that the four major components of pitch are fatty acids (namely
those of chain length C12 – C20), resin acids, triglycerides, and steryl
esters. Due to the complexity of a four component system it was

*Correspondence: David Vercoe, School of Chemistry, University of Tasmania, Private


Bag 75, Hobart, Tasmania 7001, Australia; E-mail: dqvercoe@postoffice.utas.edu.au.

115

DOI: 10.1081/WCT-200026562 0277-3813 (Print); 1532-2319 (Online)


Copyright # 2004 by Marcel Dekker, Inc. www.dekker.com
116 Vercoe et al.

decided to focus the initial study on a two component system, namely that
of a resin and a fatty acid. This paper will attempt to explain the chemical
interactions that occur at a molecular level between these two wood
extractive components. The fatty acids were divided into two studies,
involving the chain length and degrees of saturation, and the resin acids
were separated into three classes; aromatic class (dehydroabietic acid),
conjugated diene class (abietic acid), and alkene class (isopimaric
acid). The interaction has been studied using deposition studies and mol-
ecular modelling. Trends in classes of each component were contrary to
Downloaded by [Georgia Tech Library] at 15:50 11 November 2014

our initial theory that the amount of pitch deposited was related to the
solubility of the organic compounds in water. Further studies indicated
that a physical interaction in the form of hydrogen bonding was occurring
between the resin acid and fatty acid and that the method in which the
molecules interact with each other and the surrounding water molecules
affect deposition. Differences in deposition behaviour was found when
the type of resin acid, the fatty acid chain length, and the fatty acid’s
degree of unsaturation were varied. It is proposed that the aromatic
class of resin acid (dehydroabietic acid) forms hydrogen bonds over the
aromatic ring, whereas the non-aromatic classes of resin acid simply
dimerise at the carboxylic head groups. These differences in orientation
of the molecules affects the interaction of the rest of the molecule with
the surrounding water molecules and hence its solubility and deposition
abilities.

Key Words: Wood extractives; Pitch; Resin acids; Fatty acids; Gas
chromatography; Computational molecular modelling; Hydrogen
bonding.

INTRODUCTION

Paper manufacture from softwood derived mechanical pulp releases resi-


nous extractive material from the wood. This extractive material is a hetero-
genous group of substances with variable chemical behaviour. In principle,
the components can be grouped into two main classes: lipophilic (fatty) and
hydrophilic substances (Table 1).[1]
By nature, lipophilic substances are hydrophobic and tend to form col-
loidal dispersions which can agglomerate and deposit on surfaces. These
deposits are known as agglomerated pitch, and often cause problems in both
the pulp and paper mill with both the product quality and production rates
being negatively affected in a variety of ways. Pitch can decrease the effi-
ciency of pulp washing, screening, centrifugal cleaning, and refining, and it
can disrupt many paper machine operations.[2] It can also cause dirt, holes,
picking, and scabs in the final sheet.[2] Economic losses associated with
Interactions Leading to Pitch Deposition 117

Table 1. Lipophilic and hydrophilic


extractives.[1]

Lipophilic Hydrophilic

Resin acids Phenols


Free fatty acids Lignans
Triglycerides Flavanoids
Steryl esters Tannins
Free sterols Sugars
Downloaded by [Georgia Tech Library] at 15:50 11 November 2014

Monoterpenes Salts

pitch problems in kraft mills often amount to 1% of sales.[3] Research into the
pitch problems has been cited as early as 1891[4 – 9] but there is still much to
learn about the interactions that occur leading to the formation of pitch. It is
known how quantity and proportions of pitch components can vary over
season, genus, geographic location, and age of tree.[10 – 12] The mean percent-
age composition of extractives from Pinus radiata can be seen in Table 2.
There are eight common resin acids, the structures of which are shown in
Fig. 1. They can be divided into three groups; aromatic (dehydroabietic acid),
conjugated diene class (abietic acid, levopimaric acid, palustric acid, and
neoabietic acid), and alkene class (isopimaric acid, pimaric acid, and sandar-
acopimaric acid).
Straight chain saturated and unsaturated fatty acids with 16– 24 carbon
atoms (Fig. 2) are the most important fatty acids in wood plants with the
C18 mono-, di-, and tri-unsaturated acids dominating.[14]
Due to the complexity of pitch it was decided to focus this study on a
model pitch composed of two components, namely that of a resin acid and
a fatty acid. The effects of the resin and fatty acids’ chemical structure were
investigated in order to determine how changing the chemical structure may
affect pitch deposition and also the interaction between the components.
Our initial hypothesis was that the deposition of pitch was related to the solu-
bility of the components but research soon indicated that this was simply not
the case and that interactions were occurring between different molecular

Table 2. Composition of P. radiata wood extractives (%).[13]

Resin acid Fatty acid Triglycerides Other

Sapwood 29 8 36 27
Heartwood 64 3 3 30
118 Vercoe et al.
Downloaded by [Georgia Tech Library] at 15:50 11 November 2014

Figure 1. Eight common resin acids.

types. This paper will attempt to explain the chemical interactions that occur at
a molecular level between these pitch components.
The fatty acids were divided into two groups, one involving the length of
the alkyl chain and the second concerning the degrees of saturation. Each
group was studied with the three classes of resin acid. Deposition studies
and molecular modelling were carried out to investigate the interactions.
Interactions Leading to Pitch Deposition 119
Downloaded by [Georgia Tech Library] at 15:50 11 November 2014

Figure 2. Structures of common fatty acids.

RESULTS

Deposition Studies

To ensure that the concentration of the two components was high enough
to guarantee interaction, it was decided to conduct these deposition studies
above the critical micelle concentration (cmc). Micelles form only above the
cmc and are detected by noting a pronounced discontinuity in physical proper-
ties of the solution such as the surface tension[15] (Fig. 3). It was found that the
highest cmc for our studies was that for lauric acid and was approximately
9 mg/L. All deposition work was, therefore, done above that concentration

Figure 3. The typical variation of surface tension of an aqueous solution of lauric


acid close to the cmc.
120 Vercoe et al.

in our studies ranging from 12 to 15 mg/L. It was also questionable as to


whether some of the fatty acids with a higher chain length such as arachidic
acid (C20:0) would not form micelles but prefer to precipitate out of solution.
It was found that although the less soluble fatty acids did precipitate out of sol-
ution at higher concentrations a cmc was found at approximately 3.5 mg/L.
Two different studies were conducted investigating the fatty acid vari-
ables; first, the effects of chain length (C12:0 –C20:0) and second, the
degrees of saturation (C18:0 – C18:3).
Each fatty acid was dialysed to remove any solvent present and then
Downloaded by [Georgia Tech Library] at 15:50 11 November 2014

placed in a sealed polyethylene (PE) jar in a 508C water bath with constant
agitation for 2 hr. The organic components were extracted before and after
the agitation and concentrations were calculated using gas chromatography
(GC). The amount of pitch deposited was then calculated using the difference
between the two measurements (Fig. 4). The above method was validated by
extracting the deposit from the walls of the jar and running the sample on the
GC and then comparing it to the amount calculated by difference (Table 3).
The results in Table 3 indicate that the difference calculation was a valid
method for determining the amount of pitch deposited.
Figure 5 shows the GC chromatogram of the model colloidal wood resin
before deposition, after deposition, and of the wall extracts. The results indicate
that the pitch deposited is a result of physical interaction between the two com-
ponents and that no other components were formed through chemical means.
As stated earlier, our initial hypothesis was that the variability in the
amount of pitch deposited was simply directly related to the solubility of
the components present. Figure 6 indicates the solubility of the fatty acids
with varying chain lengths that are present in wood plants.[16] It can be seen
that the solubility decreases quite considerably as the fatty acid chain length
increases with the higher chain lengths being virtually insoluble in aqueous
conditions.

Figure 4. Schematic diagram of jar depositions.


Interactions Leading to Pitch Deposition 121

Table 3. Method validation data.

Concentration Concentration Concentration


before deposition in solution Accumulation of deposit on
(IN) (mg/L) (OUT) (mg/L) (IN – OUT) walls (mg/L)

Resin 19.62 15.49 4.13 4.09


acid 20.48 15.98 4.50 3.98
Fatty 11.38 7.94 3.44 3.51
acid 13.26 9.36 3.90 3.17
Downloaded by [Georgia Tech Library] at 15:50 11 November 2014

Note: Resin acid used was dehydroabietic acid and fatty acid was C18:0.

Table 4 shows the solubility of the separate resin acids under aqueous
conditions.[17] It can be seen that isopimaric acid has the lowest solubi-
lity (1.70 mg/L) with dehydroabietic acid having the highest solubility
(5.11 mg/L). The other resin acid used in our study, abietic acid, has a solu-
bility of similar order to isopimaric acid (2.75 mg/L).

Figure 5. GC chromatogram of before deposition, after deposition, and wall extracts.


1, Silyl ester of pentadecanoic acid (spike); 2, silyl ester of heptadecanoic acid (IS); 3,
silyl ester of stearic acid; 4, silyl ester of dehydroabietic acid; 5, cholesteryl stearate
(spike); 6, 1,3-dipalmitoyl-2-oleoyl-glycerol (spike); 7, triolein (spike).
122 Vercoe et al.
Downloaded by [Georgia Tech Library] at 15:50 11 November 2014

Figure 6. Solubility of fatty acids (C12 – C18) in water.[16]

If the amount of pitch deposited was related to the solubility of the fatty
acid, then an inverse relationship between solubility and deposition would be
expected. To investigate if the above statement was correct a single com-
ponent study was conducted with only the fatty acid placed into the jar. As
expected an inverse relationship was seen (Fig. 7) with the total pitch depos-
ited (TPD), or the precipitate, increasing with the increased chain length.
We then investigated a two component system, namely that of a fatty acid
and resin acid. To help understand how each of these variables behaved with
each class of resin acid, one was chosen from each class. The resin acid from
each class was chosen based on purity, availability, and cost and were dehy-
droabietic acid (aromatic), abietic acid (conjugated diene) and isopimaric acid

Table 4. Solubility of resin acids in water.[17]

Solubility
Resin acid (mg/L)

Isopimaric acid 1.70


Sandaracopimaric acid 1.82
Pimaric acid 2.17
Palustric acid 2.41
Neoabietic acid 2.31
Levopimaric acid 2.54
Abietic acid 2.75
Dehydroabietic acid 5.11
Interactions Leading to Pitch Deposition 123
Downloaded by [Georgia Tech Library] at 15:50 11 November 2014

Figure 7. Effect of fatty acid chain length on amount of fatty acid deposited (error
bars indicate standard deviation).

(alkene). When looking at the %TPD in a two-component system (Fig. 8), it


appeared to be directly proportional to the solubility of the fatty acid. The
results indicate that deposition of the two-component system is not inversely
related to the solubility of the fatty acids and that an interaction is occurr-
ing between the two components, affecting their stability in solution as
well as their deposition. The results also show that the aromatic resin acid,

Figure 8. Effect of fatty acid chain length on TPD.


124 Vercoe et al.

dehydroabietic acid, had less deposition than the non-aromatic resin acids
which may be either related to its solubility or suggesting that the aromatic
resin acid may be interacting differently with the fatty acid but further inves-
tigations would be required to confirm this.
The effect, on pitch deposition, of the saturation of the fatty acid was also
investigated (Fig. 9). Mono-, di-, and tri-unsaturated fatty acids with 18 carbon
atoms were chosen as they dominate the types of fatty acids found in woody
plants.[14] These deposition results also indicated that an interaction was
occurring between the resin acid and the fatty acid and that an increase in
Downloaded by [Georgia Tech Library] at 15:50 11 November 2014

the number of double bonds in the fatty acid resulted in greater deposition.
The results also show that, as with the fatty acid chain length deposition
study, the aromatic resin acid behaved differently to the non-aromatic
resin acids.
Although these results do indicate that an interaction is occurring between
the resin acid and fatty acid, it does not give any insight into how or where
these interactions occur.
To gain a better insight into the nature of the interactions occurring, com-
puter molecular modelling was employed. Before, we could model the inter-
action the molar ratio of the components in the deposit was required. By
observing from the GC analysis results (Table 5) the amount of each indivi-
dual component deposited it could be seen that the molar ratio of the pitch
components deposited was 1 : 1. It can also be noted that this molar ratio of
1 : 1 was observed for all the fatty acids and resin acids investigated. This
meant that when the deposits were modelled one molecule of each of the
fatty acids and resin acids was required.

Figure 9. Effect of fatty acid saturation on TPD.


Downloaded by [Georgia Tech Library] at 15:50 11 November 2014

Table 5. Typical deposition Results (varying fatty acids with dehydroabietic acid) indicating a 1 : 1 molar ratio.

Fatty acid Molecular Dehydroabietic Molecular Molar


Fatty deposited weight Number acid deposited weight Number ratio
acid (mg) (g/mol) of moles (mg) (g/mol) of moles FA : RA
Interactions Leading to Pitch Deposition

C12:0 4.15 200.32 0.0207 6.46 300.44 0.0215 0.9628


C14:0 4.97 228.37 0.0218 6.45 300.44 0.0215 1.0140
C16:0 5.49 256.43 0.0214 6.46 300.44 0.0215 0.9953
C18:0 6.18 284.48 0.0217 6.67 300.44 0.0222 0.9775
C20:0 6.52 312.54 0.0209 6.34 300.44 0.0211 0.9905
125
126 Vercoe et al.

Molecular Modelling

Computer molecular modelling was used to help understand the inter-


actions between the two components at a molecular level. Though molecular
modelling had not been applied in this area of research before it was thought
to be a useful method to investigate hydrogen bonding and any possible
p – p stacking interactions. Due to the difference in behaviour with the aro-
matic and non-aromatic resin acids in deposition studies it was hypothesised
that the method by which the fatty acid and resin acid were interacting must
Downloaded by [Georgia Tech Library] at 15:50 11 November 2014

be different for these two classes of resin acids. Two possibilities were con-
sidered for this difference; (1) p – p stacking over the aromatic ring and the
double bonds of the fatty acid chain or (2) difference in the formation of the
hydrogen bonds. There are four possible variances in the mode of hydrogen
bonding (Fig. 10). These are (1) the well published dimer form[18 – 21] in
which hydrogen bonding occurs between two carboxylic acid head groups
[Fig. 10(a)]. (2) The dimer with the a-proton[20 – 22] [Fig. 10(b)], this form
of hydrogen bonding has been noted with the acetic acid dimer and is
common with short chain fatty acids in aqueous conditions. (3) The
catemer motif[21,23] [Fig. 10(c)], this form of interaction occurs where the
R group is sufficiently small and, therefore, does not obstruct the optimum
packing. This preferred hydrogen-bonding motif is presumed to reflect
energetic differences between the two hydrogen-bonding geometries. (4) A
hydrogen bond between the carboxylic acid head group and an electron
rich region of a molecule such as an aromatic ring[24,25] [Fig. 10(d)].
An important implication of the aromatic incidence is that the aromatic
must be a good electron donor but not necessarily a highly electronegative
group.

Figure 10. Different forms of hydrogen bonding.


Interactions Leading to Pitch Deposition 127

Molecular mechanics (MM3) and molecular orbital calculations (PM3)


were used to study the binding of the two components in the gaseous
state. It can be noted that even though these calculations were conducted
in the gaseous state they do give us an understanding of the interactions in
an aqueous solution. Future work to complete this understanding could
include looking at solvation effects and/or modelling the components with
water molecules present in the calculation. An extra parameter was created
to increase the strength and probability of hydrogen bonding to occur at
the acidic proton of the carboxylic acids as the parameter originally involved
Downloaded by [Georgia Tech Library] at 15:50 11 November 2014

in these calculations was not recognising the occurrence of hydrogen bonds.


The MM3 method is used to calculate the different conformers that could
occur and in most cases the lowest energy conformer was chosen and a
molecular orbital calculation was undertaken on it. The only time that the
lowest energy conformer was not chosen was for the homogenous dehydro-
abietic dimer as the lowest energy did not contain any hydrogen bonding.
In this instance, the second lowest conformer was chosen and this was accep-
ted as the true dimer as it contained the desired hydrogen bonding and only
differed in heat of formation by 2.17 kJ/mol from the lowest conformer.
Through the modelling of the complexes in a 1 : 1 ratio it was found that
the formation of a dimer predominated over any other type of hydrogen
bonding in the non-aromatic systems, whereas the aromatic system preferred
to form their hydrogen bonds over the electron rich aromatic ring (Fig. 11).
This difference in hydrogen bonding could help to explain the differences in
behaviour between the classes of resin acids in the deposition studies.
To see if there were any correlations between the deposition and modelling
data, the stabilisation energies of the 1 : 1 complexes were considered. If we con-
sider the reaction pathway of the two components in the aqueous solution the
DHf varies for the complex and for the single components in solution.
Figure 12 shows the reaction pathway for two components A and B and two
different complexes between them. The difference in energy between the
complex and the two single components is the stabilisation energy. A lower
complex energy leads to an increase in stabilisation energy which in turn leads
to a more stable complex and hence an increased probability in its occurrence.
The difference in DHf between the MM3 and PM3 calculations was
minimal with the same trends following for each method. These results can
be seen in Table 6 with the DHf values for the single components. It was
decided that the PM3 data would be used as it is a more accurate mathematical
calculation for chemical structures.
The stabilisation energies and %TPD of the 1 : 1 complexes were
plotted as a function of fatty acid chain length (Fig. 13) and degrees of
saturation (Fig. 14) with different resin acids. Figure 13 shows that as the
fatty acid chain length increases, deposition decreases and the stabilisation
128 Vercoe et al.
Downloaded by [Georgia Tech Library] at 15:50 11 November 2014

Figure 11. Position of hydrogen bonds in aromatic and non-aromatic systems.


Interactions Leading to Pitch Deposition 129
Downloaded by [Georgia Tech Library] at 15:50 11 November 2014

Figure 12. Reaction pathway for components A and B with formation of two com-
plexes for AB.

energy of the complex also decreases. Likewise, Fig. 14 shows that as the
degree of unsaturation increases, deposition and the stabilisation energy
both increase.
If the pitch is preferentially depositing in a 1 : 1 ratio based on the fatty
acid and resin acid molar ratios, then the stabilisation energies of the most
stable complex must be less than that of the fatty acids and resin acids

Table 6. DHf values for single components using MM3 and PM3
calculations.

DHf (kJ/mol)

MM3 PM3

C12:0 2654.50 2644.59


C14:0 2703.96 2690.31
C16:0 2721.95 2750.27
C18:0 2789.90 2778.64
C18:1 2628.81 2689.46
C18:2 2568.81 2577.15
C18:3 2538.44 2497.37
C20:0 2842.78 2819.80
Abietic acid 2359.36 2477.04
Isopimaric acid 2377.31 2476.35
Dehydroabietic acid 2435.14 2484.72
130 Vercoe et al.
Downloaded by [Georgia Tech Library] at 15:50 11 November 2014

Figure 13. Effect of fatty acid chain length on stabilisation energy and deposition.

forming a complex together in a homogenous relationship. This can be seen in


Table 7 with the stabilisation energies of the heterogenous complexes being
higher than that of the homogenous complexes.

DISCUSSION

The results from both the deposition and modelling studies indicated that
there was an interaction occurring between the fatty acid and resin acid. When

Figure 14. Effect of fatty acid saturation on stabilisation energy and deposition.
Interactions Leading to Pitch Deposition 131

Table 7. Stabilisation energies for homo- and heterogenous complexes.

Stabilisation
Complex energy (kJ/mol)

Dehydroabietic acid/dehydroabietic acid 41.21


C18:0/C18:0 2254.93
Dehydroabietic acid/C18:0 53.15
Abietic acid/abietic acid 47.69
Abietic acid/C18:0 75.95
Downloaded by [Georgia Tech Library] at 15:50 11 November 2014

referring to the molecular modelling results, it must be understood what the


stabilisation energy results indicate. A higher stabilisation energy indicates
a stronger interaction between the two components and hence, stronger hydro-
gen bonding between them. This increase in hydrogen bond strength reduces
the component – water interaction which in turn affects the solubility of the
complex in water.
If we consider the interaction between the resin acid and the fatty acids of
varying chain lengths, we notice that as the chain length increases the stabili-
sation energy of the system decreases. This decrease in energy indicates that
the hydrogen bond length is greater between the two components which increa-
ses the potential for interaction with water and hence its solubility. In addition,
when we increased the number of double bonds in the system, the complex
became tighter and the hydrogen bond distance decreased (Table 8) causing
less interaction with the aqueous environment and lowering its solubility.
The deposition results also indicated in both studies that the aromatic resin
acid complex had lower deposition showing that it was more soluble in an
aqueous environment than the non-aromatic complexes. This is reinforced
by observing the computer model of this complex as it has a lower stabilisation
energy and, therefore, weaker hydrogen bonds with more hydrophilic regions
exposed to the surrounding environment due to the fact that the interaction
between the fatty acid and resin acid is occurring over the aromatic ring.

Table 8. Intermolecular distances for dehydroabietic acid and fatty acid C18:x.

C18:0 C18:1 C18:2 C18:3

Distance between two 4.758 Å 4.536 Å 4.121 Å 3.920 Å


molecules at point of
unsaturation in fatty acid
Hydrogen bond length 3.276 2.400 Å 2.417 Å 2.236 Å
132 Vercoe et al.

The modelling results have shown that the hydrophilic and hydrophobic
ends of the fatty acid and resin acid molecules could be oriented to assist in
solubility with the aromatic systems having an increased amount of hydro-
philic regions available due to the position of the hydrogen bond over the
aromatic ring. This could also be seen in the modelling results of the
dehydroabietic acid dimer (Fig. 15) and this provides an explanation as to
its increased solubility over the non-aromatic resin acids.
Downloaded by [Georgia Tech Library] at 15:50 11 November 2014

CONCLUSIONS

It can be concluded that the explanation for pitch deposition in paper mills
is not a simple one. Initial thoughts were that the deposition of these agglo-
merates of wood resins were related to the solubility of the individual com-
ponents but our work has indicated that this is not the case. We have shown
that in the deposition of pitch, there is an interaction between the components
through hydrogen bonding which affects the energy of the system, and, there-
fore, the amount that is precipitated out of solution. It can also be concluded
that the structure of the resin acid and the fatty acid chain length and degrees

Figure 15. Dehydroabietic acid dimer in an aqueous environment.


Interactions Leading to Pitch Deposition 133

of saturation affect deposition. Results showed that the non-aromatic class of


resin acid gave more deposition that the aromatic class due to differences in
how they interact and form hydrogen bonds. This affects the orientation of
the molecules and hence their interaction with the aqueous environment. Aro-
matic resin acids were found to form hydrogen bonds with the fatty acids over
the electron rich aromatic ring which enables the free hydrophilic groups of the
components to interact with water molecules and enable it to be slightly more
soluble than the non-aromatic resin acids. Shorter chain lengths and less satu-
rated fatty acid systems were found to form more stable complexes with the
Downloaded by [Georgia Tech Library] at 15:50 11 November 2014

resin acids. They were also found to cause more pitch to be deposited and, there-
fore, would be predicted to lead to reduced mill efficiency and productivity.

EXPERIMENTAL

Chemicals

All resin acids (Table 9) were purchased from Helix Biotech and all fatty
acid samples were purchased from Sigma Aldrich. All the samples’ purity was
checked by GC.

Colloidal Preparation—Dialysis

In order to study the deposition of pitch from its aqueous colloidal form to
its agglomerated form additional laboratory steps were needed to be taken
to ensure that the beginning, or pre-deposition, samples were of the aqueous

Table 9. Standard compounds used and their purity.

CAS # Compound Molecular formula % Purity

1740-19-8 Dehydroabietic acid C20H28O2 99 þ


514-10-3 Abietic acid C20H30O2 90 – 95
5835-26-7 Isopimaric acid C20H30O2 99 þ
143-07-7 Lauric acid C12H24O2 99 þ
544-63-8 Myristic acid C14H28O2 99 þ
57-10-9 Palmitic acid C16H32O2 99 þ
57-11-4 Stearic acid C18H36O2 99 þ
112-80-1 Oleic acid C18H34O2 99 þ
60-33-3 Linoleic acid C18H32O2 99 þ
463-40-1 Linolenic acid CH30O2 99 þ
506-30-9 Arachidic acid C20H40O2 99 þ
134 Vercoe et al.

colloidal form. Model pitch dispersions were prepared using a variation on


methods developed by Sundberg[26] and Stack.[10] The model pitch disper-
sions were a mixture of a fatty acid and a resin acid. The model solutions
were prepared by quickly adding acetone mixtures of the fatty acid and
resin acid, in a 1 : 1 molar ratio, to distilled water (250 mL) with stirring,
which had been adjusted to pH 5 (HNO3) and contained a slight electrolytic
residual (0.001 M KNO3). The acetone was removed by dialysis using a
cellulose membrane tubing (Sigma D-9402, 76 mm wide, .12,000 MW)
and a wash solution of distilled water, which had also been adjusted to pH 5
Downloaded by [Georgia Tech Library] at 15:50 11 November 2014

(HNO3) and contained a slight electrolytic residual (0.001 M KNO3).


The model solutions were dialyzed for 24 hr, with the wash solution being
changed every 30 min for the first 5 hr. The model solutions were then diluted
to 400 mL using fresh wash solution. The 400 mL samples were then adjusted
to pH 5.5 (KOH), which is representative of the pH encountered in pulp mills,
in order to study their depositional characteristics.

Deposition

Deposition was conducted by stirring 400 mL of the dialyzed dispersions


in PE jars using a paddle stirrer at 330 RPM (Cole Palmer, PE coated) as the
shear force generator. All depositions were conducted over a 2-hr period at a
temperature of 508C in a temperature bath.

Extraction

The model pitch components were extracted from the model pitch disper-
sions, before and after deposition, using the organic solvent, tertiary butyl
methyl ether (tBME), (Aldrich 99.8% purity HPLC grade [1634-04-4]). The
pitch-containing tBME supernatant was pipetted off the top of centrifuged
5 mL pitch dispersion aliquots to which 100 mL of internal standard (penta-
decanoic acid—Aldrich 99þ% purity [1002-84-2], heptadecanoic acid (IS)—
Sigma 99% purity [506-12-7], 1,3-dipalmitoyl-2-oleoyl-glycerol—Sigma
99% purity [2190-25-2] and cholesteryl stearate—Sigma 99% purity
[35602-69-8]) had been added. The pH was adjusted to 3.5 (HNO3) as per
Holbolm and Sjöström’s procedure.[27] This was repeated in duplicate for
each of the samples both before and after the deposition studies.

Derivatization—Silylation

The samples were blown to dryness and silylated with 100 mL of pyridine
(Aldrich 99þ% purity [110-86-1]) and 100 mL of BSA (N,O-bis(trimethyl-
silyl)-acetamide, Sigma 90% purity GC grade [10416-59-8]) followed by
Interactions Leading to Pitch Deposition 135

heating at 20 min at 608C. The GC vials were cooled to room temperature and
then filled to 1 mL with toluene (Aldrich 99.8% purity HPLC [108-88-3]).

ANALYSIS

GC Analysis

The silylated samples were then analysed using programmed injection


temperature on-column high temperature GC with a short column and a
Downloaded by [Georgia Tech Library] at 15:50 11 November 2014

flame ionisation detector (PVT-HTGC-FID).


Samples were analysed using a Varian 3800 GC equipped with a Varian
8400 autosampler. The samples were injected onto a 15 m Phenomenexw
100% polydimethylsiloxane (ZB-1, 15 m  0.53 mm ID  0.15 mm Film
Thickness) ZebronTM capillary GC column. The injector temperature was
held at 908C for the first 30 sec after injection and then ramped up to 3258C
at a rate of 2008C/min. The oven/column temperature was held at 908C for
the first 1.5 min after injection and then ramped up to 3208C at a rate of
128C/min. The FID temperature was held at 3608C for the entire duration
of the 33 min program. Helium was used as the carrier gas and was held
at a constant pressure of 3.0 psi.

Data Analysis

GC data was analysed using Varian Star 5.5 software package using hepta-
decanoic acid (Sigma 99% purity [506-12-7]) as the internal standard and 1,3-
dipalmitoyl-2-oleoyl-glycerol—Sigma 99% purity [2190-25-2] as the recovery
standard. FID correction factors were developed through the use of spikes
(petroselinic acid—Sigma 99% purity [593-39-5], dehydroabietic acid—
Helix Biotech [1740-19-8], Triolein—Sigma [122-32-7]) which were injected
at various points throughout every sample run. Pentadecanoic acid (Aldrich
99þ% [1002-84-2]) was used to verify that silylation did in fact occur.
Each sample, pre and post deposition, was extracted, silylated and
injected twice. The average of these values was used to calculate the concen-
trations of the solutions before and after deposition. The quantity of the indi-
vidual components was determined by the difference between the pre and post
deposition concentrations.

Theoretical Calculations

Initial conformer searches were carried out using the MM3 with para-
meters adjusted to allow for hydrogen bonding. These conformers were
then re-optimised with PM3 and the minimised energies noted.
136 Vercoe et al.

Surface Tension Measurements

The surface tension measurements were conducted using an Analite


surface tension meter. Solutions of varying concentrations were placed in
the sample dish and placed on the platform. This platform was then raised
until the plate just immersed into the test solution and the height control
was rotated back 0.5 turn. The surface tension was measured in mN/m.
Downloaded by [Georgia Tech Library] at 15:50 11 November 2014

ACKNOWLEDGMENTS

We would like to thank Douglas McLean for his ongoing support in the
lab and members of the computational chemistry group at the University of
Tasmania for their guidance and support. This work was funded by an Austra-
lian Research Council SPIRT Industry Grant with industry support from
Norske Skog Paper Mills (Australia) Pvt Ltd.

REFERENCES

1. Holmberg, M. Pitch and precipitate problems. In Papermaking Chem-


istry; Neimo, L., Ed.; Fapet Oy: Helsinki, 1999; Vol. 4, 222 –240.
2. Dreisbach, D.D.; Michalopoulos, D.L. Understanding the behavior of
pitch in pulp and paper mills. Tappi J. 1989, 72 (6), 129– 134.
3. Allen, L.H. Pitch control in pulp mills. In Pitch Control, Wood Resin and
Deresination; Back, E.L., Allen, L.H., Eds.; TAPPI Press: Atlanta, 2000;
265 –287.
4. Enckell, J. Rosin troubles in paper making. Pulp Paper Can. 1925, 1,
937 –940.
5. Twitchell, E. Separation of rosin from fatty acids. J. Soc. Chem. Ind.
1891, 1, 804– 805.
6. Tydén, A. Investigation of resin precipitations. Sven. Papperstidn. 1928,
31 (8), 265– 267.
7. Wells, S.D. Bentonite for pitch troubles. Paper Trade J. 1924, 79 (16),
135 –136.
8. Johnsen, B. Injurius rosin in sulphite pulp. Pulp Paper Can. 1917, 1,
577 –579.
9. Campbell, J. Resins in paper making. Pulp Paper Can. 1927, 1, 118 –121.
10. Stack, K.R.; Stevens, E.A.; Richardson, D.E.; Parsons, T.; Jenkins, S.
Factors affecting the deposition of pitch in process waters and model
dispersions. In 52nd Appita Annual General Conference Proceedings,
Brisbane, Australia, 1998; 59 – 66.
Interactions Leading to Pitch Deposition 137

11. Gutierrez, A.; del Rio, J.C.; Gonzalez-Vila, F.J.; Romero, J. Variation in
the composition of wood extractives from Eucalyptus globulus during
seasoning. J. Wood Chem. Technol. 1998, 18 (4), 439 –446.
12. Uprichard, J.M.; Lloyd, J.A. Influence of tree age on the chemical com-
position of radiata pine. NZ J. For. Sci. 1980, 10 (3), 551 –557.
13. Suckling, I.D.; Gallagher, S.S.; Ede, R.M. A new method for softwood
extractives analysis using high performance liquid chromotography.
Holzforschung 1990, 44 (5), 339 –345.
14. Ekman, R.; Holmbom, B. The chemistry of wood. In Pitch Control, Wood
Downloaded by [Georgia Tech Library] at 15:50 11 November 2014

Resin and Deresination; 1st Ed.; Back, E., Allen, L.H., Eds.; TAPPI
Press: Atlanta, 2000; 37– 76.
15. Atkins, P.W. Macromolecules and colloids. In Physical Chemistry; 6th
Ed.; Oxford University Press: Oxford, 1998; 679– 712.
16. Lide, D.R. CRC Handbook of Chemistry and Physics; 79 CRC Press:
New York, 1999.
17. Peng, G.M.; Roberts, J.C. Solubility and toxicity of resin acids. Wat. Res.
2000, 34 (10), 2779 –2785.
18. March, J. Bonding weaker than covalent. In Advanced Organic Chem-
istry,, 4th Ed.; John Wiley & Sons: New York, 1992.
19. Solomons, T.W. Physical properties and molecular structure. In Organic
Chemistry, 6th Ed.; Rose, N., Ed.; John Wiley & Sons: South Florida,
1996.
20. Jedlovszky, P.; Turi, L. A new five-site pair potential for formic acid in
liquid simulations. J. Phys. Chem. A 1997, 101, 2662– 2665.
21. Leiserowitz, L. Molecular packing modes. Carboxylic acids. Acta Cryst.
B 1976, B32, 775 –802.
22. Nakabayashi, T.; Kosugi, K.; Nishi, N. Liquid Structure of acetic acid
studied by Raman spectroscopy and ab initio molecular orbital calcu-
lations. J. Phys. Chem. A 1999, 103, 8595 –8603.
23. Beyer, T.; Price, S. Dimer or catemer? Low-energy crystal packings for
small carboxylic acids. J. Phys. Chem. B 2000, 104, 2647 –2655.
24. Green, R.D. Hydrogen Bonding by C –H Groups; Macmillan Press:
New York, 1974.
25. Pimental, G.C.; McClellan, A.L. The Hydrogen Bond; Pauling, L., Ed.;
W. H. Freeman and Company: New York, 1960.
26. Sundberg, K.; Pettersson, C.; Eckerman, C.; Holmbom, B. Preparation
and properties of a model dispersion of colloidal wood resin from
Norway spruce. J. Pulp Paper Sci. 1996, 22 (7), J248 –J252.
27. Holmbom, B. Extractives. In Analytical Methods in Wood Chemistry,
Pulping, and Papermaking, 1st Ed.; Sjöström, E., Alén, R., Eds.;
Springer: New York, 1999; Vol. 23, 125 –148.

You might also like