Вы находитесь на странице: 1из 14

Journal of Analytical and Applied Pyrolysis

60 (2001) 4154
Characterization of the water-insoluble fraction
from pyrolysis oil (pyrolytic lignin). Part I.
PYGC/MS, FTIR, and functional groups

B. Scholze
a
, D. Meier
b,
*
a
Uni6ersity of Hamburg, Institute for Physical Chemistry, Bundesstr. 45, D-20146 Hamburg, Germany
b
Federal Research Center for Forestry and Forest Products,
Institute for Wood Chemistry and Chemical Technology of Wood, Leuschnerstrasse 91,
D-21031 Hamburg, Germany
Received 25 January 2000; accepted 29 June 2000
Abstract
Pyrolytic lignins have been obtained from pyrolysis oil as ne homogeneous powder by a
novel precipitation method. The paper presents results obtained for various pyrolytic lignins
of pyrolysis oils from different fast pyrolysis processes. They were subjected to various
physico-chemical characterization methods. Analysis methods comprise chromatography,
spectroscopy and wet chemical techniques. FTIR data indicate that a changing oxygen
content mainly effects the intensity of carbonyl absorption bands. Therefore, FTIR analysis
is valuable as a fast analytical method to elucidate aging processes of pyrolysis oil. Analytical
pyrolysis combined with GC/MS, FTIR data, and results of functional group analysis show
pyrolytic lignin to be similar to technical lignins. 2001 Elsevier Science B.V. All rights
reserved.
Keywords: Pyrolytic lignin; Lignin; Fast pyrolysis; Bio oil
www.elsevier.com/locate/jaap
1. Introduction
In the past 20 years there has been an increasing number of publications on
biomass-based pyrolysis oil [13]. Considering the fact that fuel resources are

Dedicated to Professor Wolfgang Metz on the occasion of his 65th birthday


* Corresponding author. Tel.: +49-40-739517; fax: +49-40-73962503.
E-mail address: d.meier@holz.uni-hamburg.de (D. Meier).
0165-2370/01/$ - see front matter 2001 Elsevier Science B.V. All rights reserved.
PII: S0165- 2370( 00) 00110- 8
42 B. Scholze, D. Meier / J. Anal. Appl. Pyrolysis 60 (2001) 4154
limited, the combustion of pyrolysis oil in diesel engines, boilers, and gas turbines
becomes more and more attractive [4]. However, in order to establish pyrolysis oil
in the market a substantial quality improvement of the oil must be achieved. In this
respect, studies on composition and aging of pyrolysis oil are of substantial interest.
Gas chromatography (GC) is commonly used to identify and quantify volatile
substances present in organic liquids. GC analysis can only detect about 40 wt.% of
conventional pyrolysis oil. Therefore, complementary analysis techniques are
necessary.
A large number of oxygen containing reactive functional groups, e.g. carbonyl
groups, is present in pyrolysis oil. Consequently, current pyrolysis oils do not
exhibit thermal stability and cannot be effectively fractionated by standard tech-
niques such as distillation. One way to separate groups of similar compounds is
solvent fractionation. It is well known that phase separation occurs upon addition
of water to pyrolysis oil as convincingly presented recently [5]. For the determina-
tion of pyrolysis lignin contents two different methods exist: rstly the addition of
water to the oil [6] and secondly the addition of pyrolysis liquid to water [79]. The
latter seems to be suitable for analytical investigations. Adding pyrolysis oil to
water results in the precipitation of hydrophobic components mainly lignin derived
fragments, so called pyrolytic lignin. The remaining solution with the water
soluble molecules is known as liquid smoke and can be used as a source of smoke
avors and browning agents [10].
Until now there is no standardized procedure for the preparation of homoge-
neous pyrolytic lignin.
Identication of the composition and structure of pyrolytic lignin might be
essential for investigations regarding aging of pyrolysis oils. Derivatisation of
aromatic compounds with various functional groups can facilitate reactions to
larger molecules. Lignin, for instance, condenses to larger molecules upon adding
formaldehyde to its alkaline solution [11]. In addition, there is an ongoing discus-
sion about the nature of aerosol and generation of oligomers from biomass by
pyrolysis whether these compounds are formed by recombination of low molecular
weight compound or directly ejected during pyrolysis [12,13].
In this work we describe ve methods to elucidate the composition of pyrolytic
lignin. Furthermore, the analysis results of eight pyrolytic lignins originating from
different pyrolysis processes are presented and compared with some data of milled
wood lignin (MWL).
2. Experimental
2.1. Preparation and materials
The production of pyrolytic lignin was performed as follows. Sixty mililitres
pyrolysis oil was added dropwise to 1 l of ice-cooled water while stirring at 6000
rpm with an ultra turrax (oil to water ratio must be a least 1:10). A schematic
diagram of the experimental setup is depicted in Fig. 1.
43 B. Scholze, D. Meier / J. Anal. Appl. Pyrolysis 60 (2001) 4154
After slowly adding pyrolysis oil to water the precipitation of pyrolytic lignin
started. Subsequently, the pyrolytic lignin was ltered over a Bu chner funnel,
resuspended in water and slowly stirred for about 4 h in order to wash out water
soluble products. Finally, the suspension was ltered and the wet pyrolytic lignin
was dried carefully under vacuum at room temperature. Normally, the color of
pyrolytic lignin varies between light and dark brown. The origin of the eight
pyrolysis oils used for preparation of pyrolytic lignins is shown in Table 1.
Additionally, some properties of used pyrolysis oils are presented. Since reproduci-
bility of pyrolysis oils is still under investigation all pyrolytic lignins in this study
were produced only once.
2.2. Methods
2.2.1. Analytical pyrolysisgas chromatographymass spectrometry
(PyroprobeGCMS)
About 60 mg of pyrolytic lignin were pyrolyzed in a quartz boat (2.5 cm2 mm)
at 450C and a heating rate of 1000C/s using a Pyroprobe 100 (CDS) coil probe
directly coupled to a HP GC/MS series 6890. Gas chromatography was performed
using a 60 m0.25 mm capillary column (Chrompack DB 1701) with 0.25 mm lm.
Helium was used as carrier gas. The injector temperature and detector temperature
was 250 and 280C, respectively. The oven program was 4 min isothermal at 45C,
then 3C/min to 280C and nally 15 min at 280C.
2.2.2. Elemental analysis
Elemental analysis for carbon, hydrogen, and nitrogen was carried out with a
CHN-O-RAPID from HERAEUS. The oxygen content was calculated from the C,
H, and N difference as follows: O wt.%=100[C wt.%+H wt.%+N wt.%]. In
this investigation homogeneous pyrolytic lignins were employed and, therefore, a
duplicate determination for each sample was found to be sufcient.
2.2.3. Fourier transform infrared spectroscopy (FTIR-spectroscopy)
The pyrolytic lignin was investigated by FTIR-spectroscopy. About 2 mg of the
dried sample was carefully mixed with 300 mg dry KBr and pressed into a
Fig. 1. Schematic diagram of pyrolytic lignin production.
44 B. Scholze, D. Meier / J. Anal. Appl. Pyrolysis 60 (2001) 4154
T
a
b
l
e
1
O
r
i
g
i
n
,
f
e
e
d
s
t
o
c
k
,
w
a
t
e
r
c
o
n
t
e
n
t
a
n
d
e
l
e
m
e
n
t
a
l
a
n
a
l
y
s
i
s
o
f
p
y
r
o
l
y
s
i
s
l
i
q
u
i
d
s
E
l
e
m
e
n
t
a
l
a
n
a
l
y
s
i
s
a
H
i
g
h
e
r
h
e
a
t
i
n
g
W
a
t
e
r
c
o
n
t
e
n
t
F
e
e
d
s
t
o
c
k
O
r
g
a
n
i
z
a
t
i
o
n
(
%
)
v
a
l
u
e
(
M
J
/
k
g
)
O
(
%
)
b
C
(
%
)
H
(
%
)
N
(
%
)
0
.
2
2
3
8
.
6
3
2
1
.
2
5
4
.
3
5
B
e
e
c
h
6
.
8
0
F
e
d
e
r
a
l
R
e
s
e
a
r
c
h
C
e
n
t
e
r
f
o
r
F
o
r
e
s
t
r
y
a
n
d
F
o
r
e
s
t
P
r
o
d
u
c
t
s
,
3
1
.
9
I
n
s
t
i
t
u
t
e
f
o
r
W
o
o
d
C
h
e
m
i
s
t
r
y
(
I
W
C
)
,
D
E
M
i
x
e
d
5
.
9
8
0
.
3
4
3
3
.
7
9
2
2
.
8
2
6
.
4
E
N
S
Y
N
T
e
c
h
n
o
l
o
g
i
e
s
I
n
c
.
(
E
N
S
Y
N
)
,
C
A
5
9
.
8
9
H
a
r
d
w
o
o
d
3
9
.
0
7
U
n
i
o
n
E
l
e
c
t
r
i
c
a
F
E
N
O
S
A
(
F
E
N
O
S
A
)
I
,
E
S
2
0
.
2
E
u
c
a
l
y
p
t
u
s
2
5
.
6
5
4
.
4
4
6
.
1
0
0
.
3
9
3
9
.
4
1
2
0
.
3
0
.
6
4
5
3
.
5
3
3
6
.
0
6
.
4
2
U
n
i
o
n
E
l
e
c
t
r
i
c
a
F
E
N
O
S
A
(
F
E
N
O
S
A
)
I
I
,
E
S
W
h
e
a
t
/
S
t
r
a
w
0
.
3
5
P
o
p
l
a
r
3
9
.
6
0
2
0
.
3
1
6
.
1
5
3
.
5
5
6
.
5
0
A
s
t
o
n
U
n
i
v
e
r
s
i
t
y
(
A
S
T
O
N
)
,
U
K
6
.
6
4
0
.
3
0
3
6
.
8
6
2
1
.
9
N
a
t
i
o
n
a
l
R
e
n
e
w
a
b
l
e
E
n
e
r
g
y
L
a
b
o
r
a
t
o
r
y
(
N
R
E
L
)
c
,
U
S
P
o
p
l
a
r
1
6
.
3
5
6
.
2
0
6
.
5
1
0
.
3
9
3
8
.
2
6
2
1
.
0
1
8
.
6
T
w
e
n
t
e
U
n
i
v
e
r
s
i
t
y
(
B
T
G
)
,
N
L
5
4
.
8
4
M
i
x
e
d
S
o
f
t
w
o
o
d
3
8
.
0
6
2
1
.
3
P
i
n
e
0
.
1
1
5
5
.
2
3
6
.
6
0
1
7
.
4
T
e
c
h
n
i
c
a
l
R
e
s
e
a
r
c
h
C
e
n
t
e
r
o
f
F
i
n
l
a
n
d
(
V
T
T
)
,
F
I
a
B
a
s
e
d
o
n
w
a
t
e
r
f
r
e
e
o
i
l
.
b
C
a
l
c
u
l
a
t
e
d
b
y
d
i
f
f
e
r
e
n
c
e
.
c
H
o
t

l
t
e
r
e
d
o
i
l
.
45 B. Scholze, D. Meier / J. Anal. Appl. Pyrolysis 60 (2001) 4154
self-supporting pellet. FTIR measurements were performed using a BIORAD FTIR
model FTS-40. For each sample 32 spectra were accumulated between 4000 and 400
cm
1
and averaged. The resulting spectra were normalized to the highest peak in
the ngerprint region between 2000 and 700 cm
1
.
2.2.4. Methoxyl groups
Methoxyl groups were determined according to the method by Viebo ck and
Schwappach [14]. For each methoxyl group three Iodine molecules, I
2
, which are
determined by titration with thiosulfate, are formed, which make this method very
accurate and suitable for the work described here.
2.2.5. Hydroxyl groups
Hydroxyl groups were measured according to the method by Verley and Bo lsing
[15]. Fifty miligrams of pyrolytic lignin were acetylated with a solution of 0.5 ml
pyridine and 0.5 ml acetanhydride at 50C. The obtained liquid was mixed with 15
ml acetone and 1 ml water, and the solution was titrated with sodium hydroxide.
2.2.6. Heating 6alue
In general heating values can be reported as higher and lower heating value,
HHV (gross) and LHV (net), respectively. The difference between HHV and LHV
is equal to the heat of vaporization of water formed by combustion of the fuel. The
heating values in this study were obtained from calculation by Dulongs formula
[16]:
HHV(MJ/kg) =

338.2%C+1442.8

%H
%O
8
n
0.001
3. Results and Discussion
Pyrolytic lignins were obtained by adding pyrolysis oil dropwise to water under
vigorous stirring. It has been noticed during the experiments described here that at
temperatures above 20C the pyrolysis oil droplets could not be dissolved immedi-
ately, and that they tend to conglomerate and form a sticky mass. Consequently,
components may not get dissolved completely and the yield of pyrolytic lignin
seems to increase. One possible heat source is the turrax, which heats up at high
rate of rotation and, slowly increases the temperature in the solution. For this
reason the water must be cooled during the entire experiment.
Another important parameter is the stir velocity. Initial experiments showed that
pyrolytic lignin agglomerates using a magnetic stirrer and therefore a turrax was
utilized. It was found that only minor amounts of pyrolytic lignins, e.g. from
FENOSA eucalyptus pyrolysis oil, could be precipitated at a slow rotation rate.
Therefore, the rate of rotation in this study was increased to 6000 rpm. Neverthe-
less, it should be mentioned that we were also able to obtain a ne pyrolytic lignin
powder when pyrolysis oil was shaken for some hours with a large amount of water
46 B. Scholze, D. Meier / J. Anal. Appl. Pyrolysis 60 (2001) 4154
Table 2
Yield, color, and elemental analysis data of pyrolytic lignins and MWLs
Elemental analysis Color Yield
a
(%) Origin of oil
C (%) H (%) N (%) O (%)
b
66.91 6.19 13.5 0.27 Reddish brown 26.63 IWC
66.18 ENSYN 6.02 24.1 0.23 27.57 Light brown
63.68 5.79 0.19 Brown 30.34 FENOSA I 25.0
Light brown 16.5 65.22 6.13 1.02 27.63 FENOSA II
Light brown 24.0 66.03 6.11 0.21 27.65 ASTON
66.60 6.24 0.34 Brown 26.82 NREL 23.0
67.43 BTG 6.19 27.7 0.24 26.14 Light brown
70.56 6.64 0.15 Dark brown 22.65 19.0 VTT
58.84 5.92 0 35.24 MWL (beech)
60.20 6.21 0 33.59 MWL (eucalyptus)
58.92 5.73 0 35.35 MWL (poplar)
63.46 5.74 0 MWL (pine) 30.80
a
Based on water free oil.
b
Calculated by difference.
(oil water ratio at least 1:10). However, differences in the composition of pyrolytic
lignins obtained by various methods have to be expected. This will be investigated
in an upcoming study.
Pyrolytic lignins were dried at about 25C. Temperatures above 40C cause
reactions and as a result the surface of pyrolytic lignin turns black. Normally, the
color of pyrolytic lignin varies between light and dark brown. Pyrolytic lignin has
a low density and electrostatic charging of the ne powder is sometimes observed.
Precipitation yields and colors on visual inspection of the obtained pyrolytic
lignins are given in Table 2. It can be seen that the yield of pyrolytic lignin as
determined for the different pyrolysis oils varies from 13.5 to 27.7% based on water
free pyrolysis oils. The lowest yield of pyrolytic lignin in this study was obtained
from IWC pyrolysis oil (see Table 1). The high water content is not a sufcient
explanation since VTT pyrolysis oil also yielded an amount of pyrolytic lignin
below 20% based on water free oil. Piskorz et al. [17] reported already that
pyrolytic lignin yields can vary with the nature of the biomass. The authors found
yields of pyrolytic lignin ranging from 16 to 22%.
As shown in Table 2 colors of pyrolytic lignins vary between light and dark
brown. ASTON and NREL pyrolysis oils are both obtained from poplar by
ablative processes. However, different reactor congurations were utilized. ASTON
pyrolysis oil was produced with a 5 kg/h ablative reactor, whereas NREL pyrolysis
oil was made using a vortex reactor and the vapor was hot ltered before oil
condensation. In addition the two pyrolysis oils exhibited different degrees of aging.
Interestingly, different colors for pyrolytic lignins obtained from ASTON and
NREL oils were obtained, possibly resulting from the age of the oils or differing
production parameters.
47 B. Scholze, D. Meier / J. Anal. Appl. Pyrolysis 60 (2001) 4154
Mean values for elemental analysis of pyrolytic lignins are given in Table 2.
Additionally, elemental analysis data of milled wood lignins are presented [18].
Pyrolytic lignin contains more carbon than wood (pyrolytic lignin: about 66.5%,
wood: about 49%). Since hydrogen content in pyrolytic lignin and wood is similar,
pyrolytic lignin has less oxygen than wood (pyrolytic lignin: about 26.9%, wood:
about 45%). Compared to milled wood lignins the elemental composition of
pyrolytic lignin is not similar but in the same range. The carbon content of
pyrolytic lignin is between 3 and 8% higher than that one of MWL produced from
the corresponding wood. Elemental analysis data resulted in differences of 39%
and 0.30.9% for oxygen content and hydrogen content, respectively.
Composition of pyrolytic lignin was investigated by analytical pyrolysisGC/MS.
Peak identication was guided by a special library, which includes those specic
compounds that can be detected in pyrolysis oils. Fig. 2 presents the pyrogram of
pyrolytic lignin from IWC beech oil compared to MWL from beech wood. The
pyrogram of pine MWL is depicted together with VTT pine pyrolytic lignin in Fig.
3. Peak identication for all four pyrograms is given in Table 3. Since the use of the
internal standard method for peak quantication of compounds determined with a
Fig. 2. Pyrogram of pyrolytic lignin IWC and MWL (beech) mirrored at baseline. Peak identication see
Table 3.
Fig. 3. Pyrogram of pyrolytic lignin VTT and MWL (pine) mirrored at baseline. Peak identication see
Table 3.
48 B. Scholze, D. Meier / J. Anal. Appl. Pyrolysis 60 (2001) 4154
T
a
b
l
e
3
P
e
a
k
i
d
e
n
t
i

c
a
t
i
o
n
f
o
r
p
y
r
o
g
r
a
m

G
C
/
M
S
q
u
a
l
i

c
a
t
i
o
n
a
I
n
t
e
n
s
i
t
y
o
f
M
W
L
P
e
a
k
n
u
m
b
e
r
I
n
t
e
n
s
i
t
y
o
f
p
y
r
o
l
y
t
i
c
I
n
t
e
n
s
i
t
y
o
f
p
y
r
o
l
y
t
i
c
I
n
t
e
n
s
i
t
y
o
f
M
W
L
C
o
m
p
o
u
n
d
b
e
e
c
h
p
e
a
k
s
p
i
n
e
p
e
a
k
s
l
i
g
n
i
n
p
i
n
e
p
e
a
k
s
l
i
g
n
i
n
b
e
e
c
h
p
e
a
k
s
1
+
+
+
+
+
+
+
+
G
u
a
i
a
c
o
l
o
-
C
r
e
s
o
l

3
-
M
e
t
h
y
l
g
u
a
i
a
c
o
l
3
+
+

4
-
M
e
t
h
y
l
g
u
a
i
a
c
o
l
+
+
+
+
+
+
+
+
+
4
2
,
4
-
D
i
m
e
t
h
y
l
p
h
e
n
o
l

3
-
E
t
h
y
l
g
u
a
i
a
c
o
l
+
6

4
-
E
t
h
y
l
g
u
a
i
a
c
o
l

+
+
+
7
+
V
i
n
y
l
g
u
a
i
a
c
o
l
8
+
+
+
+
+
E
u
g
e
n
o
l

+
+
+
9
+
4
-
P
r
o
p
y
l
g
u
a
i
a
c
o
l
+
1
0

S
y
r
i
n
g
o
l
+
+
1
1
+
+

+
c
i
s
-
I
s
o
e
u
g
e
n
o
l

1
2
+

+
+
t
r
a
n
s
-
I
s
o
e
u
g
e
n
o
l
+
1
3
+
+
+
+
+
+
4
-
M
e
t
h
y
l
s
y
r
i
n
g
o
l
1
4
+
+
+

V
a
n
i
l
l
i
n
+
+
+
+
+
+
+
1
5

H
o
m
o
v
a
n
i
l
l
i
n

1
6

+
+
+
1
7

4
-
E
t
h
y
l
s
y
r
i
n
g
o
l
+
+
A
c
e
t
o
g
u
a
i
a
c
o
n
e

1
8
+
4
-
V
i
n
y
l
s
y
r
i
n
g
o
l
+
+
+
+

1
9

G
u
a
i
a
c
y
l
a
c
e
t
o
n

2
0

4
-
A
l
l
y
l
a
n
d
4
-
P
r
o
p
y
l
s
y
r
i
n
g
o
l
+
2
1
+

P
r
o
p
i
o
g
u
a
i
a
c
o
n
e
+
2
2

+
c
i
s
-
4
-
P
r
o
p
e
n
y
l
s
y
r
i
n
g
o
l
+

2
3

2
4
6
-
H
y
d
r
o
x
y
-
5
,
7
-
d
i
m
e
t
h
o
x
y
-
i
n
d
e
n
e
+

+
t
r
a
n
s
-
4
-
P
r
o
p
e
n
y
l
s
y
r
i
n
g
o
l
+
+
+
+
+
+

2
5
D
i
h
y
d
r
o
c
o
n
i
f
e
r
y
l
a
l
c
o
h
o
l

2
6

S
y
r
i
n
g
a
l
d
e
h
y
d
e

2
7
+
+
+
+

A
c
e
t
o
s
y
r
i
n
g
o
n
e
+
2
8

+
+
2
9
t
r
a
n
s
-
C
o
n
i
f
e
r
y
l
a
l
c
o
h
o
l

+
+
+
+

C
o
n
i
f
e
r
y
l
a
l
d
e
h
y
d
e
+

+
+
3
0
S
y
r
i
n
g
y
l
a
c
e
t
o
n
e
+
+

3
1

P
r
o
p
i
o
s
y
r
i
n
g
o
n
e

3
2
+
+
3
3

D
i
h
y
d
r
o
s
i
n
a
p
y
l
a
l
c
o
h
o
l

c
i
s
-
S
i
n
a
p
y
l
a
l
c
o
h
o
l
3
4

t
r
a
n
s
-
S
i
n
a
p
y
l
a
l
c
o
h
o
l
+
+

3
5

S
i
n
a
p
a
l
d
e
h
y
d
e

3
6

+
a
+
w
e
a
k
+
+
m
i
d
d
l
e
+
+
+
s
t
r
o
n
g

a
b
s
e
n
t
o
r
t
r
a
c
e
s
.
49 B. Scholze, D. Meier / J. Anal. Appl. Pyrolysis 60 (2001) 4154
Fig. 4. FTIR spectra of the eight different pyrolytic lignins in the ngerprint region between 2000 cm
1
and 700 cm
1
.
PyroprobeGC/MS is still under investigation, quantitative comparison of different
pyrograms is not yet possible. Therefore, only peak intensity ratios obtained for
each pyrogram measured can be compared.
One noticeable difference between the pyrograms are the low yields of vanillin
and syringaldehyde in pyrolytic lignins which could be explained by the absence of
hydroxyl groups or ethers in alpha position of the propyl sidechain. Another
remarkable difference is the very low amount, even absence, of trans-coniferylalco-
hol and coniferylaldehyde as well as trans-sinapalcohol and sinapaldehyde in
pyrolytic lignins. Normally, these are typical degradation products of lignin with an
intact sidechain. Therefore, it may be supposed that the propyl sidechain has been
largely destroyed during the fast pyrolysis process.
The pyrolytic lignin samples were measured as self-supported pellets by FTIR
spectroscopy. FTIR spectra for the eight different pyrolytic lignins are presented in
Fig. 4.
Since absorption features from different components superimpose in the nger-
print region a detailed interpretation of absorption bands is rendered difcult.
Absorption bands above 1510 cm
1
result from carbonyl and aromatic skeletal
vibrations. Below 1500 cm
1
all bands are complex and have their origin in a
variety of vibrational modes. Since bands above 1600 cm
1
, associated to conju-
gated CO groups, are often broadened by aromatic skeletal vibrations, experimen-
tal spectra were tted with gaussian prole functions in this region using a
conventional least squares algorithm (software package used: WinXAS by Ressler
[19]). As a typical example t results for ASTON and NREL pyrolytic lignin are
depicted in Fig. 5 and Fig. 6, respectively. Numbers for obtained absorption bands
are given in Table 4.
50 B. Scholze, D. Meier / J. Anal. Appl. Pyrolysis 60 (2001) 4154
FTIR data for different MWL samples were published by Faix [20]. The author
described the origins for different bands and gave the corresponding wavenumbers.
With respect to these wavenumbers FTIR data in the region between 2000 and 700
cm
1
wavenumbers of all pyrolytic lignin described in this work are given in Table
4. The absorption intensities are normalized with respect to the intensity of the
highest band in this region.
In order to nd any correlation between absorption bands and other pyrolytic
lignin parameters the experimental absorption was depicted versus elemental analy-
sis data and heating values. For all lignins with carbon contents between 67.5 and
65.2% an interesting, even if weak correlation was observed in the absorption bands
at 1701 cm
1
, 1652 cm
1
, and 1600 cm
1
. Analysis data and resulting regression
line for the absorption band at 1701 cm
1
are presented in Fig. 7. The three
investigated absorption bands originate from carbonyl groups. The data show that
an increase in carbon content results in a decrease of the corresponding absorption
bands. In addition, an increase of oxygen content results in an increase of the
absorption bands. Therefore, it can be concluded that alteration in oxygen content
mainly effects the intensity of carbonyl absorption bands. Other oxygen-related
Fig. 5. FTIR spectra of ASTON pyrolytic lignin (dot) tted with gaussian prole functions (single
functions: dashdot; resulting function: connect).
Fig. 6. FTIR spectra of NREL pyrolytic lignin (dot) tted with gaussian prole functions (single
functions: dashdot; resulting function: connect).
51 B. Scholze, D. Meier / J. Anal. Appl. Pyrolysis 60 (2001) 4154
T
a
b
l
e
4
F
T
I
R
d
a
t
a
o
f
p
y
r
o
l
y
t
i
c
l
i
g
n
i
n
s
E
N
S
Y
N
%
M
W
L
b
e
e
c
h
F
E
N
O
S
A
e
u
c
.
N
R
E
L
%
A
b
s
.
F
E
N
O
S
A
A
S
T
O
N
%
B
a
n
d
o
r
i
g
i
n
B
T
G
%
A
b
s
.
V
T
T
%
A
b
s
.
W
a
v
e
n
u
m
b
e
r
s
I
W
C
%
A
b
s
.
M
W
L
p
i
n
e
%
s
t
r
a
w
%
A
b
s
.
%
A
b
s
.
%
A
b
s
.
A
b
s
.
(
c
m

1
)
A
b
s
.
A
b
s
.
3
9
5
7
4
6
6
9
5
2
5
2
5
3
C
=
O
s
t
r
e
t
c
h
8
2
1
7
0
1

1
7
3
4
a
3
5
3
6
i
n
u
n
c
o
n
j
.
k
e
t
o
n
e
s
,
c
a
r
b
o
n
y
l
a
n
d
e
s
t
e
r
g
r
o
u
p
s
2
1
4
5
1
2
7
8
6
8
5
9
5
5
1
6
5
2

1
6
6
6
a
C
=
O
s
t
r
e
t
c
h
i
n
c
o
n
j
.
p
-
s
u
b
s
.
a
r
y
l
k
e
t
o
n
e
s
7
4
4
8
6
8
5
8
6
1
4
6
6
0
4
0
1
5
9
3

1
6
0
9
a
4
3
4
0
a
r
o
m
a
t
i
c
s
k
e
l
e
t
a
l
v
i
b
.
p
l
u
s
C
=
O
s
t
r
e
t
c
h
8
4
5
8
1
0
0
7
7
9
4
4
7
9
5
9
3
9
8
1
5
1
0

1
5
1
5
8
5
a
r
o
m
a
t
i
c
s
k
e
l
e
t
a
l
v
i
b
.
9
0
1
4
4
3

1
4
6
3
6
8
C

H
7
8
7
6
7
9
6
1
6
0
6
5
6
9
7
4
d
e
f
o
r
m
a
t
i
o
n
s
,
a
s
y
m
.
i
n

C
H
3
a
n
d

C
H
2

5
3
6
1
4
7
6
4
5
6
A
r
o
m
a
t
i
c
5
6
1
4
2
2

1
4
3
1
5
1
5
1
5
4
5
5
s
k
e
l
e
t
a
l
v
i
b
.
C
o
m
b
i
n
e
d
w
i
t
h
C

H
i
n

p
l
a
n
e
d
e
f
o
r
m
a
t
i
o
n
9
3
9
9
8
5
9
4
9
4
9
7
6
6
6
9
8
0
8
0
C

C
p
l
u
s
C

O
1
2
1
4

1
2
3
3
p
l
u
s
C
=
O
s
t
r
e
t
c
h
5
3
A
r
o
m
.
C

H
4
1
4
9
4
9
5
8
5
5
4
5
1
0
3
3

1
0
3
7
5
0
7
8
6
3
i
n
-
p
l
a
n
e
d
e
f
.
,
p
l
u
s
C

O
d
e
f
.
i
n
p
r
i
m
.
A
l
c
o
h
o
l
s
,
p
l
u
s
C
=
O
s
t
r
e
t
c
h
(
u
n
c
o
n
j
.
)
1
6
1
3
1
6
9
1
4

9
1
9
1
3
C

H
7
n
o
n
o
n
o
1
7
1
1
o
u
t
-
o
f
-
p
l
a
n
e
,
a
r
o
m
.
C

H
8
5
2

8
5
9
1
6
1
0
1
1
1
3
1
3
1
2
7
1
6
1
5
1
6
o
u
t
-
o
f
-
p
l
a
n
e
i
n
P
o
s
.
2
,
5
a
n
d
6
o
f
G
u
a
i
a
c
o
l
u
n
i
t
s
a
A
f
t
e
r

t
t
i
n
g
.
52 B. Scholze, D. Meier / J. Anal. Appl. Pyrolysis 60 (2001) 4154
Fig. 7. FTIR absorption at 17011712 cm
1
versus carbon content.
Table 5
Methoxyl groups, hydroxyl groups, and higher heating values of pyrolytic lignins
Origin of oil for pyrolytic lignin OMe
i
OH (wt.%) OH
i
HHV (MJ/kg) OMe (wt.%)
10.7 0.88 15.8 26.8 0.77 IWC
8.5 0.41 13.6 1.12 26.1 ENSYN
12.0 0.99 0.54 24.4 11.0 FENOSA I
11.9 0.81 FENOSA II 25.9 7.5 0.47
11.4 0.94 0.51 26.2 ASTON 10.4
11.3 0.93 26.7 NREL 12.6 0.61
11.9 0.81 0.43 27.0 6.9 BTG
0.48 7.7 12.0 0.82 29.4 VTT
19.8 12.0 22.1 MWL (beech)
12.2 23.3 MWL (eucalyptus) 21.7
12.2 MWL (poplar) 20.0 21.8
14.7 24.2 MWL (pine) 15.9
bands, such as hydroxyl and methoxyl bands, are less affected. If these results could
be veried also for whole pyrolysis oils, FTIR analysis might be used as a fast
technique to predict the carbonyl group content. Furthermore, these investigations
may reveal results connected with the aging of pyrolysis oils.
The results from methoxyl group determination of pyrolytic lignins are given in
Table 5. For methoxyl groups an index is calculated from the ratio of methoxyl
groups of pyrolytic lignins versus methoxyl groups of milled wood lignins. The
methoxyl group index (OMe
i
) can be calculated by (Eq. (1)). The mean values of
methoxyl groups of hardwood MWLs and that of pine MWL were taken as indices
of hardwood and softwood pyrolytic lignins, respectively.
OMe
i
=
%OMePyLig
%OMeMWL
(1)
In comparison to MWL the content of methoxyl groups of pyrolytic lignin is
reduced by pyrolysis. Methoxyl groups of MWL from hardwood and softwood are
53 B. Scholze, D. Meier / J. Anal. Appl. Pyrolysis 60 (2001) 4154
given by Fengel and Wegner [21] as 1822% and 1216%, respectively.
Demethoxylation may be caused by conversion of methoxyl groups to methanol
during thermal degradation. Except for the ENSYN sample, methoxyl indices of
hardwood pyrolytic lignins have larger numbers than those of softwood samples
indicating less demethoxylation reactions.
The content of hydroxyl groups (10.7%-13.6%) in pyrolytic lignin is similar to
that of MWLs (12.014.7%). According to the procedure for calculating the
methoxyl index, an index for hydroxyl groups is dened. The hydroxyl group index
(OH
i
) can be reduced to (Eq. (2)):
OH
i
=
%OHPyLig
%OHMWL
(2)
Hydroxyl group data as well as postulated indices are shown in Table 5. As
opposed to MWL pyrolytic lignin samples contain less amounts of hydroxyl
groups. The splitting of propyl side chain could be an explanation.
Heating values of pyrolytic lignins obtained in the range between 24.4 and 29.4
MJ/kg were calculated from elemental analysis and presented in Table 5. As
expected from the smaller amount of oxygen in the corresponding samples, heating
values decrease with the following order: pyrolytic lignin\MWL\pyrolysis oil.
4. Conclusions
As indicated by analytical pyrolysis, the basic units of pyrolytic lignins have
similarities with MWL. Splitting of methoxyl groups and cutting of the propyl
sidechain was observed by PyGC/MS. Our work has indicated that oligomeric
molecules in pyrolysis oils are desorbed from the feedstock by thermal ejection
because recombination of vaporized molecules to such pyrolytic lignin structure is
unlikely to occur.
Two alternative methods for preparation of pyrolytic lignin exist. The precipita-
tion method presented in this paper gives, in contrast to the method described by
Czernik and Piskorz [6], a ne powder without inclusions of water soluble com-
pounds. For investigations and applications a homogeneous sample is to be
preferred.
FTIR results showed a correlation between carbonyl absorption bands and
oxygen content as well as carbon content. These results are most promising for fast
analysis of pyrolytic lignins and pyrolysis oils.
Acknowledgements
Financial support from European Union through the Contract JOR3-CT95-0025
Bio fuel oil for power plants and boilers. Special thanks are due to I. Fortmann, G.
Jessusek, and S. Radtke for skilful technical assistance. V. Daiker, K. Ohltmann,
M. Thormahlen are acknowledged for technical assistance. T. Ollesch is acknowl-
54 B. Scholze, D. Meier / J. Anal. Appl. Pyrolysis 60 (2001) 4154
edged for providing IWC pyrolysis oil. The authors are grateful to VTT, Aston
University, Twente University, ENSYN Technologies, NREL and Union FENOSA
for providing the pyrolysis oil samples.
References
[1] A.V. Bridgwater, S. Czernik, D. Meier, I. Piskorz, in: R.P. Overend, E. Chornet (Eds.), Biomass.
A Growth Opportunity in Green Energy and Value-Added Products, vol. 2, Elsevier, Amsterdam,
1999, p. 1217.
[2] A. Oasmaa, S. Czernik, in: R.P. Overend, E. Chornet (Eds.), Biomass. A Growth Opportunity in
Green Energy and Value-Added Products, vol. 2, Elsevier, Amsterdam, 1999, p. 1247.
[3] S. Czernik, R. Maggi, G.V.C. Peacocke, in: R.P. Overend, E. Chornet (Eds.), Biomass. A Growth
Opportunity in Green Energy and Value-Added Products, vol. 2, Elsevier, Amsterdam, 1999, p.
1235.
[4] J.P. Diebold, A. V. Bridgwater, in: A. Bridgwater, S. Czernik, I. Diebold, D. Meier, A. Oasmaa, C.
Peacocke, J. Piskorz, D. Radlein (Eds.), Fast Pyrolysis of Biomass: A Handbook, CPL Press,
Bio-Energy Research Group, Newbury, UK, 1999.
[5] D. Radlein, Fast Pyrolysis of Biomass: A Handbook, CPL Press, Bio Energy Research Group,
Newbury, UK, 1999.
[6] D. Meier, Fast Pyrolysis of Biomass: A Handbook, CPL Press, Bio-Energy Research Group,
Newbury, UK, 1999, pp. 94101.
[7] D. Meier, B. Scholze, in: M. Kaltschmitt, A.V. Bridgwater (Eds.), Biomass Gasication and
Pyrolysis, CPL Press, Newbury, UK, 1997, p. 431.
[8] A. Oasmaa, E. Lepparnaki, P. Koponen, I Levander and E. Tapola, Physical characterisation off
biomass-based pyrolysis liquids. Application of standard fuel oil analyses. VTT Publications 3 06.
VTT Energy, Espoo (1997).
[9] K. Sipild, E. Kuoppala, L. Fagernas, A. Oasmaa, Biomass Bioenergy 14 (2) (1998) 103113.
[10] D. Radlein, Fast Pyrolysis of Biomass: A Handbook, CPL Press, BioEnergy Research Group,
Newbury, UK, 1999, p. 177.
[11] Panagiotis N. PyNe Newsletter, 6, Aston University Birmigham B4 7ET (1998) p. 6.
[12] J. Lede, J.P. Diebold, G.V.C. Peacocke, I. Piskorz, Fast Pyrolysis of Biomass: A Handbook, CPL
Press, Bio-Energy Research Group, Newbury, K, 1999, pp. 5165.
[13] J. Piskorz, P. Majerski, D. Radlein, in: R.P. Overend, E. Chornet (Eds.), Biomass. A Growth
Opportunity in Green Energy and Value-Added Products, vol. 2, Elsevier, Amsterdam, 1999, pp.
11531159.
[14] F. Viebbck, A. Schwappach, in: E. Mifer (Ed.), Houben-Weyl, Methoden der organischen Chemie,
Band 11 analytische Methoden, 4. Auage, Thieme, Berlin, 1953, pp. 406408.
[15] A. Verley, F. Bolsing, Chem. Ber. 34 (1901) 3354.
[16] C.D. Ro mpp, Chernie Lexikon, Version 1.0, Thieme, Berlin, 1995.
[17] J. Piskorz, P. Majerski, D. Radlein, D.S. Scott, Energy Fuel 3 (1989) 723726.
[18] O. Faix, personal communication (1999).
[19] T. Ressler, WinXAS: A XAS Data Analysis Program under NIS Windows, J. Synch. Rad. 5 (1998)
118.
[20] O. Faix, in: S.Y. Lin, C.W. Dence (Eds.), Methods in Lignin Chemistry, Springer, Berlin, 1992, p.
83.
[21] D. Fengel, G. Wegner, Wood. Chemistry, Ultrastructure, Reactions, de Gruyter, Berlin, 1984, p.
153.
.

Вам также может понравиться