Sunteți pe pagina 1din 16

Fuel Processing Technology, 31 (1992) 241-256 241

Elsevier Science Publishers B.V., Amsterdam


Short Communication
Characteri zati on and stabi l i ty anal ysi s of wood-deri ved
bi o-oi l
John D. Adjaye, Ramesh K. Sharma and Narendra N. Bakhshi
Catalysis and Chemical Reaction Engineering Laboratory, Department of Chemical
Engineering, University of Saskatchewan, Saskatoon, S7N 0 WO (Canada)
(Received August 28th, 1991; accepted in revised form April 5th, 1992 )
Abs t r ac t
The stability characteristics of a bio-oil, produced by the high pressure liquefaction of aspen
wood were studied by observing the changes in its physical properties, composition and distillation
characteristics with time. Distillation characteristics of the fresh bio-oil showed that maximum
amount of organic distillate was obtained at 172 Pa and 200C. This distillate fraction mainly
consisted of aromatic, aliphatic and naphthenic hydrocarbons and oxygenated compounds such
as phenols, furans, alcohols, acids, ethers, aldehydes and ketones. The bio-oil viscosity, and chem-
ical composition were found to change substantially over time probably due to polymerization of
some components. Upon storage, the concentration of aromatic hydrocarbons and phenols de-
creased while the concentration of aldehydes and ketones increased. Also, the oxygen content of
the distillate decreased from 22.7 wt% for the fresh bio-oil to 18.8 wt% after 31 days. However,
when the bio-oil was mixed with tetralin it was observed that the properties of the mixture re-
mained unchanged with time. Tetralin was found to donate hydrogen leading to the improvement
in bio-oil stability. A free radical mechanism is proposed to explain the effect of tetralin.
INTRODUCTION
The diminishing supplies of low cost fossil fuels and chemical feedstocks
have impelled researchers and industry to pursue research into alternative re-
newable resources. One alternative is the conversion of biomass to liquid fuels
and chemicals. In this area, one route has been the catalytic conversion of
vegetable oils such as castor, canola, palm, flax etc. to fuels and chemicals [ 1 ].
Another route which also has been followed is the conversion of wood to a bio-
oil by pyrolytic or catalytic liquefaction [2-4]. The bio-oil was then catalyti-
cally upgraded to fuels and chemicals. Viscosities of the wood-derived bio-oils
range from moderate to very high so t hat their direct pumping often lead to
Correspondence to: Dr. N.N. Bakhshi, Catalysis and Chemical Reaction Engineering Laboratory,
Department of Chemical Engineering, University of Saskatchewan, Saskatoon, S7N 0W0, Canada.
0378-3820/92/$05.00 1992 Elsevier Science Publishers B.V. All rights reserved.
242 J.D. Adjaye et al. / Fuel Processing Technol. 31 (1992) 241-256
pipe blockage. These bio-oils also are corrosive [5,6] and have high oxygen
contents. In addition, they are physically and chemically unstable with time
i.e. there is a continuous variation in their density, viscosity and chemical com-
position with time. The changes in the composition are accompanied by a con-
tinuous polymerization even at ambient temperature [7,8]. Therefore, these
bio-oils need to be stabilized before being catalytically upgraded.
Direct upgrading of wood-derived oils over HZSM-5 catalyst has been re-
ported to lead to severe coking due to poor stability and low H/ C ratios of the
bio-oils [9]. Sharma and Bakhshi [10] used tetralin, methanol and steam as
co-feeds during the upgrading of a wood-derived bio-oil in order to improve its
stability and minimize the coking. The best results were achieved with tetralin
which was found to donate hydrogen during upgrading [11 ]. Hydrogen-donor
solvents such as cyclohexane, decalin or tetralin have been used in coal lique-
faction also. The main advantages of using such solvents stem from an im-
proved pumping characteristics of the feed mixtures and much lower operating
pressures that can be employed during upgrading [7,12-17]. The bio-oils with
improved stability also lead to a higher yield of the upgraded product and a low
char formation [11-13 ].
There is no detailed information in literature on the stability aspects of the
bio-oils or the effect of hydrogen donor solvents on the stability. In order to
understand these aspects a rheological and compositional study of the bio-oil
with time is necessary both in the absence and presence of tetralin. The objec-
tive of this work was to investigate the changes in physical properties, chemical
composition and distillation characteristics of a wood-derived bio-oil with
storage time. Since tetralin was to be used as co-feed during upgrading of the
bio-oil [ 11 ], the effect of tetralin addition on stability also was investigated by
observing the changes in physical properties and chemical composition of bio-
oil-tetralin mixture with time.
EXPERI MENTAL
Bio-oil preparation
The bio-oil was prepared by high pressure liquefaction of air-dried aspen
poplar wood powder following the procedure described by Eager et al. [3].
Hundred grams of the wood powder (about 1 mm particle size ) was mixed with
10 g of sodium carbonate and 500 g of distilled water in an autoclave (Aminco
High Pressure Autoclave, model 406-010A) and pressurised to about 5 MPa
with carbon monoxide. The contents were heated and maintained at 360 C for
2 h before being cooled. The product gases were vented off. The liquid product
consisted of organic and aqueous layers. The organic layer consisting of bio-
oil was separated from the aqueous phase by decanting. The yield of the bio-
oil was 30 g. Water analysis by Karl Fischer titration for the fresh bio-oil showed
J.D. Adjaye et al. / Fuel Processing Technol. 31 (1992) 241-256 243
t hat it cont ai ned 1.5 wt% water. El ement al analysis showed t hat carbon, hy-
drogen and oxygen (det ermi ned by difference) cont ent s of t he bio-oil were
70.9, 8.8 and 20.3 wt%, respectively. No nitrogen could be det ect ed in t he CHN
analyses. Several bat ches of bio-oil were produced for charact eri zat i on and
stability studies. For stability measurement s, four of t hese were stored in sep-
arate 30 ml bottles, filled to t he t op and stoppered. The bot t l es were covered
with al umi ni um foil to prevent exposure to light. Two of t hese cont ai ned bio-
oil plus t et ral i n mixtures in a 2: 1 weight ratio and t he ot her two cont ai ned
only bio-oil. Various charact eri zat i on measurement s were made in duplicate
from t hese bot t l es and t he average values determined. For each analysis which
compri sed of density, viscosity and vacuum distillation measurement s, about
15 mL of t he bio-oil was removed. Once t he measurement s were completed,
only 9 mL could be returned. The bot t l e was kept st oppered all t he time. Thus,
t he bio-oil was exposed to a very small amount of air during storage.
Di s t i l l at i on charact eri st i cs
The distillation charact eri st i cs of t he fresh bio-oil were used as basis for
stability monitoring. About 4 g of t he fresh bio-oil was distilled over t he tem-
perat ure range 85-250 C in a Buchi GRK-50 distillation uni t for 30 min under
a fixed vacuum of 172 Pa. The distillate was collected in a bulb i mmersed in
liquid nitrogen and was analyzed by gas chromat ography. The amount of non-
volatile residue remai ni ng after t he distillation was determined.
St abi l i t y anal ysi s
The stability of t he bio-oil was characterized by following t he changes in
density, viscosity, distillation characteristics and chemical composi t i on of t he
fresh bio-oil after 16 and 31 days. Similar measurement s were made for bio-
oi l -t et ral i n feed mixture. In addition, el ement al analyses of t he distillate and
residue fractions obt ai ned from bio-oil and bi o-oi l -t et ral i n feed mixture, were
performed on days 1, 16 and 31. No wat er analysis were carried out on t he bio-
oil during or after storage.
For stability monitoring, about 4 g of fresh bio-oil was distilled under a vac-
uum of 172 Pa at 175, 200 and 250 C. The density, viscosity, distillation char-
acteristics and t he distillate composi t i on were measured on days 1, 16 and 31.
To st udy t he effect of t et ral i n addition on stability, tetralin and bio-oil were
mixed in 1 : 2 weight ratio. Sampl es of t he mi xt ure were distilled under vacuum
at 175, 200 and 250C on days 1, 16 and 31 and t he amount s of residue and
distillate and t he composi t i on of distillate were determined. The viscosity and
densi t y measurement s also were made at t he above time intervals.
244 J.D. Adjaye et al. /Fuel Processing Technol. 31 (1992)241-256
Analysis
The analysis of the distillate was carried out using a Carle GC (Series 500)
fitted with a 30 m long fused silica capillary column and a flame ionization
detector (FID). The temperature in the oven was programmed to operate from
45 C to 200 C at 7 C/min. The analysis took about 35 min. Various peaks in
the chromatogram were identified by GC/MS analysis and using pure com-
pounds. The components were classified into acids, alcohols, aldehydes and
ketones, aromatic hydrocarbons, aliphatic hydrocarbons, furans and pheno-
lics. The components which could not be identified were included in ' uniden-
tified fraction'.
Density and viscosity measurements were done at 25 C using a 10 ml pyc-
nometer and Brookfield cone and plate digital viscometer (model RVTDCP),
respectively. The elemental analyses of the distillate and residue fractions,
obtained at 172 Pa and 200C, were performed on days 1, 16 and 31 using a
Perkin Elmer 2400 CHN Analyzer.
RESULTS AND DISCUSSION
Distillation characteristics of the fresh bio-oil
The distillation characteristics and the composition of the distillate are pre-
sented in Table 1. The results show t hat the amount of distillate increased from
21 wt.% at 85C to 62 wt.% at 200C before decreasing to 43 wt.% at 250C.
The amount of residue decreased with an increase in temperature to a mini-
mum of 37 wt.% at 200 C. The increase in the amount of residue above 200 C
may be due to polymerization of some components as reported by Sharma and
Bakhshi [11]. Also, Molton et al. [5] have shown that bio-oils are character-
istically unstable to light, air and temperature changes and they have the ten-
dency to easily polymerize when exposed these conditions.
Composition of the distillate fraction from fresh bio-oil
The analysis of the distillate, presented in Table 1, indicates that the distil-
late is composed of a large number of oxygenated compounds and C6-Cll hy-
drocarbons. The oxygenated compounds included acids, cyclic alcohols, ali-
phatic alcohols, aldehydes, cyclic ketones, substituted furans, ethers and alkyl
substituted and methoxy phenols. The hydrocarbons consisted of aromatic,
polycyclic and long chain unsaturated hydrocarbons. In all, 85 individual com-
pounds were identified by GC/MS. The distillate contained 19-23 wt% aro-
matic hydrocarbons, 11-35 wt% phenols, 9-13 wt% naphthenes, 5-11 wt%
aliphatic hydrocarbons, 7-14 wt% aldehydes and ketones and 3-10 wt% alco-
hols. The concentrations of acids, ethers and furans were in the range 2-6 wt%
J.D. Adjaye et al. / Fuel Processing Technol. 31 (1992) 241-256 245
TABLE 1
Di sti l l ati on characteristics of fresh bio-oil showing the amount and composi ti on (wt %) of distil-
late at different temperatures
Parameter Temperature, C
85 115 140 165 175 190 200 220 250
Distillation characteristics
Distillate 2.10 22.5 30.0 35.0 38.5 47.0 62.3 52.6 43.5
Residue 79.0 77.5 70.0 65.0 61.5 53.0 37.7 47.4 56.5
Total 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0
Distillate composition
Acids 4.3 4.4 3.7 2.8 2.6 2.5 2.5 2.5 2.3
Alcohols 7.2 10.3 8.6 3.8 3.5 3.6 6.6 5.3 5.2
Aldehydes + ketones 9.1 11.0 9.0 12.4 14.2 10.0 7.8 7.7 7.5
Aliphates 11.1 10.3 7.6 6.5 6.4 6.0 5.5 5.5 5.5
Aromates 19.2 20.8 21.8 23.8 23.7 23.7 21.6 21.3 21.1
Ethers 4.9 6.2 5.4 3.1 3.8 3.5 2.5 2.6 2.5
Furans 4.0 4.0 3.8 2.2 3.0 3.2 3.4 3.4 3.6
Naphthenes 9.2 9.8 12.4 13.0 13.7 12.1 12.0 12.0 11.9
Phenols 12.6 11.0 16.8 19.0 21.4 24.7 28.1 29.8 35.2
Uni denti fi ed 18.5 12.3 10.1 13.3 7.7 11.7 9.9 9.8 4.9
fraction a
Total 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0
aDetermined by difference.
each, whereas 7- 18 wt% of t he di st i l l at e remai ned uni dent i f i ed. The phenol i c
f ract i on compos ed mai nl y of phenol , guai acol , p-cresol , p- and o-guai acol , is-
oeugenol and catechol . The acids cons i s t ed of formic acid, acetic acid and traces
of propi oni c acid. Si mi l ar c ompone nt s were reported earlier by Shar ma and
Ba khs hi [ 11 ] i n t he hi gh pressure l i quef act i on oi l and by El l i ot et al. [ 16 ] and
Sol t es and Li n [ 17] i n t he hi gh t emperat ure pyrol yti c oils.
The t ot al amount s of oxygenat es and hydrocarbons decreased above 200 C
(Fig. 1) . Wi t h i ncrease i n temperature, t he concent rat i ons of acids, ethers,
al i phat i c hydrocarbons and al cohol s decreased. On t he other hand, t he con-
cent rat i ons of aromati c hydrocarbons, al dehydes and ket ones and na pht he ne s
s howed a ma x i mum around 175 C ( Tabl e 1 ). Int erest i ngl y, t he phenol i c frac-
t i on c ont i nuous l y i ncreased wi t h t emperat ure. The i ndi vi dual concent rat i ons
of major c ompone nt s at 175, 200 and 250 C are gi ven i n Tabl e 2. Except eu-
genol whos e concent rat i on i ncreased t o about 28 wt% of t he di st i l l at e f ract i on
at 250C, t he concent rat i ons of mos t c ompone nt s were bel ow 2 wt%. Si mi l ar
c ompounds have been reported i n l i terature from t he dry di s t i l l at i on of l i gni n
[ 18] . The aromat i cs and phe nol s and t hei r al kyl s ubs t i t ut ed f ract i ons al so
246 J.D. Adjaye et al. / Fuel Processing Technol. 31 (1992) 241-256
- - ~ Pr essur e, 172 Pa
" 6
%
,~ 6 o
c
o 40
Q.
E
o
0
20.
" 6
01
o , ; o 2 ; . . . .
Temperature, oc
Fig. 1. Rel at i onshi p of t emper at ur e and composi t i on of fresh bio-oil. ( [ ] ) Oxygenates, ( ) hy-
drocarbons, and ( ) residue.
TABLE 2
Vari at i on in t he concent r at i on (wt %) of most abundant component s of bio-oil distillate wi t h
t emper at ur e and storage t i me, at 172 Pa
Component Ti me, days
1 16 31
Temperat ure, C
175 200 250 175 200 250 175 200 250
Formi c acid, 1-
met hyl propyl est er
Cycl opent anol , 2-met hyl - 0.3
Cyclohexanol, 1-methyl- -
4- ( 1-met hyl et henyl ) -
Cycl opent anone, 2-ethyl- 3.1
Benzene, (1-ethyl-2-
met hyl - ) -
Benzene, 2-but enyl - 4.3
Cyclooctadiene, 1,5- 1.5
di met hyl -
Phenol 2.6
Phenol , 4-met hyl - 4.0
Phenol , 2-met hoxy-4- (2- 5.9
propenyl ) -
- 0.2 0.3 0.5 0.1 0.2 0.1
0.9 2.1 0.1 0.3 0.9
- 0.9 1.1 0.9 0.6 0.7 0.6
1.3 1.1 1.7 1.0 0.4
- 2.9 2.4 1.9 3.6 3.0 2.5
3.8 3.6 3.4 2.1 2.0 1.7 1.8 1.7
1.2 1.1 0.8 0.2 0.8 0.3
1.9 1.6 0.8 0.4 0.3 0.4 0.3 0.3
1.6 1.9 0.7 0.6 0.4 0.7 0.6 0.3
23.0 28.3 14.5 17.1 19.5 12.4 16.8 18.6
J.D. Adj aye et al. / Fue l Processing Technol. 31 (1992) 241-256 247
have been reported as products from dry distillation of lignin [18] and wood
pyrolysis [19]. Also, it was suggested t hat such products were formed by re-
combination and cyclization reactions, via Aldol condensation, from C2, C3 and
C4 fragments which were the initial degradation products [5]. Further reac-
tions may yield furans, aldehydes and ketones. The corrosivity and the appar-
ent instability of these bio-oils are caused by the presence of aromatics and
phenols [5,6,19 ].
St abi l i t y analysis of the bio-oil
As mentioned earlier, the bio-oil stability was characterized by observing the
changes in its physical properties, distillation characteristics and composition
with time. Any significant changes in these characteristics with time indicated
a poor stability. The results are presented in Table 3. It is seen that the viscos-
ity of the bio-oil increased with time probably due to an increase in the con-
centration of high molecular weight compounds. A higher viscosity implies
poor flow characteristics. The density of the bio-oil remained unchanged. The
amount of distillate at 175 C decreased from 38 wt% for the fresh bio-oil to 32
wt% after 31 days. Similar decreases were observed at 200 and 250C. The
decrease in distillate amount with time lead to an increase in the amount of
residue. The increase in the residue suggests that some polymerization reac-
tions occurred in the bio-oil during its storage.
The compositions of the distillates obtained on days 1, 16 and 31 are given
in Table 3 and also are plotted in Figs. 2 and 3. The figures indicate that both
the total oxygenates and hydrocarbon fractions decreased with time. The con-
centrations of aromatic hydrocarbons and phenols decreased while the con-
centration of aldehydes and ketones increased. Acids, alcohols, furans and al-
iphatic hydrocarbons remained unchanged. These results suggest that the poor
stability of the bio-oil may be due to the presence of phenols, aromatic hydro-
carbons and aldehydes and ketons.
The concentrations of the individual components in the distillate at differ-
ent times, given in Table 2, indicated that there was a disappearance or de-
crease of some components alongwith the formation of others. The decrease
in the concentration of 2-methylcyclopentanol, 2-ethylcyclopentanone, 2-bu-
tenylbenzene, 1,5-dimethylcyclooctadiene, phenol, 4-methylphenol and 2-
methoxy-4,2-propenylphenol indicates that they may be involved in the poly-
merization reactions. Other compounds such as 1-methylpropyl ester formic
acid, 1-methyl-4- (1-methyl ethenyl) cyclohexanol and ethyl-2-methylbenzene
which were not present in significant amounts in the fresh bio-oil distillate
were formed during storage. Although aromatics and phenols are known to be
stable at low temperatures, phenolic units in wood extracts or in lignin building
units have been reported to be intermediates in the formation of colored con-
248 J.D. Adjaye et al. / Fuel Processing Technol. 31 (1992) 241-256
TABLE 3
Effect of storage t i me on t he amount and composi t i on (wt %) of distillate from bio-oil at different
temperatures, at 172 Pa
Paramet er Time, days
1 16 31
Temperature, C
175 200 250 175 200 250 175 200 250
Distillation characteristics
Distillate 38.5
Residue 61.5
62.3 43.5 36.0 54.1 38.3 32.4 51.0 33.4
37.7 56.5 64.0 45.9 61.7 67.6 49.0 66.6
Total 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0
Distillate composition
Acids 2.6
Alcohols 3.5
Aldehydes + ketones 14.2
Aliphates 6.4
Aromates 23.7
Et hers 3.8
Furans 3.0
Napht henes 13.7
Phenolics 21.4
Uni dent i fi ed 7.7
fraction a
2.5 2.3 2.5 2.6 2.0 2.1 1.9 1.8
6.6 5.2 3.2 5.8 5.2 2.8 4.7 4.5
7.8 7.5 14.8 11.5 11.1 16.8 14.0 13.8
5.5 5.5 6.0 5.1 5.1 5.6 4.8 4.7
21.6 21.1 21.0 20.0 20.1 20.1 18.6 18.3
2.5 2.9 5.6 4.8 6.1 6.6 5.1 6.2
3.4 3.6 3.2 3.3 3.5 3.0 3.6 3.7
12.1 11.9 15.1 12.9 11.8 15.8 12.6 11.4
28.1 35.2 20.2 26.4 29.0 19.0 25.8 29.4
9.9 4.9 8.4 7.6 6.1 8.2 8.9 6.2
Total 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0
Viscosity and densi t y of the bio-oil at 25C
Viscosity, kPa s 313
Density, kg m -3 1000
320 323
100 1000
aDetermined by difference.
densat i on product s from alkaline pulping [5]. The pol ymeri zat i on to conden-
sation product s has been report ed to be slow and occurred even at room tem-
perat ure [5]. The above observat i ons appear to confirm t hat t he bio-oil
instability may be due to t he presence of aromatics, phenols or aldehydes and
ketones.
The el ement al analyses of various fractions are present ed in Table 4. The
analyses show t hat t he oxygen cont ent of t he distillate at 175C decreased
from 22.7 wt% on day 1 to 18.8 wt% on day 31. The oxygen cont ent of t he
residue increased from 16.3 wt% on day 1 to 21.9 wt% on day 31 while its carbon
cont ent decreased from 75.6 wt% to 71.1 wt% in t he same period. The increase
J.D. Adjaye et al. / Fuel Processing Technol. 31 (1992) 241-256 249
0
4o
c
0
o
0
. o
(0
" I "
P r e s s u r e , 1 7 2 Pa
l o
15o 200
Temperature, oc
I
250
Fig. 2. Relationship of temperature and total hydrocarbon concentration in bio-oil at different
storage times. (-Q-) Bio-oil, day 1; (-m-) bio-oil, day 16; (-[~-) bio-oil, day 31; and ( - - ~- - ) bio-
oil + tetralin, days 1-31.
o
e:
c
o
c
o
0
o
10150
P r e s s u r e , 1 7 2 Pa
2OO 25O
Temperature, oc
Fig. 3. Relationship of temperature and total oxygenates concentration in bio-oil at different stor-
age times. ( - 0- ) Bio-oil, day 1; ( - . - ) bio-oil, day 16; (-[~-) bio-oil, day 31; and ( - - ~ - - ) bio-
oil + tetralin, days 1-31.
in the oxygen content of the residue suggests that the oxygenates may be the
precursors to the formation of high molecular weight compounds.
The above results indicate that the bio-oil was unstable and its properties
250 J.D. Adjaye et al. / Fuel Processing Technol. 31 (1992) 241-256
TABLE 4
Effect of storage ti me on the elemental analyses (wt %) of distillate and residue fractions at dif-
ferent temperatures, at 172 Pa
Analysis Ti me, days
1 16 31
Temperature, C
175 200 250 200 200
Bio-oil
Distillate
C 68.4 67.7 67.7 67.9 70.7
H 9.2 9.6 10.8 10.5 10.5
N 0.0 0.0 0.0 0.0 0.0
04 22.4 22.7 21.5 21.6 18.8
Residue
C 75.6 76.3 73.9 73.9 71.1
H 8.1 7.8 6.7 7.7 7.0
N 0.0 0.0 0.0 0.0 0.0
0 b 16.3 15.9 19.4 18.4 21.9
Feed (bio-oil + tetralin)
Distillate a
C 76.8 76.4 73.9 74.2 73.9
H 8.3 8.9 8.7 8.6 8.3
N 0.0 0.0 0.0 0.0 0.0
0 b 14.9 14.7 17.4 17.2 17.8
Residue
C 78.4 82.4 82.6 82.4 82.2
H 8.3 7.2 7.1 7.4 7.4
N 0.0 0.0 0.0 0.0 0.0
04 13.3 10.4 10.3 10.2 10.4
qncl udes tetralin.
bDetermined by difference.
changed with time. The poor stability probably was due to oxi dat i on/ pol y-
merization reactions which occurred during storage. Since the bio-oil was pro-
duced by the thermochemi cal conversi on of wood powder at high pressure (5
MPa) and temperature ( 360C) , it is possible that these severe condi ti ons
induced some cleavage of bonds in the bi omass that lead to free radical for-
mati on [17]. The normal C- C bond energy is about 380 kJ/ mol and thermal
exci tati on of mol ecul es becomes sufficient to break these bonds at tempera-
tures of 350- 550 C. In addition, certain bonds are excepti onal l y weak (espe-
cially O-O bonds) and therefore can break at lower temperatures. Compounds
with such bonds can initiate radical processes at 50- 150C [21]. Similar re-
J.D. Adjaye et al. / Fuel Processing Technol. 31 (1992) 241-256 251
sults have been hypothesized for coal pyrolysis and liquefaction [ 14,20 ]. As a
result, in the absence of a stabilizing agent such as a good hydrogen donor
solvent, secondary substitution and radical combination reactions (polymeri-
zation reactions) may occur during storage leading to higher molecular weight
products [ 14,20 ]. March [ 22 ] discussed coupling reactions involving free rad-
icals of aromatic nature. Based on March' s development, a mechani sm is sug-
gested for t he disappearance and reduction of the aromatic compounds (Table
2). It is based on aromatic substrate substitutions involving the coupling of
t he aromatic rings. Some aromatic free radicals, Ar , may be produced by bond
cleavage at t he liquefaction stage. If ArH is anot her aromatic ring the possible
overall reaction is,
Ar + Ar H- , Ar - Ar + H" (1)
The first step of t he polymerization process may involve the radical attacking
the ring as in reaction (2). The reaction (2) may t ermi nat e by simple coupling
leading to the formation of a higher molecular weight compound as in reaction
(3).
H Ar
-
H Ar Ar H
H Ar
(3)
Since reactions involving free radicals are slow at room t emperat ure [ 7 ] t he
increase in the residue and viscosity of t he bio-oil with time was slow. Gener-
ally, it has been considered t hat phenol at e radicals give stable molecular prod-
ucts by coupling in pairs [7]. The rate of disappearance of these radicals has
been shown to follow second order kinetics, t hus demonst rat i ng coupling or
dimerization [ 7 ] according to the reaction
ArO" + ArO" -~ Dimer (4)
Furt her reactions could take place to form polymeric products. These reactions
may be slow especially when there is steric hindrance [ 23 ]. In bio-oils, which
have high molecular weights and long molecules, the effect of steric hindrance
may be enormous leading to slow polymerization reactions involving t he phen-
olate molecules.
Stability analysis of bio-oil plus tetralin mixture
The effect of tetralin addition on bio-oil viscosity, density, chemical com-
position and distillation characteristics was investigated on days 1, 16 and 31.
252 J.D. Adjaye et al. ,/Fuel Processing Technol. 31 (1992) 241-256
The results were calculated on tetralin-free basis and are presented in Table
5. The table indicates that the viscosity, density and composition of the bio-
oil remained unchanged on days 1, 16 and 31. The concentrations of the hy-
drocarbons and oxygenates also remained unchanged with storage time as in-
dicated in Figs. 2 and 3. A comparison of the distillation characteristics of fresh
bio-oil (Table 3) and fresh feed (Table 5) showed that some bio-oil properties
changed when tetralin was added to the bio-oil. As expected, the viscosity de-
creased and there was a slight decrease in the amount of distillate. The com-
position of the distillate also changed slightly (Tables 3 and 5). For example,
the concentrations of acids, alcohols and aliphatic hydrocarbons in the feed
were higher and the concentrations of naphthenes and phenols were lower
TABLE 5
Effect of storage time on t he amount and composi t i on of (wt%) of distillate from feed (bio-
oil + tetralin ), at 172 Pa
Paramet er Time, days
1 16 31
Temperat ure, C
175 200 250 175 200 250 175 200 250
Distillation characteristics
Distillate 36.4 59.4 43.4 36.1 59.0 42.9 36.7 50.2 42.7
Residue 63.6 40.6 56.6 63.9 41.0 57.1 63.3 40.8 57.3
Total 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0
Distillate composition
Acids 5.3 5.5 5.0 6.0 5.3 4.9 5.8 5.3 4.9
Alcohols 7.7 6.9 4.3 8.0 7.1 4.2 7.9 7.1 4.6
Aldehydes + ketones 10.0 9.1 8.1 9.7 9.3 8.4 9.3 9.2 8.4
Aliphates 8.7 8.6 8.7 8.6 9.0 8.8 9.0 9.0 8.9
Aromat es 24.6 20.8 18.7 24.2 21.0 19.0 24.4 20.6 19.1
Et hers 2.4 3.1 3.6 2.4 3.4 3.6 2.1 3.3 3.6
Furans 3.5 3.8 1.7 3.6 4.0 1.6 3.2 3.9 1.8
Napht henes 10.1 10.6 10.6 10.1 10.4 10.8 10.3 10.6 11.0
Phenolics 19.7 23.8 25.1 19.4 24.0 25.0 20.0 24.3 25.6
Uni dent i fi ed 7.3 7.8 14.2 7.8 6.5 13.7 7.5 6.7 12.1
fraction a
Total 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0
Viscosity and density of the feed at 25 C
Viscosity, kPa s 26
Density, kg m -3 985
26 26
985 985
~Determined by difference.
J.D. Adjaye et al. / Fuel Processing Technol. 31 (1992) 241-256 253
than those in the bio-oil. Furans, ethers, aromatic hydrocarbons and aldehydes
and ketones remained unchanged.
The elemental analyses of the distillate and residue fractions from the bio-
oil-tetralin feed mixture are presented in Table 4. It is seen that the oxygen,
carbon and hydrogen contents of residue fraction did not change with time. It
should be noted that the elemental analysis of the distillate also includes te-
tralin in the feed mixture. These results indicate that the addition of tetralin
as a hydrogen-donor solvent helped to stabilize the bio-oil and also lowered its
viscosity.
Role of tetralin
Since the bio-oils are hydrogen defficient, the addition of hydrogen may be
done under pressure or by using a hydrogen-donor solvent which supplies hy-
drogen through dehydrogenation reactions. The dehydrogenation of tetralin
to naphthalene liberates hydrogen radicals according to reaction (5). Each
molecule of tetralin can provide a maximum of four hydrogen free radicals.
Figure 4 shows t hat the concentration of naphthalene in the bio-oil-tetralin
feed increased to about six times from t hat in
H
CA)
r
AI
It
, 4 H " ( 5 )
Step I Biomass -+ 2R
Step2 R' +DH2 - ~ RH+DH
Step 3 R +M- H~ RH +M
St ep4 R+DH - ~ RH+D
Step 5 R +M- H--,RH + M"
Step 6 M' + M" --, 2M
(6)
the bio-oil due to tetralin dehydrogenation. Based on these observations, the
apparent stability of the bio-oil in the presence of tetralin may be explained
on the basis of free radical reactions and the hydrogen donor properties of
tetralin, as reported by Bockrath [14] and McMillen et al. [20] for coal liq-
uefaction. According to the postulated mechanism representedby reaction (6),
the dissociation of some bonds in the biomass generates two free radicals, R.
The free radicals then abstract hydrogen from the donor solvent, DH2, or some
high molecular weight fraction, M, or the free radicals produced by these steps.
As a result of the dehydrogenation of solvent or some part of the high molecular
weight fraction, the fragments from the fractured bond are stabilized.
254 J.D. Adjaye et al. /Fuel Processing Technol. 31 (1992) 241-256
m
0
, I
. D
o
m
( 3
C
0
0
= =
oJ
m
m
J ~
x :
Q .
o
Z
P r e s s u r e , 1 7 2 P a
o o , ,
1 5 o 2 0 0 2 5 0
Temperature, oc
Fig. 4. Relationship of temperature and naphthalene concentration in bio-oil and feed (bio-
oil+tetralin). (bottom curve) Bio-oil, and (top curve) bi o-oi l +tetral i n.
The free radicals produced from Step 3 also may combine in Step 6 to form
high molecular weight fraction. As a result of the increase in heavier fraction,
the residual fraction increases with time.
CONCLUSIONS
(1) The bio-oil, produced from the high pressure liquefaction of aspen po-
plar wood, was found to be a complex mixture of acids, alcohols, aldehydes,
ketones, aromatic and aliphatic hydrocarbons, furans, ethers and phenols.
(2) A maximum amount of distillate product was obtained from bio-oil at
200 C.
(3) The bio-oil was unstable and its viscosity and chemical composition
changed with time due to polymerization or the oxidative coupling of some of
the bio-oil components.
(4) The stability of the bio-oil improved in the presence of tetralin. This
was explained on the basis of a free radical mechanism and the hydrogen donor
properties of tetralin.
REFERENCES
1 Prasad, Y.S., Bakhshi, N.N., Mathews, J.F. and Eager, R.L., 1986. Catalytic conversion of
canola oil to fuels and chemical feedstocks over HZSM-5. Can. J. Chem. Eng., 64: 278.
2 Scott, D.S. and Piskorz, J., 1982. The flash pyrolysis of aspen poplar wood. Can. J. Chem.
Eng., 60: 666.
J.D. Adjaye et al. / Fue l Processing Technol. 31 (1992) 241-256 255
3 Eager, R.L., Mathews, J.F. and Pepper, J.M., 1982. Liquefaction of aspen poplar wood. Can.
J. Chem. Eng., 60: 289.
4 Diebold, J. and Scahill, J., 1988. Production of primary pyrolysis oils in a vortex reactor. In:
E.J. Soltes and T.A. Milne (Eds.), Pyrolysis Oils from Biomass: Producing, Analysing and
Upgrading. ACS Symposium Series 376, Am. Chem. Soc., Washington, DC, p. 31.
5 Molton, P.M., Demmitt, T.F., Donovan, J.M. and Miller, R.K., 1978. Mechanism of conver-
sion of cellulose wastes to liquid fuels in alkaline solution. In: D.L. Klass (Ed.), Energy from
Biomass and Wastes III. Institute of Gas Technology, Chicago, IL, p. 293.
6 Eager, R.L., Pepper, J.M., Roy, J.C. and Mathews, J.F., 1983. Chemical studies on oils de-
rived from aspen poplar wood, cellulose and isolated aspen poplar lignin. Can. J. Chem., 61:
2010.
7 Vasilakos, N.P. and Austgen, D.M., 1985. Hydrogen-donor solvents in biomass liquefaction.
Ind. Eng. Chem. Process Des. Develop., 24: 304.
8 Leroy, J., Choplin, L. and Kaliaguine, S., 1988. Rheological characterisation of pyrolytic
wood-derived oils: existence of a compensation effect. Chem. Eng. Commun., 71: 157.
9 Chantal, P., Kaliaguine, S., Grandmaison, J.L. and Mahay, A., 1984. Production of hydro-
carbons from aspen poplar pyrolytic oils over HZSM-5 catalyst. Appl. Catal., 10: 317.
10 Sharma, R.K. and Bakhshi, N.N., 1989. Upgrading of biomass-derived pyrolytic oils over
HZSM-5 catalyst. Report of Contract File No. 058SZ-23283-8-6116. Bioenergy Development
Program, Energy, Mines and Resources, Canada, p. 79.
11 Sharma, R.K. and Bakhshi, N.N., 1991. Upgrading of wood-derived bio-oil over HZSM-5.
Bioresource Technol., 35: 57.
12 Renaud, M., Grandmaison, J. L., Roy, C. H. and Kaliaguine, S., 1987. Conversion of vacuum
pyrolytic oils from populus deltoides over HZSM-5. In: Producing, Analysing, and Upgrading
of Oils from Biomass. Symposium sponsored by Div. Fuel Chem. at the 193rd Meeting of the
Am. Chem. Soc. at Denver, CO, April 5-10. ACS Symp. Ser., Div. Fuel Chem., Prepr., 32 (2):
276.
13 Chen, N.Y., Walsh, D.E. and Koenig, L.R., 1988. Fluidized bed upgrading of wood pyrolysis
liquids and related compounds. In: E.J. Soltes and T.A. Milne (Eds.), Pyrolysis Oils from
Biomass: Producing, Analysing and Upgrading. ACS Symposium Series 376, Am. Chem.
Soc., Washington, DC, p. 277.
14 Bockrath, B. C., 1983. Chemistry of hydrogen donor solvents. In: M.L. Gorbaty, J.W. Larsen
and I. Wender (Eds.), Coal Science, Vol. 2. Academic Press, New York, NY, p. 65.
15 Dao, L.H., Haniff, M., Honte, A. and Lamothe, D. 1987. Reactions of model compounds of
biomass-pyrolysis oils over ZSM-5 zeolite catalyst. In: E.J. Soltes and T.A. Milne (Eds.),
Pyrolysis Oils from Biomass: Producing, Analysing and Upgrading. ACS Symposium Series
376, Am. Chem. Soc., Washington, DC, p. 328.
16 Elliot, D.C., Sealock, L.J. and Butner, S.R., 1988. Product analysis from direct liquefaction
of several high-moisture biomass feedstocks. In: E.J. Soltes and T.A. Milne (Eds.), Pyrolysis
Oils from Biomass: Producing, Analysing and Upgrading. ACS Symposium Series 376, Am.
Chem. Soc., Washington, DC, p. 179.
17 Soltes, E.J. and Lin, S.C.K., 1987. Chromatography of non-derivatized pyrolysis oils and
upgraded products. In: Producing, Analysing, and Upgrading of Oils from Biomass. Sympo-
sium sponsored by Div. Fuel Chem. at the 193rd Meeting of the Am. Chem. Soc. at Denver,
CO, April 5-10. ACS Symp. Ser., Div. Fuel Chem., Prepr., 32(2): 276.
18 Goheen, D.W., 1971. Low molecular weight chemicals in lignins. In: K.V. Sarkanen and C.H.
Ludwig (Eds.), Lignins: Occurrence, Formation, Structure and Reactions. Wiley-Intersci-
ence, New York, NY, p. 803.
19 Graef, M., Allen, G.G. and Krieger, B.B., 1981. Product distribution in the rapid pyrolysis of
biomass/lignin for production of acetylene. In: D.L. Klass (Ed.), Biomass as a Non-fossil
Fuel Source. ACS Symposium Series 144, Am. Chem. Soc., Washington, DC, p. 293.
256 J.D. Adjaye et al. / Fuel Processing Technol. 31 (1992) 241-256
20 McMillen, D.F., Malhotra, R. and Nigenda, E.S., 1987. The case for induced bond scission
during coal pyrolysis. In: Producing, Analyzing and Upgrading of Oils from Biomass. Sym-
posium sponsored by Div. Fuel Chem. at the 194th Meeting of the Am. Chem. Soc. at New
Orleans, LA, Aug. 31-Sept. 4. ACS Symp. Ser., Div. Fuel Chem., Prepr., 32(3): 180.
21 Pryor, W.A., 1966. Free Radicals. McGraw-Hill, New York, NY, p. 58.
22 March, J., 1977. Advanced Organic Chemistry: Reactions, Mechanisms and Structure.
McGraw-Hill, New York, NY, p. 620.
23 Taylor, W.I. and Battersky, A.R., 1967. Oxidative Coupling of Phenols. Marcel Dekker, New
York, NY, p. 1.

S-ar putea să vă placă și