You are on page 1of 13

Journal of Analytical and Applied Pyrolysis 117 (2016) 334–346

Contents lists available at ScienceDirect

Journal of Analytical and Applied Pyrolysis


journal homepage: www.elsevier.com/locate/jaap

Mechanisms of biomass pyrolysis studied by combining a fixed bed


reactor with advanced gas analysis
Yann Le Brech a , Liangyuan Jia a , Sadio Cissé a , Guillain Mauviel a , Nicolas Brosse b ,
Anthony Dufour a,∗
a
LRGP, CNRS, Université de Lorraine, 1 rue Grandville, 54000 Nancy, France
b
LERMAB, Université de Lorraine, BP239, 54506 Vandoeuvre les Nancy Cedex, France

a r t i c l e i n f o a b s t r a c t

Article history: Slow pyrolysis of different biomasses (miscanthus, douglas fir and oak) was conducted in a U-shape fixed
Received 25 August 2015 bed reactor combined to (1) on-line analysis of the vapors (primary tar) by soft photoionisation mass
Received in revised form 19 October 2015 spectrometry (single photoionisation time-of-flight mass spectrometer, SPI-TOFMS), (2) quantitative off-
Accepted 23 October 2015
line analysis of condensed vapors by GC*GC/MS-FID and (3) permanent gas analysis. The U-shape fixed
Available online 30 October 2015
bed reactor allows controlling external mass transfers, temperature of the bed, quenching char at targeted
temperatures and quantifying mass balances. The mass yields in solid, condensable products and gas are
Keywords:
studied between 200 ◦ C and 500 ◦ C. The soft on-line photoionisation analysis leads to a more interpretable
Biomass
Wood
mass spectra with few fragment ions and gives the evolution of the main markers of volatiles produced
Pyrolysis in real-time as a function of the temperature. The chemical structure of the markers analyzed on-line
Photoionisation by SPI-TOFMS is assigned by GC*GC/FID-MS analysis. The mechanisms of hemicelluloses, cellulose and
Mass spectrometry lignin pyrolysis in the complex network of biomass (including minerals catalytic effect) are discussed.
© 2015 Elsevier B.V. All rights reserved.

1. Introduction chemical mechanisms have been extensively investigated [6,7]


mainly by thermo-gravimetric analysis [8–11], mass spectrome-
The growing energy demand and the increase of greenhouse try analysis [4,12–15] and in-situ spectroscopic analysis [5,16,17].
gases due to the overexploitation of fossil resources make research These investigations have given important insights into the chem-
in renewable and neutral-carbon resources one of the most cru- ical mechanism of biomass pyrolysis. The analysis of the primary
cial challenges for humanity. Over the last four decades [1], a lot volatile compounds formed during biomass pyrolysis is an impor-
of research has been conducted on the conversion of biomass into tant issue to reveal the chemistry of pyrolysis.
fuel and chemical products. Therefore investigations concerning Lignocellulosic biomass including woody and herbaceous
lignocellulosic biomass, such as wood and herbaceous biomass, biomass are composed by three macromolecules (cellulose, hemi-
have garnered much interest [2]. Thermo-chemical transformation celluloses, lignin) organised in a complex network in the cell wall
processes (pyrolysis, gasification and combustion) are promising [18]. Cellulose is a polymer of ␤-d-glucopyranose. Its degradation
option to convert biomass in electricity, biofuels or chemicals [3]. occurs principally from 300 ◦ C to 400 ◦ C following different chem-
Pyrolysis is the first chemical step occurring in all these processes ical mechanisms. These mechanisms are favored or inhibited by
and consequently it is of tremendous importance to better under- the operating conditions (temperature, heat flux density) or the
stand the chemical mechanisms of pyrolysis in order to optimize presence of inorganics. Cellulose can be converted by three main
the selectivity of the reactors. pathways:
Pyrolysis is a primary decomposition step of biomass resulting
in the formation of char, primary tars and gases [4]. It con-
trols to a great extent the products distribution and composition
1) transglycosylation [4,19,20] which leads to the formation of
in thermochemical reactors [5]. Consequently pyrolysis primary
anhydro-sugars (levoglucosan);
2) dehydration [21–23] which induces the formation of unsatu-
rated bonds;
∗ Corresponding author. 3) open-ring/fragmentation [24–26] reactions leading to the for-
E-mail address: anthony.dufour@univ-lorraine.fr (A. Dufour). mation of light oxygenated compounds.

http://dx.doi.org/10.1016/j.jaap.2015.10.013
0165-2370/© 2015 Elsevier B.V. All rights reserved.
Y. Le Brech et al. / Journal of Analytical and Applied Pyrolysis 117 (2016) 334–346 335

Fig. 1. Scheme of the fixed bed reactor combined with SPI-TOFMS or cold traps for further GC*GC/MS-FID analysis.

Fig. 2. Mass balances for miscanthus slow pyrolysis (5 K min−1 ) as a function of the final temperature.

The compounds formed from cellulose (furans, The formation of anhydro-sugars (levoglucosane, levogalactosane
methylglyoxal,. . .) are produced by the combination of these or 1,4-anhydro-d-xylopyranose. . .) from glycosidic ruptures is also
different mechanisms. For instance the formation of furans needs described [31]. Furfural is one of the major compounds from xylan
open-ring reactions and dehydration in order to create furan ring pyrolysis [28]. Pyranones are also described as a product of hemi-
from the glucopyranose structure. celluloses pyrolysis [32,33].
Hemicelluloses degradation occurs on a broader range of Lignin is a phenolic macromolecule constituted of phenyl-
temperature than cellulose (from 200 to 350 ◦ C) because hemicellu- propane units linked together through various ether and
loses are formed by different molecules linked by various chemical carbon–carbon linkages [34,35]. The main conversion step of lignin
bonds which induce different thermal stabilities [27]. Xylan (the occurs over a large temperature range from 150 to 500 ◦ C [36,37].
polymer of xylose) is the most studied polymer for hemicelluloses At low temperature (150–300 ◦ C), oxygenated propyl chains are
pyrolysis [28,29]. At low temperature (200–250 ◦ C) hemicelluloses mainly converted to water by dehydration reactions [36] and
undergo dehydration reactions [6] and specific chemical function formaldehydes by the release of carbonyl functions localized on
fragmentations [28]. Acetyls, carboxylics and methoxyls moieties carbon C␥ [35]. The ether linkages (␣-O-4 and ␤-O-4) can react
are converted into formic acid, acetic acid and methanol respec- from 200 ◦ C [6] leading to the formation of unsaturated chains or
tively [28,30]. For higher temperature (250–350 ◦ C) the formation to the release of phenolic compounds with chemical structures
of furanic compounds (furans and furanones) is observed from the close to the native network. Besides, these phenolic compounds
combination of dehydration, open-ring and cyclization reactions. are composed of oxygenated aromatic rings substituted by an
336 Y. Le Brech et al. / Journal of Analytical and Applied Pyrolysis 117 (2016) 334–346

alkyl chain with two or three carbons (eugenol, 4-vinyl-guaiacol).


For temperature higher than 300 ◦ C, alkyl chain conversion is
global [38] and phenolic compounds without or with a short alkyl
chain (mainly one carbon such as methylguaiacol) are devolatilized
[4,39,40]. Methoxyl moieties become reactive from approximately
370 ◦ C (depending on the inorganic content of lignins) leading to
the formation of methanol and methane by demethoxylation and
demethylation reactions respectively. Consequently, demethoxy-
lated (phenol, methylphenol) and demethylated (catechol, cresol)
phenolic compounds are also released.
In the present work, a fixed bed reactor was combined with an
on-line analysis of volatiles by single photoionisation (SPI)-time of
flight mass spectrometry (TOF-MS) and off-line analysis after con-
densation followed by GC*GC/FID*MS analysis. The U-shape fixed
bed reactor allows controlling external mass transfer, temperature Fig. 4. Gas yields from the slow primary pyrolysis of miscanthus.
of the bed, quenching char and quantifying mass balance. The main
advantage of SPI-MS is that the ionisation is soft and leads to a more
final set temperatures (between 280 and 500 ◦ C) at a heating rate
interpretable mass spectra with few or no fragment ions [39].
of 5 K min−1 . The radial temperature gradient in the fixed bed was
studied (by various thermocouples positioned along the diameter
2. Material and methods of the fixed bed) and was below 10 ◦ C between the middle and the
border of the fixed bed of particles. Once the final set temperature
2.1. Characterization of biomass was achieved, the furnace was immediately moved down and the
reactor was immersed in ice–water in order to quench the reactions
In this study, Miscanthus × giganteus has been used because of (2 min cooling time from 500 ◦ C to 150 ◦ C). Char was cooled down
its high production yields and its low need in fertilizers. Miscant- under argon until 30 ◦ C and then collected for further analysis. The
hus was harvested in Lorraine (France, Lorraine). Douglas and oak traps used to collect condensable products were then weighted.
were harvested in the Haut-Beaujolais area (South-East France). The pyrolysis experiments with the on-line SPI-MS and the con-
They were milled and sieved to a particle size between 40 and densation system were performed separately.
100 ␮m. This particle size was kept constant for all characteriza-
tions and pyrolysis experiments. A complete characterization of 2.3. Analysis of condensable products by GC*GC/MS-FID
these biomasses is given in Supporting material.
The off-gas line which connected the quartz reactor to the cold
2.2. Pyrolysis procedure traps (Fig. 1) was heated at 350 ◦ C in order to prevent condensation.
Cold traps were glass tube fitted with glass beads (without solvent).
The pyrolysis of miscanthus, oak and douglas (800 mg) was The first trap was put inside an ice–water bath (0 ◦ C) in order to con-
conducted in a vertical U-shape quartz fixed bed reactor under dense the water and “heavy” compounds formed during pyrolysis.
atmospheric pressure. Fig. 1 presents the apparatus. The particles The two following traps were put in propanol-liquid nitrogen mix
(40–100 ␮m) were supported on a quartz sintered plate (20 mm (−60 ◦ C). A bubbler with methanol was placed after the three cold
internal diameter). This geometry allows a good control of temper- traps. All traps were washed with a total of approximately 20 mL of
ature, thanks to a thermocouple (0.5 mm diameter) placed inside methanol. An internal standard (1 ␮L, tetradecene, Sigma Aldrich
the fixed bed of biomass particles, and of mass transfers by flushing ref 87,187) was injected in the whole volume (∼20 mL methanol
the fixed bed with a given flow rate of argon (300 NmL min−1 ). The and condensables). The solution of condensed products (with the
reactor was heated by an electrical furnace from 20 ◦ C to various internal standard) was analysed by an Agilent 7890 gas chromato-

Fig. 3. Mass balances for oak and douglas slow pyrolysis (5 K min−1 ) as a function of the final temperature.
Y. Le Brech et al. / Journal of Analytical and Applied Pyrolysis 117 (2016) 334–346 337

Fig. 5. Chromatogram for the condensable products (primary tar) sampled for the slow pyrolysis of the three biomasses up to a final temperature of 500 ◦ C.
338 Y. Le Brech et al. / Journal of Analytical and Applied Pyrolysis 117 (2016) 334–346

graph coombined to an Agilent 5975C MS analyser. The solution


(1 ␮L) was injected with a split ratio of 10 into an Agilent HP-5MS
connected with a heart-cutting system (Dean’s switch) to an Agi-
lent DB-Wax123 column. These two columns were chosen because
of their complementary polarities. More details on the calibration
and predictive quantification method are available in the Support-
ing material. More than 100 oxygenated compounds have been
detected and 30 compounds have been clearly identified by MS
and the NIST database. Between 25 and 30 compounds have been
quantified for each biomass.

2.4. Analysis of permanent gases

Noncondensable gases were collected by a gas sampling bag


positioned after the bubbler. They were analysed on a ␮GC-Varian
490 equipped with four modules, composed of two molecular
sieves 5A, a PoraPlot U and a CP-Wax 52CB columns. ␮GC-490 Sig-
nal was calibrated using standard bottles (Air Liquide, France). Four
major gases (H2 , CO, CO2 and CH4 ) were quantitatively determined.

2.5. On-line analysis of volatiles by SPI-MS

The on-line SPI-MS system is a customized gas analyzer (Photo-


TOF-MS, Photonion GmbH, Germany). It was connected to the
off-gas line of the quartz reactor. A small fraction of the gas phase
products (4 mL/min) was sampled through a capillary line at 250 ◦ C
and injected to the photoionisation region of the SPI-MS. The SPI-
MS is composed of a reflectron time-of-flight mass spectrometer
(TOF-MS, resolution ∼2000 at m/z 78) equipped with an elec-
tron beam pumped argon-excimer vacuum ultraviolet (VUV) lamp
system (EBEL, modified E-Lux 126, Optimare, Germany) [41]. The
photon energy of VUV light is 9.8 eV (126 nm). Data were acquired
from the mass spectrometer with Acqiris AP240 (Agilent Technolo-
gies) average cards and displayed by a special software (Photonion
GmbH, Germany).

3. Results

3.1. Mass yields in gas, liquid and char

Fig. 2 presents the mass yields of gaseous, liquid and solid


products as a function of the final pyrolysis temperature for mis- Fig. 6. Effect of the final pyrolysis temperature on the yield of some key primary
canthus. The global mass balance is between 95 and 108 wt%. Char tar compounds from miscanthus: (a) low molecular weight molecules, (b) furan
derivatives, (c) phenol derivatives.
yield decreases sharply in the temperature range of 280 and 350 ◦ C
(75–40 wt%) whereas the condensable and gaseous yields increase
from 25 wt% to 48 wt% and 3 wt% to 8 wt%, respectively. From 350 ◦ C stabilities [6]. At the temperature of 500 ◦ C the char yield is quite
product yields evolve slowly, char yield decreases until 30 wt%, similar between the two wood species (∼23%). For miscanthus, char
condensable and gaseous yields slightly increase up to 56 wt% and yields at 300 ◦ C and 500 ◦ C are higher mainly due to the higher inor-
10 wt%, respectively. The water content in the condensed phase was ganic content (see Supplementary material). The high condensable
not determined. However, water was reported as a major compo- yields (∼67%) and the low gas yields (∼6%) indicate that the sec-
nent of the condensed phase [42]. The char yields obtained after ondary reaction of volatiles was not significant during our fixed bed
rapid quenching in the fixed bed were in good agreement with experiments.
thermogravimetric analysis demonstrating that the pyrolysis was
conducted close to the chemical regime (see Supporting informa- 3.2. Gas composition
tion).
Fig. 3 presents the yields in gaseous, liquid and solid products as Gases released during pyrolysis have been quantified as a func-
a function of temperature for oak and douglas. For these biomasses tion of final temperature and are presented in Fig. 4. CO2 and CO
only the final temperatures of 300 ◦ C and 500 ◦ C have been inves- display a similar pattern. H2 was detected at the temperature of
tigated. At the temperature of 300 ◦ C, the yield of gases is very low 500 ◦ C but not quantified because of a very low concentration. The
(∼1 wt% for douglas and ∼1.5 wt% for oak) and the char yield from formation of CH4 starts at 350 ◦ C and increases until 500 ◦ C. The
oak is lower compared to the char yield from douglas. This observa- emission of CO2 is mainly due to the cleavage of carboxyl functions
tion was confirmed by thermogravimetric analysis (See Supporting present in hemicelluloses and lignin [28,36]. The formation of CO
information). This could be explained by a different composition arises from the cleavage of carbonyl functions naturally present in
of oak, douglas and miscanthus. Douglas hemicelluloses are com- hemicelluloses and lignin but also from the cleavage of C O and
posed predominantly of glucomannan whereas oak hemicelluloses C O C bonds. The production of methane (CH4 ) occurs at a higher
are composed of xylan, therefore this induces different thermal temperature highlighting the demethylation of lignin methoxyl
Y. Le Brech et al. / Journal of Analytical and Applied Pyrolysis 117 (2016) 334–346 339

Table 1
Yields (mg/g dry biomass) of main compounds quantified by GC*GC/MS-FID.

No. Compounds CAS no. MW (g/mol) Formula Miscanthus Oak Douglas Ref. for qualitative
assignment

1 Acetic acid 64-19-7 60 C2 H4 O2 18.88 40.41 7.91 NIST


2 2-Propanone, 1-hydroxy 116-09-6 74 C3 H6 O2 10.25 4.38 3.60 NIST
3 Ethyl-1-propenyl ether 928-55-2 86 C5 H10 O 0.71 0.65 0.26 NIST
4 Methylglyoxal 78-98-8 72 C3 H4 O2 3.60 6.72 10.30 NIST
5 1-Penten-3-one 1629-58-9 84 C5 H8 O 0.00 2.25 0.58 NIST
6 Propanal, 2,3 dihydroxy 367-47-5 90 C3 H6 O3 14.24 17.39 29.49 NIST
7 Propanoic acid, 2-oxo-, methyl ester 600-22-6 102 C4 H6 O3 0.53 2.89 3.20 NIST
8 furfural 98-01-1 96 C5 H4 O2 6.82 9.19 3.71 NIST
9 2-Furanmethanol 98-00-0 98 C5 H6 O2 2.03 2.79 2.88 NIST
10 2(5H)-Furanone 497-23-4 84 C4 H4 O2 n.qa 2.99 2.15 NIST
11 2,cyclopenten-1-one, 2 hydroxy 10493-98-8 98 C5 H6 O2 6.42 4.05 2.94 NIST
12 Phenol 108-95-2 94 C6 H6 O 1.49 0.34 0.61 NIST
13 4-Hydroxy-5,6-dihydro-(2H)-pyran2one – 114 C5 H6 O3 0.00 1.25 3.33 [44]
14 Corylon 80-71-7 112 C6 H8 O2 0.95 0.48 0.35 NIST
15 Phenol 2-methyl 95-48-7 108 C7 H8 O 0.43 n.q* 0.44 NIST
16 Phenol, 4-methyl- 106-44-5 108 C7 H8 O 0.61 0.17 0.57 NIST
17 Phenol, 2-methoxy- 90-05-1 124 C7 H8 O2 1.30 0.84 2.05 NIST
18 Phenol, 2 methoxy-4-methyl 93-51-6 138 C8 H10 O2 n.qa 0.82 5.34 NIST
19 4 vinyl phenol 496-16-2 120 C8 H8 O 3.48 0.00 0.00 [45]
20 2-Furaldehyde, 5-(hydroxymethyl)- 67-47-0 126 C6 H6 O3 0.00 3.32 2.79 NIST
21 2-Methoxy-4-vinylphenol 7786-61-0 150 C9 H10 O2 1.84 1.05 2.71 NIST
22 Phenol, 2,6-dimethoxy- 91-10-1 154 C8 H10 O3 1.34 1.91 0.00 NIST
23 vanillic acid 121-34-6 168 C8 H8 O4 0.00 1.79 0.00 NIST
24 Isoeugenol 5932-68-3 164 C10 H12 O2 n.qa 0.43 4.41 NIST
25 Levoglucosan 498-07-7 162 C6 H10 O5 0.00 19.57 25.55 NIST
26 Coniferyl alcohol (Douglas)/ 4-vinyl syringol (Oak and Miscanthus) 458-35-5/28343-22-8 180 C10 H12 O3 n.qa 1.07 0.31 [45]
27 2-Propenal, 3-(4-hydroxy-3-methoxyphenyl)- 127321-19-1 178 C10 H10 O3 n.qa 0.96 0.04 [45]
28 Syringyl acetone 19037-58-2 210 C11 H14 O4 0.00 0.66 0.00 [45]
29 3,5-Dimethoxy-4-hydroxycinnamaldehyde 87345-53-7 208 C11 H12 O4 0.00 0.93 0.00 [45]

The molecular structures of all the species are given in the Supporting information.
a
n.q means not quantified, numbers refer to numbers in the chromatograms presented in Fig. 5.

groups remaining into the char and also the breaking of methy- occurs mainly in the range of 200–300 ◦ C (Fig. 6) corresponding
lene groups [36]. Concerning douglas and oak, the amount of gas to the degradation of hemicelluloses. For oak, acetic acid is still
released during slow pyrolysis (shown in Supplementary material) formed after 300 ◦ C, this observation is in agreement with TG analy-
are smaller but in agreement with literature [37,43]. sis which shows that oak degradation occurs at higher temperature
probably due to lower inorganics content. 2-Propanone,1-hydroxy
3.3. GC/MS analysis of condensed primary tar is more important for miscanthus pyrolysis whereas methylglyoxal
is a major product of douglas pyrolysis. Propanal, 2,3,dihydroxy
The condensed phase contains a wide variety of products as is an important product for miscanthus, oak and douglas. Lev-
revealed by GC/MS analysis (Fig. 5). Their complete quantification oglucosan is detected and quantified for oak and douglas but not
appears to be very difficult, some compounds being too heavy to be in miscanthus highlighting the catalytic effect of minerals which
quantified by GC. The GC-analysed products of miscanthus, oak and inhibit the transglycosylation pathway [46,47]. Before 300 ◦ C, no
douglas represent about 8.2%, 13.3% and 11.7% of the initial weight levoglucosan is observed for oak and douglas showing that trans-
of dry biomass respectively. The major identified compounds have glycosylation reaction is not yet started at this temperature. All
been quantified and they are presented in Table 1. Their chemi- compounds assigned to carbohydrate products formed before this
cal structures are given in the Supplementary material. Concerning temperature (300 ◦ C) mainly come from hemicelluloses degrada-
miscanthus, the evolution profiles of key compounds as a function tion. For instance, the formation of furfural which is a good indicator
of temperature have been plotted on Fig. 6. of hemicelluloses degradation (for xylan mainly) is important for
The markers from carbohydrates pyrolysis can be divided in dif- oak at 300 ◦ C indicating that the hemicelluloses degradation starts
ferent categories which depict the different chemical pathways of before 300 ◦ C. Furfural is not so important in douglas products high-
pyrolysis as functions of temperature and biomass composition: lighting the low xylose content in the native biomass. The formation
of 4-hydroxy-5,6-dihydro-(2H)-pyran-2-one is more important in
douglas and oak than in miscanthus. It could be explained by the
1) the light oxygenated compounds (acetic acid; 2-propanone,1-
catalytic effects of minerals [33]. The presence of an anhydro-
hydroxy; ethyl-1-propenyl ether; methylglyoxal; 1-penten-3-
sugar compound (not identified clearly) at 45.72 min of the GC
one; propanal 2,3,dihydroxy and propanoic acid, 2-oxo methyl
chromatogram was only detected in douglas products showing
ester) testify to the fragmentation/dehydration pathways;
the specific hemicelluloses composition of douglas as compared to
2) the furanic compounds (furfural; 2-furanmethanol; 2-
miscanthus and oak. This compound which is formed from hemicel-
furaldehyde, 5-(hydroxymethyl)-), furanones (2(5H)-furanone);
luloses evidences that hemicelluloses with a low mineral content
and cyclopentanone (2,cyclopenten-1-one, 2 hydroxy; corylon)
undergo the formation of anhydrosugar [33]. Futhermore, the for-
highlight the dehydration/open-ring/cyclisation mechanism;
mation of 4-hydroxy-5,6-dihydro-(2H)-pyran-2-one which is a C5
3) anhydro-sugars (levoglucosan) evidence the transglycosylation
compound is observed for douglas whereas douglas hemicellu-
pathway.
loses are mainly composed of C6 sugars. Further investigations
are necessary to understand the mechanism of formation of this
Acetic acid is more important according to miscanthus and oak
compound.
pyrolysis. Concerning miscanthus the evolution during pyrolysis
340 Y. Le Brech et al. / Journal of Analytical and Applied Pyrolysis 117 (2016) 334–346

HO O

HO O

6 OH O 1
O 6 2
OH
5 O
Dehydraon
O O
5 HO
4 OH O 3
HO 1 HO 4
O 2 O O OH O
3
O O
OH OH
HO

Release of Cyclizaon
levoglucosan
OH
6
O H OH
6
OH OH
O O
Transglycosylaon Open-ring 4
5
1
5 O HO O
4 OH O
HO HO 2
2
1
n
O
O O 3
O
3
OH OH
OH
HO

K-Uncatalyzed

OH OH
6 HO

O O
5 O
4 O
HO OH
1 HO
O 2 O n
O
3
OH OH
HO

Linear cellulose chain

K-catalyzed

Open-ring
OH

OH OH HO 1
6 6 Aldol
OH
5
OH Dehydraon 5 reacon
4 1 4 1 5 2
HO O HO O
2 O 2
O
3 3 4 3
OH OH O OH

Fig. 7. Simplified mechanism of cellulose pyrolysis highlighting the main degradation pathways and key volatile markers.

The formation of cyclopentenones (2,cyclopenten-1-one, 2 evolution as a function of temperature showing that they could be
hydroxy; corylon) is enhanced in miscanthus whereas the for- produced by the same chemical mechanism.
mation of anhydrosugar (levoglucosan), furan (2-furaldehyde,
5-(hydroxymethyl)-) is more important for oak and douglas. 3.4. SPI-TOFMS on-line analysis of volatiles
Futhermore, a furanone is detected (m/z 128, retention time
26.64 min, not quantified) and appears to be an important com- The mass spectra obtained for the three investigated biomasses
pound for oak and douglas. This compound has been assigned (fixed bed, 5 K min−1 , final temperature 500 ◦ C) are displayed in
to 5-methyl-4-oxotetrahydrofuran-2-carbaldehyde in a previous Fig. 8. By comparison with products identified by GC/MS and based
investigation from our group by synchrotron SPI/MS*MS [39]. The on previous studies [4,12,39,49], the main m/z which are molecular
selectivity concerning the formation of these compounds seems to ions from the key markers of pyrolysis mechanisms have been iden-
be led by the inorganic catalytic effects. As already described [11], tified. The main products from the pyrolysis of cellulose are furfuryl
furanones and anhydro-sugars are more important from biomasses alcohol and 2-cyclopenten-1-one, 2 hydroxy (m/z 98); furfural (m/z
with low minerals contents whereas cyclopentanone derivatives 96); 2-furaldehyde, 5-(hydroxymethyl)- (m/z 126) and 5-methyl-4-
and phenol are more important [48] for biomasses with high inor- oxotetrahydrofuran-2-carbaldehyde (m/z 128). The products from
ganics contents (especially potassium). Cyclopentenone could be transglycosylation are levoglucosan (m/z 162) or levoglucosan
formed by rearrangement (aldol reaction) and dehydration of car- fragments (m/z 144, 126, 98 and 97) induced by dissociative pho-
bohydrates. A global simplified mechanism for cellulose pyrolysis toionisation as already described [39], are observed. Peaks at m/z
is proposed in Fig. 7. 43, 57 and 73 are also analysed by SPI-MS and could originate from
Oxygenated aromatic compounds are good markers of lignin the fragmentations of carbohydrate product [13,50,51]. It corre-
pyrolysis and the substituted degree of these aromatic compounds sponds mainly to low molecular weight carbonyl compounds [13]
is a good indicator of the thermal severity. Concerning douglas, (m/z 43, C2 H3 O+ ). 3-Hydroxy-2-penteno-1,5-lactone (m/z 114) is
no syringyl derivative was detected which is in agreement with also observed. This compound has been identified in previous stud-
the low syringyl units content in coniferous woods. Concerning ies as a hemicelluloses marker [39,52]. It has been confirmed by
miscanthus, the evolution of phenol, 2,6-dimethoxy; 2-methoxy- GC/MS analysis and by comparison between literature data [44].
4-vinylphenol; phenol, 2 methoxy-4-methyl; phenol, 2-methoxy- Lignin markers are also detected for each biomass highlight-
and 4-vinylphenol has been plotted in Fig. 6. It is shown that these ing the difference in the native network of macromolecules.
lignin products are mainly formed between 280 and 350 ◦ C. Phenol, The main m/z ions were assigned to the following markers:
2,6-dimethoxy and phenol, 2-methoxy- follow exactly the same
Y. Le Brech et al. / Journal of Analytical and Applied Pyrolysis 117 (2016) 334–346 341

Fig. 8. Total mass spectra (SPI-TOFMS, 9.8 eV) accumulated on whole experimental time until 500 ◦ C.

phenol (m/z 94); phenol, 2 methoxy-4-methyl (m/z 138); 2- temperature around 250 ◦ C [6]. The m/z at 150 (2-methoxy-4-
methoxy-4-vinylphenol (m/z 150); phenol, 4-ethyl-2-methoxy vinylphenol) and m/z 180 (4-vinyl syringol), observed at slightly
(m/z 152, not quantified, 37.2 min on GC/MS chromatogram); phe- higher temperatures, could be produced probably by the combi-
nol, 2,6-dimethoxy- (m/z 154); isoeugenol (m/z 164); 2-Propenal nation of dehydration and the breaking of ether bounds [6]. The
3-(4-hydroxy-3-methoxyphenyl)- (m/z 178); coniferyl alcohol conversion of carbohydrates is highlighted by the evolution of
(Douglas) and 4-vinyl syringol (Oak) (m/z 180); propenylsyringol 3-hydroxy-2-penteno-1,5-lactone (m/z 114) which comes mainly
(m/z 194, not quantified, 61.4 min on the GC/MS chromatogram); from hemicelluloses thermal degradation [33,39] and carbohydrate
3,5-Dimethoxy-4-hydroxycinnamaldehyde (m/z 208). The profiles fragments (m/z 43, 57 and 86) from 220 ◦ C until 370 ◦ C with a max-
concerning the main markers for cellulose, hemicelluloses and imum around 300 ◦ C. This degradation becomes more important
lignin pyrolysis have been plotted as a function of temperature at higher temperature (300–350 ◦ C) with the release of furfuryl
and are presented in Figs. 9–11 for miscanthus, oak and douglas alcohol and 2-cyclopenten-1-one, 2 hydroxy (m/z 98) and the m/z
respectively. 126. The m/z 126 may be assigned to levoglucosenone (detected by
Concerning miscanthus, the first products (from 220 ◦ C) ana- GC/MS but not quantified).
lysed by SPI-MS are 4-vinylphenol (m/z 120) and 2-propenal The degradation of lignin structure is significant from 250 ◦ C to
3-(4-hydroxy-3-methoxyphenyl)-(m/z 178) (Fig. 9c). These prod- 450 ◦ C (Fig. 9d) with the release of many different compounds as
ucts could mainly come from the thermal degradation of the phenol, 2-methoxy-(m/z 124); phenol, 2,6-dimethoxy-(m/z 154);
acetyl functions localized at C␥ of the propyl side chains of catechol, 3-methoxy (m/z 140, not quantified, 36.06 min); phenol,
lignin monomers in the native network. The native lignin in Mis- 2 methoxy-4-methyl (m/z 138); isoeugenol (m/z 164); methylphe-
canthus × giganteus is highly acetylated at the C␥ possibly with nol (m/z 108); phenol (m/z 94). The profiles of these compounds
acetate or p-coumarate groups [53]. Consequently, the formation released as a function of temperature are in close agreement with
of 4-vinylphenol (m/z 120) and 2-propenal 3-(4-hydroxy-3- the comprehensive study of Fendt et al. [54]. These compounds
methoxyphenyl)-(m/z 178) could be due to the breaking of acetate come mainly from the conversion of the alkyl chain (phenol, 2-
functions. C␥ implied into carbonyl function is likely to break at methoxy-; phenol, 2,6-dimethoxy-; phenol, 2 methoxy-4-methyl)
342 Y. Le Brech et al. / Journal of Analytical and Applied Pyrolysis 117 (2016) 334–346

Fig. 9. On-line analysis of key markers by SPI-MS as a function of temperature for miscanthus pyrolysis, (a) hemicelluloses and carbohydrate fragments, (b) main cellulose
markers, (c) lignin markers formed at low temperature, (d) lignin markers formed at high temperature.

which induce C C linkage breaking. Compounds which require These products show the formation of a double bond in conjugation
demethoxylation reaction are also observed. Indeed phenol (m/z with the aromatic ring in the early stage of pyrolysis [4] which is
94) and methylphenol (m/z 108) were detected from 250 ◦ C to high described to come from the enhancement of H-donor species [58].
temperature (500 ◦ C). The catechol, 3-methoxy is also observed but H-addition reactions [59] could provoke the formation of prod-
only from 270 to 400 ◦ C. This compound is formed by the demethy- ucts with double bond in conjugation with the aromatic ring as
lation of syringol units enhanced by the catalytic effect of minerals isoeugenol and vinylguaiacol. H-donor species are formed from
[55]. hydrogen radicals [58] during aromatization mechanisms. These
The formation of volatiles starts at 220 ◦ C for oak slow radicals could come from the aromatization of hemicelluloses start-
pyrolysis (Fig. 10). Indeed, some lignin markers are detected 3,5- ing in the same range of temperature.
dimethoxy-4-hydroxycinnamaldehyde (m/z 208) and 2-propenal, As for miscanthus pyrolysis, the degradation of carbohydrates
3-(4-hydroxy-3-methoxyphenyl)-(m/z 178). The formation of in oak is highlighted by the evolution of 3-hydroxy-2-penteno-1,5-
these phenolic aldehydes in the first step of pyrolysis has been lactone (m/z 114) and carbohydrate fragments [50] (m/z 43, 73,
already described [56]. It is due to ether ␣-O-4 and/or ␤-O-4 split- 57). Ethyl-1-propenyl ether (m/z 86) is less significant compared
ting at about 200 ◦ C and 250 ◦ C respectively [57]. This reaction to m/z 114. The formation of 3-hydroxy-2-penteno-1,5-lactone is
should be combined with the formation of C O moieties on the favoured for oak. Minor fragmentation reactions are induced in
C␥ of the propyl side chain probably by dehydration [56] of the oak hemicelluloses probably due to the low concentration of inor-
C␥H2 OH moiety or conversion of the carboxyl group on ␥–position. ganics. Furfural (m/z 96) is also detected from 220 ◦ C to 400 ◦ C
The formation of these aldehydes could be also due to the large with a maximum at 270 ◦ C highlighting that the conversion of
quantities of radicals. Radicals are formed in the first step of lignin oak’s hemicelluloses (mainly xylan) occurs at this temperature. This
pyrolysis through the homolytic cleavage of the lignin ether link- observation is in accordance with the thermogravimetric analysis
ages which react with the C␥-hydrogen at the conjugated allyl (see Supplementary material).
position [58] to form aldehydes. At higher temperature, the formation of phenolic compounds
At slightly higher temperature, propenyl syringol (m/z 194, not with small alkyl chain (with one carbon) or without alkyl side chain
quantified, at 54.6 min on GC chromatogram); 4 vinyl syringol is observed. Vanillic acid (m/z 168, not quantified, at 45.81 min
(m/z 180); isoeugenol (m/z 164); 2-methoxy-4-vinylphenol (m/z on GC chromatogram); phenol, 2,6-dimethoxy-(m/z 154), phe-
150) and 2-propenal, 3-(4-hydroxy-3-methoxyphenyl)-(m/z 178) nol, 2-methoxy-(m/z 124), phenol, 2 methoxy-4-methyl (m/z 138)
are detected. These products are formed from ether-linkage (for are detected between 280 and 470 ◦ C. The catechol, 3-methoxy
propenyl phenolic compounds: m/z 164, m/z 194) combined to (m/z 140, not quantified, at 36.06 min on GC chromatogram) is
C C breaking (for vinyl phenolic compounds: m/z 150, m/z 180). also observed. This compound is formed by the demethylation of
Y. Le Brech et al. / Journal of Analytical and Applied Pyrolysis 117 (2016) 334–346 343

Fig. 10. On-line analysis of key markers by SPI-MS as a function of temperature for oak pyrolysis, (a) hemicelluloses and carbohydrate fragments, (b) main cellulose markers,
(c) lignin markers formed at low temperature, (d) lignin markers formed at high temperature.

syringol units induced by mineral catalytic effects as for miscant- dimer [4] formed by C␥ elimination [35]. The formation of phenolic
hus. compounds with small alkyl chain (phenol, 2 methoxy-4-methyl,
The phenolic volatiles composed of S-units seem to be produced m/z 138) or without alkyl chain (phenol, 2-methoxy-, m/z 124) is
at lower temperature than G-units. Indeed, phenol, 2,6-dimethoxy- observed.
(m/z 154) and catechol, 3-methoxy (m/z 140) are formed at lower The production of 3-hydroxy-2-penteno-1,5-lactone (m/z 114)
temperature as compared to phenol, 2-methoxy-(m/z 124) and is observed from 220 ◦ C to 400 ◦ C indicating the degradation of
phenol, 2 methoxy-4-methyl (m/z 138) (Fig. 10). Syringyl units are the hemicelluloses. The m/z 126 is assigned to 2-furaldehyde,
described to be depolymerized at lower temperature range base 5-(hydroxymethyl)-, m/z 98 to 2-cyclopenten-1-one, 2-hydroxy
on different studies comparing softwood and hardwood [56,60]. and 2-furanmethanol. The m/z 128 may be assigned to 5-
The m/z 124 signal at high temperatures could be also assigned to methyl-4-oxotetrahydrofuran-2-carbaldehyde. These compounds
dihydroxytoluene (not quantified, 35.63 min on GC chromatogram) are markers of cellulose degradation on the range of 300–380 ◦ C.
produced by demethylation of 2 methoxy-4-methyl (m/z 138). Fur- Despite their different structural compositions, oak and douglas
ther investigations are required to understand this mechanism. display relative similar pyrolysis profiles. Aldehyde phenolic (i.e. 2-
The m/z 126 is assigned to 2-furaldehyde, 5-(hydroxymethyl)- propenal, 3-(4-hydroxy-3-methoxyphenyl)-) compounds start to
, m/z 98 to 2-cyclopenten-1-one, 2-hydroxy and 2-furanmethanol be volatilized at low temperature (around 250 ◦ C to 400 ◦ C) fol-
and the m/z 128 was assigned to 5-methyl-4-oxotetrahydrofuran- lowed at slightly higher temperatures by alkyl phenolic compounds
2-carbaldehyde. These compounds are markers of cellulose (i.e. methoxy-4-vinylphenol). The formation of phenolic com-
degradation on the range of 300–380 ◦ C. The m/z 128 was clearly pounds with small alkyl chain (i.e. phenol, 2-methoxy-4-methyl
relatively more important for oak than miscanthus probably due to (m/z 138)) or without alkyl chain (i.e. phenol, 2-methoxy-(m/z 124))
lower inorganic catalytic effect as already discussed. is then observed at higher temperature (Fig. 12). Regarding mis-
Concerning douglas pyrolysis, the formation of volatiles starts canthus, the formation of these volatiles is less distinguishable
at 200 ◦ C (Fig. 11) with the release of 2-propenal, 3-(4-hydroxy- probably due to the catalytic effect of inorganic materials.
3-methoxyphenyl) (m/z 178). As for oak, the formation of this
compound is consistent with the formation of aldehyde compounds
in the early stage of pyrolysis. At slightly higher temperatures, prod- 4. Conclusion
ucts formed by ether-linkage (coniferyl alcohol; m/z 180) combined
to C C breaking (2-methoxy-4-vinylphenol, m/z 150; isoeugenol, The geometry of the fixed bed reactor allows good control of
m/z 164) are observed. The m/z 272 is assigned to enol ether temperature and mass transfers and thus studying accurately the
primary mechanisms of biomass pyrolysis. The combination of
344 Y. Le Brech et al. / Journal of Analytical and Applied Pyrolysis 117 (2016) 334–346

Fig. 11. On-line analysis of key markers by SPI-MS as a function of temperature for douglas pyrolysis, (a) hemicelluloses and carbohydrate fragments (b) main cellulose
markers (c) lignin markers formed at low temperature (d) lignin markers formed at high temperature.

Fig. 12. Simplified mechanism of lignin pyrolysis showing the formation of the different type of markers as a function of temperature.

SPI/MS for on-line analysis with GC/MS for the assignment of the molecules (lignin, cellulose, hemicelluloses) present in the biomass
chemical structure of main m/z markers is a powerful approach to network. The three investigated biomasses, grass (miscanthus) and
understand the chemistry of biomass pyrolysis and notably for the wood (oak and douglas fir), exhibit different compositions (inor-
evolution of lignin markers. This investigation has unraveled some ganics, sugars, lignin structure) and give rise to various volatiles.
of the different steps of pyrolysis according to the different macro- Inorganics present inside the native biomass is the most important
Y. Le Brech et al. / Journal of Analytical and Applied Pyrolysis 117 (2016) 334–346 345

parameter that influences the chemical mechanism during pyroly- [16] G.A. Zickler, W. Wagermaier, S.S. Funari, M. Burghammer, O. Paris, In situ
sis. This methodology (combining fixed bed and SPI-MS) could be X-ray diffraction investigation of thermal decomposition of wood cellulose, J.
Anal. Appl. Pyrolysis 80 (2007) 134–140, http://dx.doi.org/10.1016/j.jaap.
applied to study the thermal decomposition of all other carbona- 2007.01.011.
ceous materials. The EBEL lamp (9.8 eV) used for SPI-TOFMS is a [17] K. Kirtania, J. Tanner, K.B. Kabir, S. Rajendran, S. Bhattacharya, In situ
soft ionisation method although there is still the need of an even Synchrotron IR study relating temperature and heating rate to surface
functional group changes in biomass, Bioresour. Technol. (2014) 36–42,
softer and more selective ionisation method to detect anhydrosug- http://dx.doi.org/10.1016/j.biortech.2013.10.034.
ars without any fragmentation. [18] J. Zakzeski, P.C.A. Bruijnincx, A.L. Jongerius, B.M. Weckhuysen, The catalytic
valorization of lignin for the production of renewable chemicals, Chem. Rev.
110 (2010) 3552–3599, http://dx.doi.org/10.1021/cr900354u.
Acknowledgments [19] J. Piskorz, D. Radlein, D.S. Scott, On the mechanism of the rapid pyrolysis of
cellulose, J. Anal. Appl. Pyrolysis 9 (1986) 121–137, http://dx.doi.org/10.1016/
0165-2370(86)85003-3.
The financial support from the ANR French Agency through
[20] V. Mamleev, S. Bourbigot, M. Le Bras, J. Yvon, The facts and hypotheses
the project PYRAIM is gratefully acknowledged. Fundings from relating to the phenomenological model of cellulose pyrolysis:
the French State and Lorraine Region through the project CPER Interdependence of the steps, J. Anal. Appl. Pyrolysis 84 (2009) 1–17, http://
dx.doi.org/10.1016/j.jaap.2008.10.014.
MEPP is also kindly acknowledged for a part of the purchase of the
[21] W. Chaiwat, I. Hasegawa, T. Tani, K. Sunagawa, K. Mae, Analysis of
SPI-TOFMS system. We also wish to thank Michel Mercy (CNRS, cross-linking behavior during pyrolysis of cellulose for elucidating reaction
Nancy) for his technical support and kind help with some experi- pathway, Energy Fuels 23 (2009) 5765–5772.
mental measurements. Dr. Sven Ehlert and Dr. Matthias Bente-von [22] J. Scheirs, G. Camino, W. Tumiatti, Overview of water evolution during the
thermal degradation of cellulose, Eur. Polym. J. 37 (2001) 933–942, http://dx.
Frowein (Photonion Gmbh, Germany) are kindly acknowledged for doi.org/10.1016/s0014-3057(00)00211-1.
the development of the customised SPI-TOFMS system. [23] F.J. Kilzer, A. Broido, The nature of cellulose pyrolysis, Pyrodynamics 2 (1965)
151–163.
[24] G.N. Richards, Glycolaldehyde from pyrolysis of cellulose, J. Anal. Appl.
Appendix A. Supplementary data Pyrolysis 10 (1987) 251–255.
[25] D.K. Shen, S. Gu, The mechanism for thermal decomposition of cellulose and
its main products, Bioresour. Technol. 100 (2009) 6496–6504.
Supplementary data associated with this article can be found, in [26] J. Piskorz, D.S.A.G. Radlein, D.S. Scott, S. Czernik, Pretreatment of wood and
the online version, at http://dx.doi.org/10.1016/j.jaap.2015.10.013. cellulose for production of sugars by fast pyrolysis, J. Anal. Appl. Pyrolysis 16
(1989) 127–142.
[27] K. Werner, L. Pommer, M. Broström, Thermal decomposition of
References hemicelluloses, J. Anal. Appl. Pyrolysis (2014) 130–137, http://dx.doi.org/10.
1016/j.jaap.2014.08.013.
[1] D. Mohan, P.H. Pittman Steele, Pyrolysis of wood/biomass for bio-oil: a critical [28] D.K. Shen, S. Gu, A.V. Bridgwater, Study on the pyrolytic behaviour of
review, Energy Fuels 20 (2006) 848–889, http://dx.doi.org/10.1021/ xylan-based hemicellulose using TG-FTIR and Py-GC-FTIR, J. Anal. Appl.
ef0502397. Pyrolysis 87 (2010) 199–206, http://dx.doi.org/10.1016/j.jaap.2009.12.001.
[2] A.J. Ragauskas, C.K. Williams, B.H. Davison, G. Britovsek, J. Cairney, C.A. Eckert, [29] N. Worasuwannarak, T. Sonobe, W. Tanthapanichakoon, Pyrolysis behaviors
et al., The path forward for biofuels and biomaterials, Science 311 (2006) of rice straw, rice husk, and corncob by TG-MS technique, J. Anal. Appl.
484–489, http://dx.doi.org/10.1126/science.1114736. Pyrolysis 78 (2007) 265–271, http://dx.doi.org/10.1016/j.jaap.2006.08.002.
[3] N. Brosse, A. Dufour, X. Meng, Q. Sun, A. Ragauskas, Miscanthus: a [30] Y. Peng, S. Wu, The structural and thermal characteristics of wheat straw
fast-growing crop for biofuels and chemicals production, Biofuels Bioprod. hemicellulose, J. Anal. Appl. Pyrolysis 88 (2010) 134–139, http://dx.doi.org/
Bior. 6 (2012) 580–598, http://dx.doi.org/10.1002/bbb.1353. 10.1016/j.jaap.2010.03.006.
[4] R.J. Evans, T.A. Milne, Molecular characterization of the pyrolysis of biomass. [31] T. Hosoya, H. Kawamoto, S. Saka, Pyrolysis behaviors of wood and its
1. Fundamentals, Energy Fuels 1 (1987) 123–137. constituent polymers at gasification temperature, J. Anal. Appl. Pyrolysis 78
[5] A. Dufour, M. Castro-Diaz, N. Brosse, M. Bouroukba, C. Snape, The origin of (2007) 328–336, http://dx.doi.org/10.1016/j.jaap.2006.08.008.
molecular mobility during biomass pyrolysis as revealed by In situ 1H NMR [32] A. Ohnishi, K. Katō, E. Takagi, Pyrolytic formation of
spectroscopy, ChemSusChem 5 (2012) 1258–1265, http://dx.doi.org/10.1002/ 3-hydroxy-2-penteno-1,5-lactone from xylan, xylo-oligosaccharides, and
cssc.201100442. methyl xylopyranosides, Carbohydr. Res. 58 (1977) 387–395, http://dx.doi.
[6] F.-X. Collard, J. Blin, A review on pyrolysis of biomass constituents: org/10.1016/s0008-6215(00)84365-7.
mechanisms and composition of the products obtained from the conversion [33] G.R. Ponder, G.N. Richards, Thermal synthesis and pyrolysis of a xylan,
of cellulose, hemicelluloses and lignin, Renewable Sustainable Energy Rev. 38 Carbohydr. Res. 218 (1991) 143–155, http://dx.doi.org/10.1016/0008-
(2014) 594–608, http://dx.doi.org/10.1016/j.rser.2014.06.013. 6215(91)84093-t.
[7] J. Lédé, Cellulose pyrolysis kinetics: an historical review on the existence and [34] M.P. Pandey, C.S. Kim, Lignin depolymerization and conversion: a review of
role of intermediate active cellulose, J. Anal. Appl. Pyrolysis (2011) 17–32, thermochemical methods, Chem. Eng. Technol. 34 (2011) 29–41, http://dx.
http://dx.doi.org/10.1016/j.jaap.2011.12.019. doi.org/10.1002/ceat.201000270.
[8] A. Jensen, K. Dam-Johansen, M.A. Wójtowicz, M.A. Serio, TG-FTIR Study of the [35] H. Kawamoto, S. Horigoshi, S. Saka, Pyrolysis reactions of various lignin model
influence of potassium chloride on wheat straw pyrolysis, Energy Fuels 12 dimers, J. Wood Sci. 53 (2007) 168–174, http://dx.doi.org/10.1007/s10086-
(1998) 929–938, http://dx.doi.org/10.1021/ef980008i. 006-0834-z.
[9] E. Jakab, O. Faix, F. Till, Thermal decomposition of milled wood lignins studied [36] E. Jakab, O. Faix, F. Till, T. Székely, Thermogravimetry/mass spectrometry
by thermogravimetry mass spectrometry, J. Anal. Appl. Pyrolysis 40-1 (1997) study of six lignins within the scope of an international round robin test, J.
171–186, http://dx.doi.org/10.1016/s0165-2370(97)00046-6. Anal. Appl. Pyrolysis 35 (1995) 167–179, http://dx.doi.org/10.1016/0165-
[10] M.J.J. Antal, G. Varhegyi, Cellulose pyrolysis kinetics: the current state of 2370(95)00907-7.
knowledge, Ind. Eng. Chem. Res. 34 (1995) 703–717, http://dx.doi.org/10. [37] S. Wang, K. Wang, Q. Liu, Y. Gu, Z. Luo, K. Cen, et al., Comparison of the
1021/ie00042a001. pyrolysis behavior of lignins from different tree species, Biotechnol. Adv. 27
[11] D.J. Nowakowski, J.M. Jones, Uncatalysed and potassium-catalysed pyrolysis (2009) 562–567, 10.1016/j.biotechadv.2009.04.010.
of the cell-wall constituents of biomass and their model compounds, J. Anal. [38] R.K. Sharma, J.B. Wooten, V.L. Baliga, X.H. Lin, W.G. Chan, M.R. Hajaligol,
Appl. Pyrolysis 83 (2008) 12–25, http://dx.doi.org/10.1016/j.jaap.2008.05.007. Characterization of chars from pyrolysis of lignin, Fuel 83 (2004) 1469–1482,
[12] O. Faix, I. Fortmann, J. Bremer, D. Meier, Thermal degradation products of http://dx.doi.org/10.1016/j.fuel.2003.11.015.
wood: A collection of electron-impact (EI) mass spectra of polysaccharide [39] A. Dufour, J. Weng, L. Jia, X. Tang, B. Sirjean, R. Fournet, et al., Revealing the
derived products, Holz Als Roh-Werkst 49 (1991) 299–304, http://dx.doi.org/ chemistry of biomass pyrolysis by means of tunable synchrotron
10.1007/BF02663795. photoionisation-mass spectrometry, RSC Adv. 3 (2013) 4786–4792, http://dx.
[13] A. Fendt, T. Streibel, M. Sklorz, D. Richter, N. Dahmen, R. Zimmermann, doi.org/10.1039/c3ra40486b.
On-line process analysis of biomass flash pyrolysis gases enabled by soft [40] D.K. Shen, S. Gu, K.H. Luo, S.R. Wang, M.X. Fang, The pyrolytic degradation of
photoionization mass spectrometry, Energy Fuels 26 (2012) 701–711, http:// wood-derived lignin from pulping process, Bioresour.Technol. 101 (2010)
dx.doi.org/10.1021/ef2012613. 6136–6146, 10.1016/j.biortech.2010.02.078.
[14] W. Genuit, J.J. Boon, O. Faix, Characterization of beech milled wood lignin by [41] F. Mühlberger, J. Wieser, A. Morozov, A. Ulrich, R. Zimmermann,
pyrolysis-gas chromatography–photoionization mass spectrometry, Anal. Single-photon ionization quadrupole mass spectrometry with an electron
Chem. 59 (1987) 508–513, http://dx.doi.org/10.1021/ac00130a029. beam pumped excimer light source, Anal. Chem. 77 (2005) 2218–2226,
[15] T. Adam, R.R. Baker, R. Zimmermann, Investigation, by single photon http://dx.doi.org/10.1021/ac048319f.
ionisation (SPI)-time-of-flight mass spectrometry (TOFMS), of the effect of [42] E. Butler, G. Devlin, D. Meier, K. McDonnell, Fluidised bed pyrolysis of
different cigarette-lighting devices on the chemical composition of the first lignocellulosic biomasses and comparison of bio-oil and micropyrolyser
cigarette puff, Anal. Bioanal. Chem. 387 (2006) 575–584, http://dx.doi.org/10. pyrolysate by GC/MS-FID, J. Anal. Appl. Pyrolysis 103 (2013) 96–101, http://
1007/s00216-006-0945-9. dx.doi.org/10.1016/j.jaap.2012.10.017.
346 Y. Le Brech et al. / Journal of Analytical and Applied Pyrolysis 117 (2016) 334–346

[43] P.T. Williams, S. Besler, The influence of temperature and heating rate on the [53] R. El Hage, N. Brosse, L. Chrusciel, C. Sanchez, P. Sannigrahi, A. Ragauskas,
slow pyrolysis of biomass, Renew. Energy 7 (1996) 233–250. Characterization of milled wood lignin and ethanol organosolv lignin from
[44] O. Faix, I. Fortmann, J. Bremer, D. Meier, Thermal degradation products of miscanthus, Polym. Degrad. Stab. 94 (2009) 1632–1638, http://dx.doi.org/10.
wood, Holz Als Roh-Werkst 49 (1991) 213–219, http://dx.doi.org/10.1007/ 1016/j.polymdegradstab.2009.07.007.
bf02613278. [54] A. Fendt, R. Geissler, T. Streibel, M. Sklorz, R. Zimmermann, Hyphenation of
[45] O. Faix, D. Meier, I. Fortmann, Thermal degradation products of wood, Holz two simultaneously employed soft photo ionization mass spectrometers with
Als Roh-Werkst 48 (1990) 351–354, http://dx.doi.org/10.1007/bf02639897. thermal analysis of biomass and biochar, Thermochim. Acta 551 (2013)
[46] P.R. Patwardhan, J.A. Satrio, R.C. Brown, B.H. Shanks, Influence of inorganic 155–163, http://dx.doi.org/10.1016/j.tca.2012.10.002.
salts on the primary pyrolysis products of cellulose, Bioresour. Technol. 101 [55] M. Kleen, G. Gellerstedt, Influence of inorganic species on the formation of
(2010) 4646–4655. polysaccharide and lignin degradation products in the analytical pyrolysis of
[47] A. Trendewicz, R. Evans, A. Dutta, R. Sykes, D. Carpenter, R. Braun, Evaluating pulps, J. Anal. Appl. Pyrolysis 35 (1995) 15–41, http://dx.doi.org/10.1016/
the effect of potassium on cellulose pyrolysis reaction kinetics, Biomass 0165-2370(95)00893-j.
Bioenergy 74 (2015) 15–25, http://dx.doi.org/10.1016/j.biombioe.2015.01. [56] M.F. Nonier, N. Vivas, N. Vivas de Gaulejac, C. Absalon, P. Soulié, E. Fouquet,
001. Pyrolysis-gas chromatography/mass spectrometry of Quercus sp. wood, J.
[48] I.-Y. Eom, K.-H. Kim, J.-Y. Kim, S.-M. Lee, H.-M. Yeo, I.-G. Choi, et al., Anal. Appl. Pyrolysis 75 (2006) 181–193, http://dx.doi.org/10.1016/j.jaap.
Characterization of primary thermal degradation features of lignocellulosic 2005.05.006.
biomass after removal of inorganic metals by diverse solvents, Bioresour. [57] F.-X. Collard, J. Blin, A. Bensakhria, J. Valette, Influence of impregnated metal
Technol. 102 (2011) 3437–3444. on the pyrolysis conversion of biomass constituents, J. Anal. Appl. Pyrolysis 95
[49] E.M. Hodgson, D.J. Nowakowski, I. Shield, A. Riche, A.V. Bridgwater, J.C. (2012) 213–226, http://dx.doi.org/10.1016/j.jaap.2012.02.009.
Clifton-Brown, et al., Variation in miscanthus chemical composition and [58] T. Kotake, H. Kawamoto, S. Saka, Mechanisms for the formation of monomers
implications for conversion by pyrolysis and thermo-chemical bio-refining and oligomers during the pyrolysis of a softwood lignin, J. Anal. Appl.
for fuels and chemicals, Bioresour. Technol. 102 (2011) 3411–3418, http://dx. Pyrolysis 105 (2014) 309–316, http://dx.doi.org/10.1016/j.jaap.2013.11.018.
doi.org/10.1016/j.biortech.2010.10.017. [59] T. Kotake, H. Kawamoto, S. Saka, Pyrolysis reactions of coniferyl alcohol as a
[50] M.W. Jarvis, T.J. Haas, B.S. Donohoe, J.W. Daily, K.R. Gaston, W.J. Frederick, model of the primary structure formed during lignin pyrolysis, J. Anal. Appl.
et al., Elucidation of biomass pyrolysis products using a laminar entrained Pyrolysis 104 (2013) 573–584, http://dx.doi.org/10.1016/j.jaap.2013.05.011.
flow reactor and char particle imaging, Energy Fuels 25 (2011) 324–336, [60] T. Kotake, H. Kawamoto, S. Saka, Pyrolytic formation of monomers from
http://dx.doi.org/10.1021/ef100832d. hardwood lignin as studied from the reactivities of the primary products, J.
[51] R. Sykes, B. Kodrzycki, G. Tuskan, K. Foutz, M. Davis, Within tree variability of Anal. Appl. Pyrolysis 113 (2015) 57–64, http://dx.doi.org/10.1016/j.jaap.2014.
lignin composition in populus, Wood Sci. Technol. 42 (2008) 649–661, http:// 09.029.
dx.doi.org/10.1007/s00226-008-0199-0.
[52] G.R. Ponder, G.N. Richards, T.T. Stevenson, Influence of linkage position and
orientation in pyrolysis of polysaccharides: a study of several glucans, J. Anal.
Appl. Pyrolysis 22 (1992) 217–229, http://dx.doi.org/10.1016/0165-
2370(92) 85015-d.

You might also like