You are on page 1of 6

Thermal Pretreatment of

Lignocellulosic Biomass
Wei Yan, Tapas C. Acharjee, Charles J. Coronella, and Victor R. Vasquez
Chemical and Metallurgical Engineering Department, University of Nevada, MS170 Reno, NV 89557; coronella@unr.edu
(for correspondence)
Published online 5 August 2009 in Wiley InterScience (www.interscience.wiley.com). DOI 10.1002/ep.10385

Torrefaction is a process to convert diverse lignocellulosic biomass feedstocks into an energy dense homogeneous solid, a pretreatment for subsequent thermochemical conversion. Loblolly pine was treated by
wet torrefaction (hot compressed water, 2002608C)
and dry torrefaction (nitrogen, 2503008C), with
mass yield of solid product ranging between 57 and
89%, and energy densication to 108136% of the
original feedstock. The solid product has been characterized, including proximate analysis, ber analysis,
ultimate analysis, and equilibrium moisture. In both
dry and wet torrefaction, increasing temperature
results in decreased mass yield and increased energy
densication, and results in a solid with increased
carbon content, decreased oxygen content, and
decreased volatiles. The biomass is transformed into a
fuel similar to a low-rank coal. Generally, the wet torrefaction process produces a solid with greater energy
density than dry torrefaction, with the same mass
yield. The ber analysis indicates that hemicellulose
is quickly removed during wet torrefaction, and the
solid product contains substantial quantities of aqueous soluble compounds. The equilibrium moisture
content of solids produced by both processes is somewhat decreased from that of the biomass feedstock,
indicating a hydrophobic solid suitable for storage and
transportation. 2009 American Institute of Chemical
Engineers Environ Prog, 28: 435440, 2009

Keywords: biomass pretreatment, torrefaction, lignocellulosic biomass, thermochemical conversion


INTRODUCTION

Thermochemical conversion is a promising route


to convert lignocellulosic biomass to fuel, chemicals,
and renewable power [1, 2]. However, both pyrolysis
and gasication of raw biomass feedstock have significant, but distinct, technical barriers that must be
2009 American Institute of Chemical Engineers

eliminated for widespread successful commercialization [2, 3]. In both cases, feedstock handling is
complex due to the diverse nature of important feedstocks, including commercial timber, agricultural residues as diverse as rice hulls, corn stover and wheat
chaff, and energy crops such as switch grass. Seasonal availability and low forest density make feedstock logistics complex and expensive. To solve these
problems, there is a need for a process to homogenize the feedstocks, while simultaneously producing
a stable, energy-dense, solid fuel. Two distinct thermal pretreatment technologies are investigated in this
study: wet torrefaction and dry torrefaction.
In the process we call wet torrefaction, also
known as hydrothermal pretreatment, biomass is
treated in hot compressed water, resulting in three
products, including gases, aqueous chemicals, and a
solid fuel [46]. The temperature is in the range of
2002608C, and the pressures are up to 700 psi. The
solid product contains about 5590% of the mass and
8095% of the fuel value of the original feedstock.
The gas product is about 10% by mass of the feedstock, and the aqueous chemicals, primarily sugars,
make up the balance [79]. On the other hand, dry
torrefaction, sometimes called low-temperature pyrolysis, is a process in which the biomass is heated in
an inert gas environment at temperatures ranging
from 200 to 3008C [3, 1012]. The dry torrefaction
reaction has two products, a solid and a gas. The
solid has about 6080% of the mass, and approximately 7090% of the fuel value of the original biomass [1012].
Both torrefaction processes exhibit some common
features. The solid product has a higher energy density relative to the starting biomass feedstock in both
cases. Both solid products are easily friable and have
an aroma similar to wood char. The reaction mechanisms of both torrefaction processes are poorly understood. The objective of this work is to investigate how

Environmental Progress & Sustainable Energy (Vol.28, No.3) DOI 10.1002/ep

October 2009 435

both torrefaction processes affect product distribution


and to assess the quality of solid products for thermochemical conversion. Characterization of solid products is performed, which includes proximate analysis,
ultimate analysis, higher heating value, ber analysis,
and equilibrium moisture content. A comparative
study is made to elucidate the capability of both thermal pretreatment technologies to homogenize lignocellulosic biomass and produce the immediate solid
fuel for thermochemical conversion.

MATERIALS AND METHODS

Biomass and Chemicals


Loblolly pine (Alabama, USA) was acquired as a
typical lignocellulosic biomass for torrefaction. The
solutions and lter bags for ber analysis were purchased from ANKOM Technology Inc (Macedon,
NY).

Wet Torrefaction
Wet torrefaction of loblolly pine was performed in
a 100-mL Parr bench-top reactor (Moline, IL) at temperatures ranging from 200 to 2608C. The temperature of the reactor was controlled using PID controller. The reactor pressure was not controlled but indicated by a pressure gauge and ranged from 200 to
700 psi, approximately in accord with the water
vapor pressure. For each run, a mixture of loblolly
pine and water (water:biomass 5 5:1, w/w) was
loaded into the reaction vessel. The biomass sample
size was about 2 g. Prior to the reaction, the biomass
was stirred manually to ensure complete wetting.
Nitrogen was passed through the reactor for 10 min
to purge oxygen out of the reactor. The reactor was
heated up in 1530 min and maintained at the
desired temperature for a desired period of time
(5 min). Then, the reactor was cooled to room temperature rapidly by immersion in an ice bath. The gas
sample was released. The liquid sample and solid
sample were separated and collected for further analysis.

Dry Torrefaction
Dry torrefaction of loblolly pine was carried out in
the 100-mL Parr reactor described above, but tted
with a glass sample holder to separate the biomass
from the hot reactor walls. The pretreatment was performed with a nitrogen ow rate of 10 mL (STP)/min.
The glass sample loader with biomass particles
(5 g) were placed in the center of reactor. The torrefaction temperature was between 250 and 3008C
and was controlled using a PID controller. The temperature program consists of a dynamic heating period following by an isothermal period. After a
desired isothermal reaction period (80 min), the
sample was cooled to room temperature with continuous nitrogen ow through the reactor. The solid
sample was collected for further analysis.
436 October 2009

Analytical Methods
Proximate Analysis

Solid samples were analyzed with Perkin Elmer


TGA-7 Thermogravimetric Analyzer (Waltham, MA) to
determine the contents of moisture, volatile matter,
and xed carbon [13]. Thermogravimetric analysis
was carried out under a nitrogen purge at a constant
rate of 50 mL/min to prevent oxidation of samples.
Samples were rst heated up to 1058C at the rate of
108C/min, maintained at 1058C for 10 min, then jump
to 9008C at the rate of 508C/min and hold for 60 min.
Mass evolved at 1058C is said to be moisture. Mass
evolved at temperatures between 105 and 9008C
belongs to volatiles. All mass remaining after heating
to 9008C consists of xed carbon and ash. Ash content of the pretreated biomass was measured separately using a standard method for ash in biomass
[14]. The solid sample was rst dried at 1058C for 24
h and brought to constant weight by igniting at 5758C
for 5 h in a mufe furnace.
Ultimate Analysis

Solid samples were dried in an oven at 1058C for


24 h. Ultimate analysis of both biomass feedstock and
pretreated biomass was performed with a FlashEA
1112 (Pittsburgh, PA) elemental analyzer for full
determination of C, H, N, S, and O [15].
Higher Heating Value

The higher heating value (HHV) of solid samples


was measured in a Parr 1241 adiabatic oxygen bomb
calorimeter (Moline, IL) tted with continuous temperature recording [16]. The sample (0.20.5 g) was
dried at 1058C in the oven for 24 h prior to analysis.
Fiber Analysis

A modied Van Soest method using the ANKOM


A200 Filter Bag Technique (FBT) was used to determine the contents of hemicellulose, cellulose, lignin,
and aqueous soluble compounds in solid samples
[17, 18]. The solid samples are rst crushed and
sieved to the desired particle size (2065 mesh) and
were dried at 1058C for 24 h prior to the analysis.
The contents of hemicellulose, cellulose, lignin, and
solubles were calculated from the difference of neutral detergent ber (NDF), acid detergent ber (ADF),
acid detergent liquid (ADL), and ash.
Equilibrium Moisture Content

Previous literature has suggested that torrefaction


solid products are relatively hydrophobic [12]. To
evaluate that report, the equilibrium moisture content
(EMC) was measured at 308C by the static desiccator
technique. The solid sample is exposed to an environment with constant humidity and temperature
over a long period of time, until the moisture in the
solid reaches an equilibrium value. The humidity in
the chamber is maintained at a constant value by
keeping the air in equilibrium with an aqueous solution that is saturated with a particular salt [19].
Two salt solutions were used, with relative humidities
11.3 and 83.6% for lithium chloride and potassium

Environmental Progress & Sustainable Energy (Vol.28, No.3) DOI 10.1002/ep

Table 1. Thermal pretreatment of loblolly pine.

Pretreatment
Wet torrefaction
Dry torrefaction

Temperature
(8C)

Mass
yield (%)

HHV
(cal/g)

Energy
densication ratio

Energy
yield (%)

200
230
260
250
275
300

88.7
70.6
57.0
83.8
74.2
60.5

5043.3
5276.8
6342.5
5005.4
5207.2
5627.4

1.08
1.13
1.36
1.07
1.12
1.21

95.8
79.8
77.5
89.7
83.1
73.2

The HHV of dried loblolly pine is 4669.8 cal/g.

Table 2. Ultimate analysis of biomass feedstock and thermal pretreated solid products.

Pretreatment
Biomass
Wet torrefaction
Dry torrefaction

Temperature (8C)

200
230
260
250
275
300

C (%)
50.25
54.72
56.05
72.07
50.73
52.27
54.81

chloride, respectively. The solid sample (3 g) was


dried at 1058C for 24 h and then immediately transferred into the chamber for 811 days to reach the
equilibrium. Equilibrium was determined when the
sample mass was constant for three consecutive days.
RESULTS AND DISCUSSION

Wet and dry torrefaction of loblolly pine were performed at various temperatures. All thermal pretreatments and analytical experiments were performed at
least twice, and mean values are reported. Mass yield,
energy densication ratio, and energy yield are three
important measures in this study, which are dened
as:
Mass yield

Mass of dried pretreated solid


3 100%
Mass of dried biomass

Energy densification ratio


HHV of dried pretreated solid

HHV of dried biomass


Energy yield mass yield 3 energy densification ratio
As summarized in Table 1, the results show that
reaction temperature is a signicant variable in both
pretreatment processes. In both cases, increasing temperature causes a reduction in both mass and energy
yields, while energy densication increases. Over the
temperature ranges studied, with similar mass yields,

H (%)
5.97
6.03
5.94
4.90
6.21
6.13
5.94

N (%)
0
0.14
0.09
0.16
0.12
0.15
0.14

S (%)
0
0
0
0
0
0
0

O (%)
43.34
39.11
37.92
22.89
42.94
41.45
39.11

wet torrefaction yields a higher energy densication,


and hence higher energy yield, than dry torrefaction.
For example, wet torrefaction at 2608C has a mass
yield of 57.0% and an energy yield of 77.5%, while
the mass yield of torrefaction at 3008C is 60.5% and
the energy yield is 73.2%. Apparently, wet torrefaction is more successful at energy densication.
Table 2 shows ultimate analysis of biomass feedstock and thermal pretreated solids. Both torrefaction
processes produce a solid fuel with increased carbon
content and decreased oxygen content. This is consistent with an increase in energy density of thermal
pretreated solids. The change in carbon and oxygen
content is signicant, but less dramatic in the case of
dry torrefaction. Wet torrefaction can signicantly
increase carbon content from 50 up to 72%, and
decrease oxygen content from 43 down to 23%.
Proximate analysis is commonly used to establish
the rank of coals [13] and other solid fuels for suitability of combustion, pyrolysis, and gasication.
Table 3 shows the proximate analysis of loblolly pine
and torreed loblolly pine. The proximate analysis of
raw feedstock is similar to the proximate analysis of
biomass wet torreed at low temperatures (200 and
2308C), and it is also similar to that of biomass dry
torreed at low temperatures (250 and 2758C). However, when processed at higher temperatures (2608C
in the case of wet torrefaction, 3008C in the case of
dry torrefaction), the pretreated biomass exhibits
decreased volatiles and increased xed carbon content. This result is consistent with trends reported in
the literature [20] correlating increased heating value
with increasing xed carbon content. The proximate
and ultimate analyses suggest that torrefaction converts biomass from a low energy-density feedstock

Environmental Progress & Sustainable Energy (Vol.28, No.3) DOI 10.1002/ep

October 2009 437

Table 3. Proximate analysis of biomass feedstock and thermal pretreated solid products.

Pretreatment
Biomass
Wet torrefaction
Dry torrefaction

Temperature (8C)

Moisture (%)

Volatiles (%)

Fixed carbon (%)

Ash (%)

200
230
260
250
275
300

3.6
1.3
1.3
1.3
0
0
0

83.7
87.1
83.8
73.2
87.7
83.0
82.3

12.3
12.4
15.8
26.3
11.8
16.4
17.0

0.4
0.5
0.4
0.5
0.5
0.6
0.7

Table 4. Fiber analysis of biomass feedstock and thermal pretreated solid products; conversion of individual
ber components, relative to initial starting feedstock.

Pretreatment

Temperature
(8C)

Hemicellulose
(%)

Cellulose
(%)

Lignin
(%)

Aqueous
solubles (%)

A. Fiber analysis of biomass feedstock and thermal pretreated solid products


Biomass

11.9
54.0
25.0
8.7
Wet torrefaction
200
0.4
47.4
27.8
24.3
230
0
44.1
30.2
25.3
260
0
33.9
33.8
31.8
Dry torrefaction
250
9.2
51.0
34.1
5.2
275
6.5
52.7
36.7
3.5
300
2.3
32.7
62.1
2.2
B. Conversion of individual ber components, relative to initial starting feedstock
Wet torrefaction
200
97.0
22.1
1.4
2147.7
230
100.0
42.3
14.7
2105.3
260
100.0
64.2
22.9
2108.3
Dry torrefaction
250
35.2
20.9
214.3
49.9
275
59.5
27.6
28.9
70.1
300
88.3
63.4
250.3
84.7

Ash
(%)
0.4
0.5
0.4
0.5
0.5
0.6
0.7

A negative conversion indicates a ber compound that was apparently produced by torrefaction.

into a high energy-density feedstock, which may


have favorable properties for application as a fuel for
thermal conversion processes.
Fiber analysis is used to show the reactions of the
primary components of lignocellulosic biomass,
which may help elucidate the reaction mechanism in
both pretreatments. Table 4a, as described above the
distribution of hemicellulose, cellulose, lignin, aqueous soluble compounds, and ash of the biomass
feedstock and pretreated solids. If each ber fraction
is considered as a separate component, it is possible
to determine the conversion of each component by
torrefaction with the aid of the mass yield shown in
Table 1. Table 4b, as described above the conversion
of each of hemicellulose, cellulose, lignin, and soluble
compounds using this approach. A negative conversion indicates that more of that compound is present
after torrefaction than was present in the raw biomass.
Inspection of Table 4 indicates several noteworthy
trends. First, with respect to wet torrefaction, the
hemicellulose is nearly completely removed at 2008C,
and it is completely reacted at 230 and 2608C. Hemicellulose is clearly quite reactive in hot water. Also
438 October 2009

note that both cellulose and lignin are partially


reacted by hot water, and the extent of reaction
increases with temperature. Most interesting is the
nding that aqueous soluble compounds are apparently produced by the wet torrefaction reaction. This
can be explained by considering the products of the
degradation of hemicellulose and cellulose, assumed
to consist primarily of sugars. The remaining solid
product is very porous, and water soluble products
are apparently deposited onto the pores of the solid,
leaving behind signicant quantities of these compounds. Apparently, the solvents used in the ber
analysis are strong enough to remove these compounds. A similar conclusion was reached in a prior
study [6].
Next consider the ber analysis of solid product
produced by dry torrefaction, also shown in Table 4.
Hemicellulose and aqueous soluble compounds are
quite reactive, and are reacted at rates increasing with
temperature. Cellulose exhibits similar, albeit slower,
behavior. Dry torrefaction apparently produces compounds that behave in a manner identical to that of
lignin in the NDF/ADF/ADL ber analysis, especially

Environmental Progress & Sustainable Energy (Vol.28, No.3) DOI 10.1002/ep

Table 5. Equilibrium moisture content of biomass feedstock and thermal pretreated solid products.

Pretreatment
Biomass
Wet torrefaction
Dry torrefaction

Temperature
(8C)

EMC (%),
RH 5 11.3%

EMC (%),
RH 5 83.6%

200
230
260
250
275
300

3.5
1.8
0.9
0.4
2.3
2.2
2.3

15.6
12.8
8.2
5.3
10.4
8.7
8.7

RH, relative humidity. All EMC measurements were done at 308C.

at 3008C. This was unexpected, but might be


explained in consideration of the chemistry of the
torrefaction reaction. Previous work [21] described
low temperature pyrolysis of cellulose as a scission
reaction with signicant aromatization and cross-linking, producing a solid which has characteristics similar to that of lignin. At 250 and 2758C, generation of
lignin-like compounds was quite limited, but at
3008C, signicant lignin was generated; this is seen as
consistent with the aromatization reactions described
above.
Equilibrium moisture content (EMC) can be used
as an indicator of hydrophobicity of a solid. For
lignocellulosic biomass, moisture can be absorbed
into the cell walls and hydrogen-bonded to the
hydroxyl groups of the cell wall components [22].
Breakdown of these hydroxyl groups with thermal
pretreatments results in lower EMC and the solid is
said to become more hydrophobic. A solid with
reduced EMC can be stored stably over time, with
low risk of biological deterioration. Also, transportation of stored hydrophobic solids is less expensive,
because there will less moisture to transport along
with the biomass. Table 5 clearly shows that pretreated solids become more hydrophobic in comparison with biomass feedstock. The EMC of pretreated
biomass decreases with increasing pretreatment temperature, and wet torrefaction produces a solid more
hydrophobic than that produced by dry torrefaction.
At RH 5 84%, EMC of torreed biomass decreases
from 16% for raw biomass to 5% (wet process) and
to 9% (dry process).

biomass has been transformed into a fuel with properties resembling low rank coal. The reduced EMC
indicates that the pretreated solid is more hydrophobic and can be easily stored to accommodate seasonal availability. Although both thermal pretreatments show some common features, the reaction
mechanism is quite different as reected by the ber
analysis. In wet torrefaction, hemicellulose is very reactive even at 2008C, while both cellulose and lignin
can be dissolved partially at higher temperatures.
Both cellulose and hemicellulose are somewhat reactive in dry torrefaction, producing a solid that resists
dissolution, similar to lignin. The solid product produced by dry torrefaction is not at all soluble by ber
analysis measurements.
This study reports analysis on initial scoping
experiments; there is no attempt to optimize the conditions of torrefaction. In particular, the optimal reaction time and temperature have not been identied,
nor have reaction kinetics. Pretreated biomass presumably will be pelletized for feeding to a thermochemical conversion process, and there is no attempt
here to describe the solid characteristics for required
handling (friability, morphology, etc.). To establish
techo-economic feasibility, mass and energy balance
of both torrefaction processes and the subsequent
thermal conversion process (e.g. gasication or fast
pyrolysis) must be investigated as well. We anticipate
further reports from our lab and from other labs on
these issues.

ACKNOWLEDGMENTS
CONCLUSION

Wet and dry torrefaction are promising methods to


convert diverse lignocellulosic biomass feedstocks to
a homogeneous solid fuel for subsequent thermochemical conversion. In both thermal pretreatments,
reaction temperature signicantly affects the product
distribution and energy densication, and character
of solid product. The energy density of pretreated
loblolly pine can be increased by 736%, depending
on the pretreatment conditions. Torreed biomass
has increased xed carbon (proximate analysis) and
atomic carbon (ultimate analysis), indicating that the

We gratefully acknowledge nancial support from


the US Department of Energy (Award Number: DEFG36-01GO11082). The authors acknowledge meaningful conversations with project partners, including
Larry Felix of the Gas Technology Institute (GTI),
Kent Hoekman of the Desert Research Institute (DRI),
and Craig Einfeldt of Changing World Technologies
(CWT). The authors gratefully acknowledge assistance of Jeremey Riggle of DRI for his help in conducting the ultimate analyses, and assistance of Jason
Hastings at UNR who conducted most of the torrefaction experiments.

Environmental Progress & Sustainable Energy (Vol.28, No.3) DOI 10.1002/ep

October 2009 439

LITERATURE CITED

1. Tester, J.W. (2005). Sustainable energy, Cambridge, Massachusetts: MIT Press.


2. U.S. Department of Energy. (2008). Biomass multiyear program, Ofce of the Biomass Program,
Energy Efciency and Renewable Energy.
3. Prins, M.J., Ptasinski, K.J., & Janssen, J.J.G.F.
(2006). More efcient biomass gasication via torrefaction, Energy, 31, 34583470.
4. Ando, H., Sakadi, T., Kokusho, T., Shibata, M.,
Uemura, Y., & Hatate, Y. (2000). Decomposition
behavior of plant biomass in hot-compressed
water, Industrial and Engineering Chemistry
Research, 39, 36883693.
5. Sasaki, M., Adschiri, T., & Arai, K. (2003). Fractionation of sugarcane bagasse by hydrothermal
treatment, Bioresource Technology, 86, 301304.
6. Yan, W., Coronella, C.J., & Vasquez, V.R. Wet torrefaction of lignocellulosic biomass, Biomass Bioenergy, Submitted for publication.
7. Sun, Y. & Cheng, J. (1992). Hydrolysis of lignocellulosic material from ethanol production, Bioresource Technology, 39, 107115.
8. Bobleter, O. (1994). Hydrothermal degradation of
polymers derived from plants, Progress in Polymer Science, 19, 797841.
9. Pertersen, M., Larsen, J., & Thomsen, H.M.
(2009). Optimization of hydrothermal pretreatment of wheat straw for production of bioethanol
at low water consumption without addition of
chemicals, Biomass Bioenergy, 33, 834840.
10. Prins, J.M., Ptasinski, J.K., & Janssen, J.J.G.F. (2006).
Torrefaction of wood. I. Weight loss kinetics, Journal of Analytical and Applied Pyrolysis, 77, 2834.
11. Prins, J.M., Ptasinski, J.K., & Janssen J.J.G.F.
(2006). Torrefaction of wood. II. Analysis of
products, Journal of Analytical and Applied Pyrolysis, 77, 3540.

440 October 2009

12. Bergman, P.C.A., Boersma, A.R., Zwart, R.W.R., &


Kiel, J.H.A. (2005). Torrefaction for biomass coring in existing coal-red power stations (Biocoal), Energy Research Centre of the Netherlands,
ECN-C-05-013.
13. Donahue, J.C. & Rais, A.E. (2009). Proximate
analysis of coal, Journal of Chemical Education,
86, 222224.
14. Ehrman, T. (1994). Standard method for ash in
biomass, NREL Laboratory Analytical Procedure
LAP-005.
15. ASTM D3176-89.(2002). Standard practice for ultimate analysis of coal and coke, West Conshohocken, PA: ASTM International.
16. Operating instructions for the 1241 oxygen bomb
calorimeter, Parr Instruments, No. 203M, pp. 1
13.
17. Goering, H.K. & Van Soest, P.J. (1970). Forage
ber analysis, USDA Agriculture Handbook, 379,
19.
18. ANKOM. (1995). Acid detergent and neutral detergent ber using ANKOMs ber analyzer F200,
Fairport, NY: Ankom Technology Corporation.
19. Mehta, S. & Sing, A. (2006). Adsorption isotherms
for red chili, European Food Research Technology, 223, 849852.
20. Ayhan, D. (2003). Relationships between heating
value and lignin, xed carbon and volatile material contents of shells from biomass products,
Energy Source, 25, 62935.
21. Agrawal, K.R. (1988). Kinetics of reactions
involved in pyrolysis of cellulose. I. The three
reaction model, Canadian Journal of Chemical
Engineering, 66, 403412.
22. Andersson, M. & Tillman, A.M. (1989). Acetylation of jute: effects on strength, rot resistance and
hydrophobicity, Journal of Applied Polymer Science, 37, 34373447.

Environmental Progress & Sustainable Energy (Vol.28, No.3) DOI 10.1002/ep

You might also like