Sei sulla pagina 1di 9

Akwasi A.

Boateng
1
e-mail: akwasi.boateng@ars.usda.gov
Charles A. Mullen
Eastern Regional Research Center,
Agricultural Research Service,
U.S. Department of Agriculture,
600 E. Mermaid Lane,
Wyndmoor, PA 19038
Logan Osgood-Jacobs
Peregrine Carlson
Nelson Macken
Department of Engineering,
Swarthmore College,
Swarthmore, PA 19081
Mass Balance, Energy, and
Exergy Analysis of Bio-Oil
Production by Fast Pyrolysis
Mention of trade names or commercial products in this publication is solely for the purpose
of providing specic information and does not imply recommendation or endorsement by
the U.S. Department of Agriculture (USDA). USDA is an equal opportunity provider and
employer. Mass, energy, and exergy balances are analyzed for bio-oil production in a
bench-scale fast pyrolysis system developed by the USDAs Agricultural Research Service
(ARS) for the processing of commodity crops to fuel intermediates. Because mass balance
closure is difcult to achieve due, in part, to the systems small size and complexity a linear
programming optimization model is developed to improve closure of elemental balances
without losing the overall representation of the pyrolysis products. The model results pro-
vide an opportunity to analyze true energy and exergy balances for the system. While
energy comparisons are based on heating values, exergy ows are computed using statisti-
cal relationships and other standard techniques. Comparisons were made for a variety of
biomass feedstocks including energy crops and various byproducts of agriculture and bio-
energy industry. The mass model allows for proper accounting of sources of mass loss and
suggestions for improved system performance. Energy recovery and exergetic efciency
are compared for a variety of pyrolysis product utilization scenarios including use of bio-
char and noncondensable gases as heat sources. Exergetic efciencies show high potential
for energy utilization when all the pyrolysis product streams can be recycled to recuperate
their internal energy. The exergy analysis can be benecial to developing exergetic life
cycle assessments (ELCA) for the fast pyrolysis process as sustainable technology for
advanced biofuels production. [DOI: 10.1115/1.4007659]
Keywords: fast pyrolysis, mass balance, energy, exergy analysis
Introduction
The U.S. Departments of Energy and Agriculture are commit-
ted to achieving the countrys energy security through the devel-
opment of domestic renewable energy and advanced biofuels
which will at the same time create opportunities for the farm
industry. Of the various biomass conversion technologies being
studied, fast pyrolysis has received the farmers attention due to
its small footprint and potential ease of deployment on-farm. A
pilot uidized-bed fast pyrolyzer (maximum input rate is 5 kg/h
biomass, typical rates are 2.5 kg/h) has been developed at the
USDAs ARS and tested for the conversion of a wide variety of
biomass generating useful data such as energy requirements and
product yields that can be used for larger scale design. Details of
system operation and production results for some selected agricul-
tural biomass feedstocks (e.g., switchgrass, alfalfa stems, and bar-
ley) can be found in Refs. [13].
However, like all pilot scale systems, the reactors small size
lead to system uctuations, high heat losses, gas leaks, etc. These,
in turn, contribute to premature condensation of the viscous bio-
oils in the tracks and with the systems shorter run times makes
full recovery of the liquid and solid products problematic. Overall
mass closures typically result in an imbalance of 1540% which
makes true evaluation of system performance difcult. It is imper-
ative to improve mass closure through the use of optimization
modeling.
Optimization models have been used extensively to understand
and improve the performance of engineering systems [4]. A typical
optimization problem was addressed by Szargut and Stanek [5] in
the design of a solar collector. For this application, the objective
was to minimize the depletion of nonrenewable natural exergy
resources in such systems. In another application, Ansari and Tade
[6] developed a nonlinear constrained optimization model which
they applied to control a uid catalytic cracking system. A nonlin-
ear multivariable dynamic control algorithm was developed and
applied to the model. Although these applications involve thermo-
chemical processes and in some instances involving renewable
energy production, to our knowledge, an analytical process model
for fast pyrolysis does not exist. One approach to achieving mass
closure and employed herein, is to adapt a commercially available
linear optimization tool formulated in Excel using the SIMPLEX
Solver. This model has previously been used extensively for a wide
variety of applications [4] similar to the pyrolysis system in this
study including, for example, that described by Papadatos et al. [7],
who used Solver to optimize net revenue in cheese manufacture.
The model developed herein utilizes fundamental relationships and
experimental data to achieve improved elemental balances without
losing the overall representation of the pyrolysis products providing
a unique framework for a fast pyrolysis system and in its applica-
tion to biomass conversion systems. Accurate determination of
mass ows is essential for several applications including improving
system performance in areas such as energy and exergy accounting,
locating components for design improvement, and providing infor-
mation for life cycle analysis (LCA).
Since biomass is heterogeneous in nature, the quality of pro-
duced energy carriers is susceptible to changes hence biomass
conversion technologies are best described by exergy analysis.
Exergy is an expression of the maximum theoretical available
work from a substance if it were to achieve equilibrium with the
environment. The fundamentals of exergy are based on the rst
and second laws of thermodynamics that have been used in
1
Corresponding author.
Contributed by the Advanced Energy Systems Division of ASME for publication
in the JOURNAL OF ENERGY RESOURCES TECHNOLOGY. Manuscript received August 25,
2011; nal manuscript received July 12, 2012; published online October 19, 2012.
Assoc. Editor: Gunnar Tamm.
Journal of Energy Resources Technology DECEMBER 2012, Vol. 134 / 042001-1 Copyright VC
2012 by ASME
Downloaded From: http://asmedigitalcollection.asme.org/ on 05/07/2014 Terms of Use: http://asme.org/terms
process analysis since the middle of the last century. Most recently,
however, it has been a useful tool in life cycle analysis in estimating
depletion of renewable resources. Many energy and exergy analy-
ses have been developed for biomass conversion systems including
Nilsson [8], for example, who compared energy and exergy losses
in the use of straw as fuel to power district heating plants. Exergy
analysis for bioenergy applications has also been performed for bio-
diesel [9], biomass gasication [1013], and bioethanol production
[14]. Most recently, exergy analysis has provided a useful tool in
LCA studies of renewable energy [1520].
In the present study, we report on details of a mass closure
model and apply the results to the conversion of several biomass
commodity crops. Energy and exergy calculations are made and
results discussed. Energy comparisons of input and output mate-
rial ows are based on heating values. Exergy ow relations are
developed and evaluated. The chemical exergy for the biomass,
bio-oil and biochar are computed from statistical regression equa-
tions found in the literature [21]. Chemical exergy of nonconden-
sable gases are computed using standard methods [22]. Energy
recovery and exergetic efciency is computed and compared for
various input and output scenarios.
Methods
Fast Pyrolysis System. The fast pyrolysis process system
modeled comprises a bubbling uidized bed of quartz sand oper-
ated at temperatures in the 400550

C range. Typical biomass


feed rates are 2.5 kg/h. The converted feedstock we have modeled
and optimized in this process include energy crops such as switch-
grass [1] and other agricultural residues including barley straw
[3], alfalfa stems, and oil seed presscakes (e.g., pennycress press-
cake) [23]. The reactor, shown in Fig. 1, is a 7.62 cm (3-in.) diam-
eter uidized bed equipped with two cyclones in series used for
biochar collection. These are followed by four cold water chilled
condensers in series and a bank of three electrostatic precipitators
(ESP), also in series which are employed for pyrolysis-oil collec-
tion. The functions of each of the components have been detailed
by Boateng et al. [1]. Pyrolysis products comprise biochar, bio-
oil, or pyrolysis oil (condensers and ESP catch), and nonconden-
sable gases (NCG) that are typically exhausted out of the system.
Product yields are quantied by weighing the feedstock, bio-oil,
and biochar products at all collection points. The NCG composi-
tion is analyzed by gas chromatography (GC, Agilent MicroGC
3001A). Total NCG produced is determined by the difference
between the uidizing gas, measured by mass ow meter (Alicat
Scientic, Tucson, AZ) and the efuent gas is measured with a
bulk gas ow meter (Metris).
In a typical run such measurements typically account for about
6085% mass balance. This imbalance is normally attributed to
several factors some of which include the trapping of biomass and
biochar in the bubbling sand bed, cyclone inefciencies, prema-
ture cooling of tarry, viscous uids within ow paths. These con-
ditions including short run times and relatively small production
capacity exacerbate difculties in complete recovery of the liquid
product.
Product Characterization. Product characterization is neces-
sary in order to establish any material balance; some of the experi-
mental approaches are provided herein. Pyrolysis-oil water
content is determined by Karl-Fischer (K-F) titration using 3:1
methanol: chloroform as solvent and HYDRANAL Karl-Fischer
Composite 5 (Fluka) as titrant. Elemental analyses (C, H, N, S) of
the feedstocks and products were determined using a Thermo
Flash EA1112 CHNS/O analyzer, by complete combustion of the
material followed by GC quantication of the combustion prod-
ucts. Oxygen is then determined by difference after accounting for
CHNS, water, and ash. Ash is determined as the percentage
remaining after heating a sample in a mufe furnace in air to
650

C for 6 h.
Mass Closure Theory. The mass models developed are based
on the linear programming model approach, a mathematical mod-
eling method that searches for the best solution to the pertinent
equation (objective function) when given a set of linear con-
straints. The models developed herein use a multigoal weighted
method in which any deviations from the given constraints are
assigned specic weights that reect their importance in the prob-
lem. The sum of these deviations that make up the objective func-
tion and which must always be minimized were run through Excel
Solver. The Excel program employs the SIMPLEX algorithm a
standard method for solving linear programming models for the
optimal solution to the optimization model [4].
In our pyrolysis system application, the model uses eleven deci-
sion variables, which correspond to the eleven different outputs of
the pyrolysis process listed in Table 1. Each of these variables rep-
resents a fraction of the corresponding product in the total output
Fig. 1 USDA-ARS bench-scale fast pyrolysis system. (a) Gas preheater, (b) feed hopper,
(c) injection screw, (d) uidized-bed reactor, (e) cyclone, (f) condenser train, (g) electrostatic
preceptor (ESP), and (h) exhaust (NCG to GC).
042001-2 / Vol. 134, DECEMBER 2012 Transactions of the ASME
Downloaded From: http://asmedigitalcollection.asme.org/ on 05/07/2014 Terms of Use: http://asme.org/terms
between (0 and 1). The total amount of each product can be calcu-
lated by multiplying the decision variable by the total ash-free
input mass. All of elemental fractions to be discussed later are cal-
culated on an ash-free basis. Two types of constraints are estab-
lished, i.e., hard constraints that cannot be broken and loose
constraints that can be broken.
The rst two constraints employed in the pyrolysis model are
considered hard constraints, and these can be written as follows:
(1) Equality of molar input and output of each element (C, H,
O, N)
e C; H; O; N :

11
i1
f
i;e
M
e
x
i

f
in;e
M
e
(1)
(2) There must be at least as much of each product as was
measured, i.e.,
i 1; 2; ; 11 : x
i

m
i
m
in
(2)
The next three constraints applied were loose constraints.
The amount that each of these constraints breaks or devi-
ates from the given constraint is what makes up the objec-
tive function. These three loose constraints may be invoked
as follows:
(3) The calculated mass fraction of each NCG must equal
measured mass fraction of each NCG
i 7; 8; ; 10 :
x
i

10
j7
x
j

m
i

10
j7
m
j
(3)
For this to be a linear constraint and to include the devia-
tion term, it must be expressed mathematically as
i 7; 8; ; 10 : x
i

m
i

10
j7
m
j

10
j7
x
j
_ _
s

3;i
s

3;i
0 (4)
(4) The calculated water fraction of bio-oil and biochar must
equal measured water fraction in the bio-oil and biochar, i.e.,
i 2; 4; 6 :
x
i
x
i
x
i1

m
i
m
i
m
i1
(5)
Again, for this to be a linear constraint and to include the
deviation term it must be rewritten as
i 2; 4; 6 : x
i

m
i
m
i
m
i1
x
i
x
i1
s

4;i
s

4;i
0 (6)
The last constraint concerns the biochar. The only possible
deviation is above the expected value because of constraint
2 described above, dening the measured value as the mini-
mum possible solution for all products.
(5) Fraction of calculated biochar out must equal fraction of
measured biochar output
x
5
x
6
s

5

m
5
m
6
m
in
(7)
The objective function for the model may be written as follows:
Y a
3

10
i7
s

3;i
s

3;i
_ _
a
4

4;6
j2
s

4;j
s

4;j
_ _
a
5
s

5
(8)
The objective is to minimize Eq. (8) above. To do so each of the
deviation terms is multiplied by a weight, a. These weights can be
used to indicate an assumed accuracy of measurement. For exam-
ple, a higher assigned weight would be used if the system is
thought to have very little error in the corresponding constraint. In
the current pyrolysis system model, little is known about which
measurements are more accurate; therefore a range of weights is
tested.
The Excel solver is used to solve the system of optimization
equations with the following conditions for convergence: (i)
assume a linear model, (ii) apply a precision of 0.01 (precision of
measurement), a tolerance of 5% (default value, not used in the
analysis), and (iii) a convergence criterion of 0.001 (chosen as one
order less than precision). These results are then transferred into a
table that shows the percent of each product in the total output,
the percent each product contributes to each of the elements, and
the difference between the measured and calculated biochar,
water, and NCG fraction.
Energy Balance Model. For energy computation, energy
streams in and out of the pyrolysis system are considered in the
control volume illustrated in Fig. 2. Energy inputs considered are
the electrical energy for the reactor and the energy of the biomass.
Energy outputs are the energy in the bio-oil, the biochar and the
NCG. As Fig. 2 indicates the system boundary is dened at ambi-
ent conditions. This eliminates heat transfer as an energy contribu-
tor. Other energy sources and outputs are small and considered
negligible. These include the electricity used in the electrostatic
precipitator, the energy of ash produced, the energy given up by
cooling water in the condensers, the input energy of the nitrogen
used as a uidizing medium and the energy of the water produced.
All material streams are evaluated by their heat of combustion
or higher heating value (HHV). Hence, the energy per unit mass is
simply assigned the HHV, i.e.,
E HHV (9)
From here, we dene energy recovery as
Energy recovery Useful energy output = Energy input (10)
Both energy input and output were evaluated using appropriate
heating values. These represent the energy content for the starting
Table 1 Denition of decision variables used in the mass
model
Decision variable Corresponding product
x
1
Pyrolysis oil from ESP
x
2
Water from ESP
x
3
Pyrolysis oil from condenser
x
4
Water from condenser
x
5
Char from cyclones
x
6
Water from char in cyclones
x
7
CO from NCG
x
8
CO
2
from NCG
x
9
H
2
from NCG
x
10
CH
4
from NCG
x
11
NH
3
from NCG
Fig. 2 Energy and exergy ows for the system
Journal of Energy Resources Technology DECEMBER 2012, Vol. 134 / 042001-3
Downloaded From: http://asmedigitalcollection.asme.org/ on 05/07/2014 Terms of Use: http://asme.org/terms
materials and products. Per this denition, energy recovery will
depend on the utilization of energy input and output streams.
Because there are several possible ways that the internal energy of
the pyrolysis-oil byproducts can be recuperated, energy recovery
can be based on several scenarios of product utilization. These
include (i) the target product bio-oil for use as the only energy car-
rier, e.g., for say vehicular application, (ii) if bio-oil and biochar
are both useful energy carriers, and (iii) if bio-oil, biochar, and
NCG are all useful energy carriers which can be recycled as
potential heat source for the endothermic pyrolysis reactions.
These scenarios will be applied to energy recovery results.
Exergy Balance Model. Exergy is an expression of the maxi-
mum theoretical work available from a substance if it were to
achieve equilibrium with the environment. It is used in evaluating
the potential use of energy resources. The reference (environmen-
tal) values used for this study are the generally accepted values of
temperature at 298 K and pressure at 1 atm. For a uid stream per
unit mass, exergy is expressed by invoking the laws of thermody-
namics as
E
x
h h
o
T
o
s s
o

v
2
2
gz E
xch
(11)
where the rst four terms are the thermomechanical exergy and
the last term is the chemical exergy. In the pyrolysis system, (see
Fig. 2), we consider all entering and exiting uid streams as well
as our boundary for heat transfer to be at reference conditions.
Hence the rst four terms of Eq. (11) are zero and therefore the
chemical exergy is the only contribution to uid exergy streams.
In several applications like ours, the concept of exergy is fre-
quently applied to a system as illustrated in Fig. 2. Input exergy
consists of the biomass (including its water content) and the elec-
trical exergy required to heat the reactor. The latter is normally
equivalent to the electrical energy. Like the energy, the exergy
associated with the electrostatic precipitator, the exergy of ash
produced, the input exergy of the nitrogen used as a uidizing me-
dium and the exergy of the water produced are small and ignored.
Our preliminary calculations indicate that they each contribute
less than 1% to the total exergy input or output. Again, since our
system boundary is at the reference environment, exergy associ-
ated with heat transfer and the thermomechanical exergy of the
cooling condenser water is neglected.
Chemical exergy is usually found by formulating the reactions
of a given chemical with the elements in the environment and
nding the maximum theoretical work that could come from this
reaction [22]. However, the chemical exergy of the biomass, bio-
char, and bio-oil are all not as well dened. Therefore, to evaluate
this, statistical methods are applied and regression equations
reported for a large number of organic compounds and fuels [21]
are employed. These expressions (dened by b) show a relation
between the atomic ratios or mass fractions of H/C, O/C, N/C, and
S/C and the chemical exergy. The general form of these exergy
equations (for dry substances with negligible ash and sulfur) is
expressed as
E
xch
b LHV (12)
where LHV is the lower heating value. The equations below
dene b with z indicating mass fraction [19]
b
biomass

1:0412 0:2160
z
H
z
C
0:2499
z
O
z
C
1 0:7884
z
H
z
C
_ _
0:0450
z
N
z
C
1 0:3035
z
O
z
C
(13)
b
biochar
1:0437 0:1896
z
H
z
C
0:0617
z
O
z
C
0:0428
z
N
z
C
(14)
b
bio-oil
1:0401 0:1728
z
H
z
C
0:0432
z
O
z
C
0:2169
z
S
z
C
1 2:0628
z
H
z
C
_ _
(15)
These specic relationships have been used in previous publica-
tions [810]. The chemical exergies of the NCG are found by
standard relationships [22]
E
xch
e
ch
RT
o
lny (16)
The exergy associated with electrical power (considered work) is
equivalent to the electrical energy [22].
From these expressions, the exergetic efciency can be dened
as follows:
W Exergy output = Exergy input (17)
By expanding on Eq. (17), the exergetic efciency may be com-
puted as
W
E
xbio-oil
E
xbiochar
E
xNCG
E
xbiomass
E
electricity
(18)
The exergetic efciency is useful for evaluating environmental
performance and the comprehensive energy efciency of a sys-
tem. In the best system, exergy would be conserved and the
exergetic efciency would be one. Reduced forms of Eq. (18) can
be computed if, for example, the exergy of the biochar and NCG
are considered unimportant.
Results and Discussion
Mass Model. We present results for the biomass commodities
used and described in the methods section, including switchgrass,
pennycress presscake, barley straw, rye grass, and alfalfa stems.
Before we do so, it is important to caution that our results are spe-
cic to the biomass used and it is to evaluate the model in the py-
rolysis system. We do not hypothesize that our results are
consistent with all varieties of the biomass discussed as that will
require use of a larger pool of feedstocks which is impossible in
the current evaluation. Our purpose in this study is only to demon-
strate the usefulness of our mass model, energy, and exergy analy-
sis and not to draw specic conclusions about biomass inputs or
values computed. Since the mass model allows for deviations
from expected values, several solutions are theoretically possible,
i.e., there are different solutions for different weighting factors
assigned in the constraint Eqs. (4), (6), and (7). We selected
results that would give us the maximum predicted bio-oil output.
A comparison with measured values can provide a range of out-
puts for different biomass inputs.
042001-4 / Vol. 134, DECEMBER 2012 Transactions of the ASME
Downloaded From: http://asmedigitalcollection.asme.org/ on 05/07/2014 Terms of Use: http://asme.org/terms
Table 2 and Fig. 3 compare measured and mass model results.
The percentages shown are the mass fraction of output compared
to input on an ash-free basis. The rst three rows of Table 2
(Overall Products) represent the total output. The bio-oil and bio-
char are as is, i.e., as collected with water (but not ash)
included. The lost calculation is the percent of unaccounted mass.
The remaining entries in the table are details for products that cor-
respond to the decision variables in Table 1.
The model results show that the difference between predicted
and measured bio-oil ranges from 10% to 45%. This is not unex-
pected, since the bio-oil may remain in the system and not appear
in the output. For switchgrass and pennycress presscake used
here, the differences between measured and predicted are about
10%. An examination of the detailed outputs allows for predicting
possible locations where the bio-oil unaccounted for is held up,
i.e., either in the condensers or electrostatic precipitators. As
Table 2 shows we predict that the missing bio-oil is situated in the
electrostatic precipitator for switchgrass, barley straw, and alfalfa
stems and in the condenser for rye grass. The results also show
considerable losses in the NCG and, for switchgrass, loss of bio-
char. In a practical application of the model, this could suggest
that improvements need to be made in gas outlet measurement
and to reduce leakages. Also, we note that these particular com-
parisons are made for the model parameters that predict maximum
bio-oil output. A more exhaustive analysis over a range of solu-
tion parameters would yield more denitive results. The discus-
sion here is presented to demonstrate the usefulness of the model
in identifying losses in bio-oil.
The mass model results also provide useful information on how
the elements (C, H, N, and O) that make up the organic portion of
the biomass feedstock are distributed into the products, therefore
allowing for the estimation of the true carbon conversion. Since
carbon from biomass is considered renewable, it is desirable to
retain as much carbon as possible and as low oxygen as possible
in the bio-oil. Bio-oil with less oxygen has a higher energy con-
tent. Hydrogen retention is equally important since it reduces the
hydrogen necessary for post-production upgrading of the bio-oil.
Table 3 provides the distribution of the carbon, oxygen, hydro-
gen and nitrogen from the biomass to the pyrolytic products (pyrol-
ysis oil, biochar, and noncondensable gases) using the predicted
results of the mass balance model. The mass balance model results
revealed that 5360% of the biomass carbon ended up in the bio-oil
while oxygen conversion to bio-oil ranged from 26% for pennycr-
ess presscake up to 57% for switchgrass. Oxygen rejection as water
accounted for between 23% of the biomass oxygen in switchgrass
up to 39% of the oxygen for pennycress presscake. Carbon con-
verted from biomass to biochar has potential to be sequestered, and
therefore, make the fast pyrolysis process overall carbon negative
[24,25]. The mass balance model reveals that between 17% and
28% of the biomass carbon is converted to biochar. Nitrogen distri-
bution was also measured. Only the pennycress presscake and
alfalfa stems had signicant nitrogen content in the feedstocks, and
the distribution of the nitrogen to bio-oil varied from 51% for pen-
nycress presscake to 85% for alfalfa. In each case the biochar con-
tained between 12% and 17% of the biomass nitrogen. Organic
nitrogen in protein can also be converted into the gaseous fraction
during pyrolysis, with NH
3
being the dominant gaseous nitrogen
containing product produced at <600

C [26,27]. The amount in


the NCG (NH
3
) also varied from no nitrogen converted to NH
3
for
alfalfa stems, to 36% pennycress presscake. This wide range is
likely reective of a wider variety of structures containing nitrogen
(proteins) in the feedstocks than there among the oxygenated struc-
tures (carbohydrates and lignin).
Energy. Energy analysis results are illustrated in Tables 4
and 5. As discussed earlier, material energy inputs are evaluated
using HHV. Biomass and the reactor electrical energy are inputs
and bio-oil (dry), biochar (dry), and NCG are energy outputs. A
comparison is made of measured results and those based on ows
predicted by the mass model. The model results used are based on
mass values for maximum bio-oil output as reported above. The
full predicted range of bio-oil output, i.e., measured to maximum
bio-oil, should be useful in evaluating thermal performance.
Table 4 gives quantitative results for input and output. This
table indicates a difference of 1540% between measured and pre-
dicted total energy outputs. If the bio-oil alone is considered as
output, the difference in energy is as large as 50%. This illustrates
the importance of improving mass balance. These computations
can also be used to compare the energy ranking of different bio-
mass inputs. Values are in energy rate (MJ/h). The absolute values
are system dependent, i.e., only applicable to this fast pyrolysis
process. Table 4 indicates switchgrass and pennycress presscake
provided more energy input and barley straw and alfalfa stems the
least in the runs reported. The energy ranking in bio-oil outputs
(using the model results) follows the inputs except for ryegrass
which has the lowest bio-oil output but not the lowest input.
Table 5 reports energy recovery as the ratio between energy
output to energy input as illustrated in Fig. 2 and following the
scenarios discussed earlier. The entrees under Product Output
Table 2 Measured and mass model product yields (wt. %)
Switchgrass Pennycress presscake Barley straw Rye grass Alfalfa stems
Measured Model Measured Model Measured Model Measured Model Measured Model
Overall products
Bio-oil (as is) 62.3 69.5 53.6 59.2 42.70 64.5 46.6 59.5 32.6 60.6
Char (as is) 9.79 15.6 20.7 20.7 13.9 13.9 17.6 17.6 13.2 13.2
NCG 11.6 14.9 10.8 19.3 NA 21.7 7.25 23.0 17.8 26.2
Lost (Measured) 16.3 14.96 28.6 36.4
Detailed products
ESP oil 25.0 32.2 11.6 13.7 13.2 33.5 21.5 21.5 10.6 30.9
ESP water 1.65 1.65 0.85 0.85 0.97 2.47 1.44 1.44 2.84 10.6
Condenser oil 26.3 26.3 26.8 30.36 18.0 18.0 19.9 20.9 15.1 15.1
Condenser water 9.44 9.44 14.4 14.4 10.5 10.5 3.70 15.59 4.03 4.03
Char 9.42 15.07 20.7 20.7 13.7 13.6 17.3 17.3 12.7 12.2
Char water 0.37 0.50 0.00 0.01 0.20 10.53 0.23 0.32 0.48 4.03
CO 6.14 6.14 1.57 1.57 1.97 12.52 1.91 6.06 3.14 4.61
CO
2
4.94 4.94 8.96 8.96 1.18 4.58 5.24 15.5 14.2 20.8
H
2
0.04 0.04 0.01 0.01 0.02 1.29 0.00 1.14 0.02 0.16
CH
4
0.48 3.65 0.27 5.95 0.24 3.32 0.09 0.30 0.47 0.69
NH
3
0.00 0.18 0.00 2.84 0.00 0.00 0.00 0.00 0.00 0.00
Note: NANot measured.
Journal of Energy Resources Technology DECEMBER 2012, Vol. 134 / 042001-5
Downloaded From: http://asmedigitalcollection.asme.org/ on 05/07/2014 Terms of Use: http://asme.org/terms
illustrate product use, i.e., the useful energy output in Eq. (10).
The entrees under products utilized as input illustrate the poten-
tial for recuperation of the energy in the byproducts (recycle).
Under none, we mean energy input is biomass electricity
for the reactor. For the other columns, char, NCG,
char NCG, the electricity is replaced by using the biochar
and/or NCG as heat sources where appropriate. The results are
highly dependent on biomass. The model predicts switchgrass
(6092%) and barley straw (6099%) to have the highest energy
recoveries. It is clear that considering biochar and/or NCG as
products in addition to bio-oil has the potential to produce very
high energy recovery. The use of the biochar and NCG output as
input (recycle) also provides substantial increases in energy recov-
ery. It is important to note that the results for these recycle oppor-
tunities are upper bounds and assume complete energy conversion
is achievable.
Exergy. Table 6 presents a list of exergy production for the
inputs and outputs and the predicted exergetic efciencies. The
row listed as all products useful assumes that the output exergy
as bio-oil, biochar, and NCG as indicated in Eq. (18). This pro-
vides a measure of the efciency for the overall process. The sec-
ond row bio-oil useful considers only the bio-oil exergy as
output. In the third row, we consider the possibility of the biochar
and the NCG exergy to be used to provide the electrical power for
the process. This results in the exergetic efciency dened as the
ratio of bio-oil to biomass exergy.
We can compare various biomass inputs and outputs to deter-
mine their relative value for biomass conversion. Table 6 shows a
wide range of exergy production (MJ/h) values. If we were to base
our system performance on exergy production alone, we would
conclude that switchgrass (43.5 MJ/h) is perhaps a better candidate
for our system as compared to alfalfa stems (22.1 MJ/h). However,
the ranking of bio-oil candidates additionally depends on the con-
version efciencies, discussed more thoroughly below. Generally,
the ranking of bio-oil values follow the input biomass. However,
we see some exceptions in comparing rye grass to barley straw or
alfalfa stems.
One aspect of looking at exergetic efciency is that it can be
used to dene the price one would pay in the fast pyrolysis pro-
cess. In this perspective, it can be said that the most valuable
energy conversion processes would have the highest energy ef-
ciency and the smallest exergy destruction. For this model results,
we nd efciencies are very high for all products useful in all
cases with the exception of rye grass. These results are also illus-
trated in Fig. 4. This suggests that fast pyrolysis is a viable energy
conversion process. Energy efciencies are dramatically reduced
if bio-oil is the only useful product output as shown in the table
entries labeled bio-oil useful. If the energy in the byproducts is
recuperated into the process itself, recycle, then there is a sub-
stantial rise in efciency as exergy is utilized. For example, the
exergetic efciency for production of bio-oil from switchgrass is
60% if only bio-oil is useful, but if the other products can be
recycled and utilized in the process the exergetic efciency of
bio-oil production from switchgrass increases to 71%. As stated
above in the energy discussion, these results for recycle opportu-
nities assume complete energy conversion is possible and must be
considered as upper bounds.
One application of exergy analysis of such systems provides
can be found in life cycle assessment. ELCA has a distinct advant-
age as a tool in evaluating the environmental problem of depletion
of natural resources [16,17,21]. In these analyses, values of exergy
are required for all steps in the production chain being evaluated.
Our methodology can be used to supply the information required
by ELCA for the fast pyrolysis process in the conversion of bio-
mass to biofuel.
Exergy and Energy Compared. Examining model results in
Tables 5 and 6, we can compare the recovery of energy with the
Fig. 3 Comparison of measured and model yields for fast py-
rolysis of selected biomass
042001-6 / Vol. 134, DECEMBER 2012 Transactions of the ASME
Downloaded From: http://asmedigitalcollection.asme.org/ on 05/07/2014 Terms of Use: http://asme.org/terms
exergetic efciency. Comparing exergetic efciency for the sce-
nario where all products useful (Table 6), with the equivalent
scenario in energy recovery where bio-oil char NCG are
product output (Table 5), we see almost identical values for all
biomasses. The same results are found if we make comparisons
for bio-oil useful (Table 6) with Bio-oil (Table 5) and
recycle (Table 6) with corresponding results in Table 5. This
means the percentage of exergy reduction is almost the same as
efciency in energy recovery. Hence, the fast pyrolysis process
we have studied of the commodities selected as feedstock is about
as efcient in converting heat as it is in destruction of the ability
to produce useful work.
Table 4 Energy input and output
Energy input (MJ/HR) Switchgrass Pennycress presscake Barley straw Rye grass Alfalfa stems
Biomass 39.8 30.8 20.4 25.5 18.1
Reactor power 7.66 7.66 7.66 7.66 7.66
Total input 47.5 38.4 28.1 33.2 25.8
Energy output (MJ/HR) Measured Model Measured Model Measured Model Measured Model Measured Model
Bio-oil 24.9 28.5 15.7 18.0 10.1 16.7 10.5 10.7 6.5 13.0
Char 5.53 8.85 8.80 8.80 5.00 4.98 6.78 6.75 7.01 6.73
NCG 2.03 5.93 0.46 5.90 0.45 6.07 0.34 3.24 0.62 1.10
Total output 32.4 43.8 25.0 32.7 15.6 27.8 17.6 20.7 14.1 20.8
Table 5 Energy recovery in products (%)
Products utilized as input
Measured Model
Product output None Char NCG Char NCG None Char NCG Char NCG
Switchgrass Bio-oil 52.3 59.2 54.7 62.3 60.0 71.5 68.5 71.5
Bio-oil char 64.0 66.9 78.6 89.8
Bio-oil NCG 56.6 64.1 72.5 86.4
Bio-oil char NCG 68.3 92.3
Pennycress presscake Bio-oil 40.9 53.1 41.4 53.9 46.9 58.6 55.4 58.6
Bio-oil char 63.8 64.6 69.8 82.5
Bio-oil NCG 42.1 52.6 62.3 77.8
Bio-oil char NCG 65.8 85.2
Barley straw Bio-oil 36.1 44.0 36.7 44.8 59.6 72.5 76.1 82.0
Bio-oil char 53.9 54.8 77.4 98.7
Bio-oil NCG 37.7 45.9 81.2 98.8
Bio-oil char NCG 55.6 99.0
Rye grass Bio-oil 31.7 39.8 32.0 40.4 32.2 40.4 35.7 41.8
Bio-oil char 52.1 52.7 52.5 58.2
Bio-oil NCG 32.7 41.1 42.0 52.7
Bio-oil char NCG 53.2 62.3
Alfalfa stems Bio-oil 25.1 34.5 25.7 35.7 50.3 69.1 52.5 71.5
Bio-oil char 52.3 53.6 76.4 79.8
Bio-oil NCG 27.5 37.8 54.6 74.9
Bio-oil char NCG 54.7 80.7
Table 3 Distribution of biomass elements to pyrolysis products (%)
Switchgrass Pennycress presscake Barley straw Rye grass Alfalfa stems
Product C H O N C H O N C H O N C H O N C H O N
Bio-oil (DB) 59.9 56.5 57.3 41.7 56.9 29.8 26.0 51.2 58.4 36.4 45.6 80.6 53.0 41.4 31.3 58.0 56.6 59.6 32.0 84.7
Char (DB) 25.9 12.2 3.9 29.6 28.3 12.7 12.7 12.6 22.3 7.5 5.3 19.3 31.8 11.6 2.8 42.0 24.8 10.4 0.8 15.3
NCG (DB) 14.1 13.8 15.8 28.5 14.7 31.1 21.7 36.0 19.2 33.2 23.1 0.0 15.1 18.3 32.2 0.0 18.6 5.0 37.7 0.0
Water 17.4 22.8 26.2 39.4 22.7 25.8 28.7 33.6 25.0 29.5
Journal of Energy Resources Technology DECEMBER 2012, Vol. 134 / 042001-7
Downloaded From: http://asmedigitalcollection.asme.org/ on 05/07/2014 Terms of Use: http://asme.org/terms
Conclusions
A linear programming optimization model is developed to
improve mass closure for a bench-scale fast pyrolysis system.
Computational details and results are discussed for switchgrass,
pennycress presscake, barley straw, rye grass, and alfalfa stems.
Mass model results are compared to actual measurements. Sources
of losses and areas for measurement improvement are identied.
Computations of exergy and energy clearly demonstrate the
necessity and advantages of improved mass closure. Comparisons
are made of the various biomass inputs and their outputs for
energy and exergy. Energy recovery and exergetic efciency are
compared for a variety of product utilization scenarios. Substan-
tial benets are indicated when biochar and NCG are used as
inputs (recycle) or usable as outputs. Exergetic efciencies show
high potential for energy utilization when all products are useful
as energy carriers. The exergy analysis presented here can be
applied to supply input to ELCA in future developments of the
fast pyrolysis process for biofuels production.
Acknowledgment
The authors would like to thank Professor Arthur McGarity of
Swarthmore College for his help in developing the mass model.
Nomenclature
E energy, kJ/kg
E
x
exergy, kJ/kg mole
E
x
exergy, kJ/kg
M molecular weight, kg/kg mole

R universal gas constant, kJ/kg mole K


T absolute temperature, K
e standard chemical exergy, kJ/kgmole
f measured mass fraction of element, Eq. (1)
g gravity constant, m
2
/s
h enthalpy, kJ/kg
m mass, kg
P pressure, Pa
s entropy, kJ/kg K
s

deviation above the expected value, Eqs. (4), (6)(8)


s

deviation below the expected value, Eqs. (4), (6), and (8)
T temperature, K
v velocity, m/s
x calculated mass fraction of product in Table 1
Y objective function, Eq. (8)
y partial pressure ratio
z height in exergy, Eq. (11)
z mass fraction in exergy regression relations
a weights in objective function, Eq. (8)
b regression equation for exergy
W exergetic efciency
Subscripts
i,j summation indices
ch chemical
o environmental condition
e element
in input
References
[1] Boateng, A. A., Daugaard, D. E., Goldberg, N., and Hicks, K. B., 2007,
Bench-Scale Fluidized-Bed Fast Pyrolysis of Switchgrass for Bio-Oil
Production, Ind. Eng. Chem. Res., 46, pp. 18911897.
[2] Boateng, A. A., Mullen, C. A., Goldberg, N., and Hicks, K. B., 2008,
Production of Bio-Oil From Alfalfa Stems by Fluidized-Bed Fast Pyrolysis,
Ind. Eng. Chem. Res., 47(12), pp. 41154122.
[3] Mullen, C. A., Boateng, A. A., Hicks, K. B., Goldberg, N., and Moreau, R. A.,
2010, Analysis and Comparsion of Bio-Oil Produced by Fast Pyrolysis From
Three Barley Biomass/Byproduct Streams, Energy Fuels, 24, pp. 699706.
[4] Hamdy, T. A., 2007, Operations Research: An Introduction, Pearson Prentice-
Hall, New Jersey.
[5] Szargut, J., and Stanek, W., 2007, Thermo-Ecological Optimization of a Solar
Collector, Energy, 32, pp. 584590.
[6] Ansari, R. M., and Tade, M. O., 2000, Constrained Nonlinear Multivariable Con-
trol of a Fluid Catalytic Cracking Process, J. Process Control, 10, pp. 539555.
[7] Papadatos, A., Berger, A. M., Pratt, J. E., and Barbano, D. M., 2002, A Nonlin-
ear Programming Optimization Model to Maximize Net Revenue in Cheese
Manufacture, J. Dairy Sci., 85, pp. 27682785.
[8] Nilsson, D., 1997, Energy, Exergy and Emergy Analysis of Using Straw as
Fuel in District Heating Plants, Biomass Bioenergy, 13(12), pp. 6373.
Fig. 4 Comparison of exegetic efciency for fast pyrolysis of var-
ious biomass. In this case, all pyrolysis products (bio-oil, biochar,
noncondensable gases are considered useful products).
Table 6 Exergy analysis
Exergy input (MJ/HR) Switchgrass Pennycress presscake Barley straw Rye grass Alfalfa stems
Biomass 41.7 32.2 21.4 27.0 19.3
Reactor power 7.66 7.66 7.66 7.66 7.66
Total input 49.3 39.9 29.2 34.6 26.9
Exergy output (MJ/HR) Measured Model Measured Model Measured Model Measured Model Measured Model
Bio-oil 25.8 29.6 16.2 18.6 10.7 17.5 10.6 10.8 6.6 13.3
Char 5.66 9.05 9.20 9.20 5.21 5.19 6.97 6.93 7.28 7.00
NCG 1.34 4.92 0.99 5.15 0.39 4.62 0.56 3.45 1.14 1.79
Total output 32.8 43.5 26.4 33.0 16.3 27.4 18.1 21.1 15.0 22.1
Exergetic efciency
All products useful (%) 66.5 88.3 66.2 82.6 56.0 93.8 52.3 61.0 55.7 82.0
Bio-oil useful (%) 52.4 59.9 40.7 46.6 36.9 60.2 30.6 31.1 24.5 49.4
Recycle (%) 62.0 71.0 50.3 57.7 82.1 39.9 34.2 69.0
042001-8 / Vol. 134, DECEMBER 2012 Transactions of the ASME
Downloaded From: http://asmedigitalcollection.asme.org/ on 05/07/2014 Terms of Use: http://asme.org/terms
[9] Hovelius, K., and Hansson, P., 1999, Energy- and Exergy Analysis of Rape
Seed Oil Methyl Ester (RME) Production Under Swedish Conditions, Biomass
Bioenergy, 17, pp. 279290.
[10] Ptasinski, K. J., Prins, M. J., and Pierik, A., 2007, Exergetic Evaluation of Bio-
mass Gasication, Energy, 32, pp. 568574.
[11] Prins, M. J., Ptasinski, K. J., and Janssen, F. J. J. G., 2007, From Coal to Bio-
mass Gasication: Comparison of Thermodynamic Efciency, Energy, 32, pp.
12481259.
[12] Braun, R. J., Hanzon, L. G., and Dean, J. H., 2011, System Analysis of
Thermochemical-Based Bioreneries for Coproduction of Hydrogen and Elec-
tricity, ASME J. Energy Resour. Technol., 133(1), p. 012601.
[13] Srinivas, T., Gupta, A. V. S. S. K. S., and Reddy, B. V., 2009, Thermodynamic
Equilibrium Model and Exergy Analysis of a Biomass Gasier, ASME J.
Energy Resour. Technol., 131(3), p. 031801.
[14] Ojeda, K. A., Sanchez, E. L., Suarez, J., Avila, O., Quintero, V., El-Halwagi,
M., and Kafarov, V., 2011, Application of Computer-Aided Process Engineer-
ing and Exergy Analysis to Evaluate Different Routes of Biofuels Production
From Lignocellulosic Biomass, Ind. Eng. Chem. Res., 50, pp. 27682772.
[15] Szargut, J., Ziebik, A., and Stanek, W., 2002, Depletion of the Non-
Renewable Natural Energy Resources as a Measure of Ecological Cost,
Energy Convers. Manage., 43, pp. 11491163.
[16] Cornelissen, R. L., and Hirs, G. G., 2002, The Value of Exergetic Life Cycle
Assessment Besides the LCA, Energy Convers. Manage., 43, pp. 14171424.
[17] Dewulf, J., Van Langenhove, H., Muys, B., Bruers, S., Bakshi, B. R., Grubb, G. F.,
Paulus, D. M., and Sciubba, E., 2008, Exergy: Its Potential and Limitations in Envi-
ronmental Science and Technology, Environ. Sci. Technol., 42(7), pp. 22212232.
[18] Dewulf, J., Van Langenhove, H., and Van De Velde, B., 2005, Exergy-Based
Efciency and Renewable Assessment of Biofuel Production, Environ. Sci.
Technol., 39(10), pp. 38783882.
[19] Rubio Rodriguez, M. A., De Ruyck, J., Roque Diaz, P., Verma, V. K., and
Bram, S., 2011, An LCA Based Indicator for Evaluation of Alternative Energy
Routes, Appl. Energy, 88, pp. 630635.
[20] Talens Peiro, L., Lombardi, L., Villalba Mendez, G., and Gabarrell i Durany,
X., 2010, Life Cycle Assessment (LCA) and Exergetic Life Cycle Assessment
(ELCA) of the Production of Biodiesel From Used Cooking Oil (UCO),
Energy, 35, pp. 889893.
[21] Szargut, J., Morris, D. R., and Steward, F. R., 1988, Exergy Analysis of
Thermal, Chemical, and Metallurgical Processes, Hemisphere Publishing
Corporation, New York.
[22] Moran, M. J., and Shapiro, H. N., 2000, Fundamentals of Engineering Thermo-
dynamics, John Wiley & Sons, Inc., New York.
[23] Boateng, A. A., Mullen, C. A., and Goldberg, N. M., 2010, Producing Stable
Pyrolysis Liquids From the Oil-Seed Presscakes of Mustard Family Plants: Pen-
nycress (Thlaspi arvense L.) and Camelina (Camelina sativa), Energy Fuels,
24, pp. 66246632.
[24] Laird, D. A., 2008, The Charcoal Vision: A Win-Win-Win Scenario for Simul-
taneously Producing Bioenergy, Permanently Sequestering Carbon, While
Improving Soil and Water Quality, Agron. J., 100, pp. 178181.
[25] Boateng, A. A., Mullen, C. A., Goldberg, N. M., Devine, T. E., Lima, I. M., and
McMurtrey, J. E., 2010, Sustainable Production of Bioenergy and Biochar
From the Straw of High Biomass Soybean Lines via Fast Pyrolysis, Environ.
Prog. Sustainable Energy, 29, pp. 175183.
[26] Yan, S., Chen, X., Li, W., Liu, H., and Wang, F., 2011, Nitrogen Conversion
Under Rapid Pyrolysis of Two Types of Aquatic Biomass and Corresponding
Blends With Coal, Bioresour. Technol., 102, pp. 1012410130.
[27] Yaun, S., Zhou, Z., Li, J., Chen, X., and Wang, F., 2010, HCN and NH
3
Released From Biomass and Soybean Cake Under Rapid Pyrolysis, Energy
Fuels, 24, pp. 61666171.
Journal of Energy Resources Technology DECEMBER 2012, Vol. 134 / 042001-9
Downloaded From: http://asmedigitalcollection.asme.org/ on 05/07/2014 Terms of Use: http://asme.org/terms

Potrebbero piacerti anche