You are on page 1of 57

Critical Review

Pyrolysis for biochar purposes: a review to establish


current knowledge gaps and research needs.
Joan Josep Many
Environ. Sci. Technol., Just Accepted Manuscript DOI: 10.1021/es301029g Publication Date (Web): 09 Jul 2012
Downloaded from http://pubs.acs.org on July 10, 2012

Just Accepted
Just Accepted manuscripts have been peer-reviewed and accepted for publication. They are posted
online prior to technical editing, formatting for publication and author proofing. The American Chemical
Society provides Just Accepted as a free service to the research community to expedite the
dissemination of scientific material as soon as possible after acceptance. Just Accepted manuscripts
appear in full in PDF format accompanied by an HTML abstract. Just Accepted manuscripts have been
fully peer reviewed, but should not be considered the official version of record. They are accessible to all
readers and citable by the Digital Object Identifier (DOI). Just Accepted is an optional service offered
to authors. Therefore, the Just Accepted Web site may not include all articles that will be published
in the journal. After a manuscript is technically edited and formatted, it will be removed from the Just
Accepted Web site and published as an ASAP article. Note that technical editing may introduce minor
changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers
and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors
or consequences arising from the use of information contained in these Just Accepted manuscripts.

Environmental Science & Technology is published by the American Chemical


Society. 1155 Sixteenth Street N.W., Washington, DC 20036
Published by American Chemical Society. Copyright American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the
course of their duties.

Page 1 of 56

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Environmental Science & Technology

Pyrolysis for biochar purposes: a review to establish


current knowledge gaps and research needs
Joan J. Many
Thermo-chemical Processes Group (GPT), Aragn Institute of Engineering Research (I3A),
University of Zaragoza, Technological College of Huesca, crta. Cuarte s/n, E-22071 Spain.
E-mail address: joanjoma@unizar.es

ACS Paragon Plus Environment

Environmental Science & Technology

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Page 2 of 56

ABSTRACT
According to the International Biochar Initiative (IBI), biochar is a charcoal which can be applied to
soil for both agricultural and environmental gains. Biochar technology seems to have a very
promising future. Nevertheless, the further development of this technology requires continuing
research. The present paper provides an updated review on two subjects: the available alternatives to
produce biochar from a biomass feedstock and the effect of biochar addition to agricultural soils on
soil properties and fertility. A high number of previous studies have highlighted the benefit of using
biochar in terms of mitigating global warning (through carbon sequestration) and as a strategy to
manage soil processes and functions. Nevertheless, the relationship between biochar properties
(mainly physical properties and chemical functionalities on surface) and its applicability as a soil
amendment is still unclear and does not allow the establishment of the appropriate process
conditions to produce a biochar with desired characteristics. For this reason, it is highlighted the
need of enhancing the collaboration among researchers working in different fields of study:
production and characterization of biochar on one hand, and on the other, measurement of both
environmental and agronomical benefits linked to the addition of biochar to agricultural soils. In this
sense, when experimental results concerning the effect of the addition of biochar to a given soil on
crop yields and/or soil properties are published, details regarding the properties of the used biochar
should be well reported. The inclusion of this valuable information seems to be essential in order to
establish the appropriate process conditions to produce a biochar with more suitable characteristics.

Keywords:

Biochar; Pyrolysis; Soil fertility; Carbon sequestration.

ACS Paragon Plus Environment

Page 3 of 56

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Environmental Science & Technology

TOC/Abstract art

Operating conditions

Pyrolysis
gas
Biomass

Slow Pyrolysis

Charcoal
(40-60% yield)

Liquid
fraction

BIOCHAR USE

ACS Paragon Plus Environment

Environmental Science & Technology

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Page 4 of 56

1. INTRODUCTION
Concerns about climate change and food productivity have recently generated interest in biochar,
a form of charred organic matter which is applied to soil in a deliberate manner as a means of
potentially improving soil productivity and carbon sequestration.1 The idea of adding charcoal to
soil in order to increase its fertility is to be inspired by the ancient agricultural practices, by means of
which terra preta soils were created.2 These soils, which may occupy up to 10% of Amazonia,3 are
characterized by high levels of soil fertility compared to other soils where no organic carbon
addition occurred. Besides the potential of biochar to enhance the fertility of agricultural soils, its
apparent ability to increase the capacity of soil to retain water makes biochar a very promising
alternative in the current context of climate uncertainty.
A high number of recent studies have highlighted the benefit of using biochar in terms of
mitigating global warming and as a strategy to manage soil health and productivity.48 In the most of
cases, these studies are constrained by limited experimental data and are geographically limited.
This fact can be considered as expected because the complexity of the experimental tasks.
Biochar can be produced by several thermochemical processes: conventional carbonization or
slow pyrolysis, fast pyrolysis, flash carbonization, and gasification. Slow pyrolysis has the
advantage that can retain up to 50% of the feedstock carbon in stable biochar.8 Biomass pyrolysis
and gasification are well-known technologies for the production of biofuels and syngas. However,
commercial exploitation of biochar as a soil amendment is still in its infancy.2 Pyrolysis process and
its parameters (principally final temperature, heating rate, pressure, and residence time at the final
temperature) greatly condition the biochar production and quality. In addition to this, the intrinsic
nature of the biomass feedstock also interacts with the rest of variables in determining the properties
of the produced biochar.9,10

ACS Paragon Plus Environment

Page 5 of 56

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Environmental Science & Technology

The relationship between biochar properties and its potential to enhance agricultural soils is still
unclear and does not allow the establishment of the appropriate process conditions in order to
produce a biochar with desired characteristics.11 Several recent studies have been focused on
providing a characterization methodology of biochars.1114 These studies represent an initial step,
but further efforts are needed to perform soil tests in order to establish an appropriate formulation of
desired biochar properties.
The specific aim of the present study is to review and analyze the available published studies
related to biochar production, characterization, and its addition into agricultural soils. As a result of
this review process, the objective of the author is to highlight the research needs for this exciting
field of study. Among other potential research gaps, this paper focuses on the interaction between
biochar production and its potential applicability to agricultural soils. In this sense, the knowledge of
the effect of the operating conditions governing the pyrolysis process on the properties of the
resulting biochar (degree of aromaticity, cation exchange capacity...) for a given biomass feedstock,
seems to be necessary to facilitate future research on this topic.
2. THE BIOCHAR CONCEPT
Biochar is a carbon-rich, fine-grained, porous substance; which is produced by thermal
decomposition of biomass under oxygen-limited conditions and at relatively low temperatures (<
700 C).1,2 The definition adopted by the International Biochar Initiative (IBI) furthermore specifies
the need for purposeful application of this material to soil for both agricultural and environmental
gains.2 This fact distinguishes biochar from charcoal, which is used as a fuel for heat, as an
adsorbent material, or as a reducing agent in metallurgical processes.1
One of the interesting properties of biochar, that makes it attractive as a soil amendment, is its
porous structure, which is believed responsible for improved water retention and increased soil
surface area.2 Furthermore, the addition of biochar to soil has been associated with an increase of the
5

ACS Paragon Plus Environment

Environmental Science & Technology

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Page 6 of 56

nutrient use efficiency, either through nutrients contained in biochar or through physico-chemical
processes that allow better utilization of soil-inherent or fertilizer-derived nutrients.2 In addition to
the above-mentioned potentially beneficial effects, a key property of biochar is its apparent
biological and chemical stability. In fact, studies of charcoal from natural fire and ancient
anthropogenic activity indicate millennial-scale stability.1516 This property can allow biochar to act
as a carbon sink.2
According to the above-explained considerations, the conversion of biomass to long-term stable
soil carbon species can result in a long-term carbon sink, as the biomass removes atmospheric
carbon dioxide through photosynthesis.17 For this reason, the use of biochar can imply a net removal
of carbon from the atmosphere.1 Furthermore, three complementary goals can be achieved by using
biochar applications for environmental management: soil improvement (from both productivity and
pollution points of view), waste valorization (if waste biomass is used for this purpose), and energy
production (if energy is captured during the biochar production process). In light of this, the
production of biochar from agricultural residues and/or forest biomass appears to be a very
promising alternative to integrate carbon sequestration measures and renewable energy generation
into conventional agricultural production.17
3. BIOCHAR PRODUCTION
Biochar can be produced as a co-product from several different processes. The properties of a
given biochar strongly depend on each process characteristics and also on the material to which the
process is applied. In the next sections, several technologies currently in use or under development
are reviewed.
Slow pyrolysis
Conventional carbonization or slow pyrolysis processes, in which a relatively long vapor
residence time and a low heating rate are the key process parameters, have been used to generate
6

ACS Paragon Plus Environment

Page 7 of 56

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Environmental Science & Technology

charcoal from many years ago.18 As a result of many relatively recent studies focused on increasing
charcoal yields,9,10,1925 several variables and factors that play a critical role during the pyrolysis
process have been identified; among these are peak temperature, pressure, vapor residence time, and
moisture content.19
The peak temperature is the highest temperature reached during the process.19 As a general rule,
the charcoal yield decreases as temperature increases. However, an increase of the peak temperature
results in an increase of the fixed-carbon content in biochar.19,26,27 This increase is especially
pronounced in the temperature range from 300 to 500 C. In addition, the peak temperature has
influence on surface area and pore size distribution (both properties generally related to specific
adsorptive properties) of charcoals. Khalil28 reported very low surface areas for charcoals (from a
wide variety of biomass feedstocks) pyrolyzed at temperatures near 550 C. However, setting peak
temperatures higher than 700 C does not seem appropriate to generate charcoals with potentially
better adsorptive properties.2931
Pyrolysis or carbonization at elevated pressure (1.03.0 MPa) seems to improve the charcoal yield
as a consequence of the increase of the vapor residence time within the solid particle. This effect,
which results in a substantial increase of the secondary charcoal production (as a consequence of the
decomposition of vapors onto the solid carbonaceous matrix), is magnified when the gas flow
through the particle bed is small.19 Furthermore, it should be kept in mind that the energy demand of
the pyrolysis process is closely related to the production of charcoal by primary (endothermic) and
secondary (exothermic) reactions.20,32 In line with this, an increase of the charcoal produced by
secondary reactions can significantly reduce the amount of energy required to sustain the process.
Pyrolysis pressure also produces an effect on the porosity of produced charcoals. Cetin and coworkers33 reported a slight decrease of the total surface area by increasing pressure during the
pyrolysis of several biomass feedstocks (radiata pine, eucalyptus wood, and sugarcane bagasse).
7

ACS Paragon Plus Environment

Environmental Science & Technology

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Page 8 of 56

However, in a recent study conducted by Melligan and co-workers,34 a dramatic decrease of the
BET (Brunauer, Emmett and Teller) surface area of charcoals obtained by slow pyrolysis (at 13 K
min1 and at a peak temperature of 550 C) of miscanthus is reported (from 161.7 m2 g1 at 0.1 MPa
to 0.137 m2 g1 at 2.6 MPa). The authors attributed this result to a clogging of the pores by tar
deposits as a consequence of the high pressure. In addition to this, Melligan and co-workers also
reported that chars formed at high pressure had more extended fused aromatic structures, reflected
also in the higher carbon contents, than those obtained at atmospheric pressure.
Regarding the moisture content of the biomass feedstock, results obtained in previous studies2035
indicated that high moisture contents (in the range of 4262%) can improve the yield of charcoal at
elevated pressures. This finding makes certain agricultural residues, which are characterized by high
moisture contents, particularly attractive for biochar purposes. In addition to the moisture effect, it
must also be taken into account that the charcoal yield from a given biomass feedstock is influenced
by its inherent composition (holocellulose, lignin, extractives, and inorganic matter). In this sense,
pyrolysis of biomass species with high lignin contents can produce higher charcoal yields.19,36 An
increase of the charcoal production was also observed by Di Blasi and co-workers37 when pyrolyzed
extractive-rich woods (e.g., chestnut) instead of another wood varieties with lower extractives
contents (e.g., beech).
Special attention has been focused on discussing the influence of the inorganic matter on pyrolysis
product distribution. During biomass pyrolysis, inorganic matter, especially alkali and alkali earth
metals, catalyses biomass decomposition and char forming reactions.38 Several researchers have
obtained lower charcoal yields when the biomass feedstock was pre-treated with hot water (at 80 C)
as a measure to reduce the ash content.3943
Additional process variables that might affect charcoal yields are the soak time at peak
temperature and the particle size. Regarding the first one, Antal and Gronli19 stated that the soaking
8

ACS Paragon Plus Environment

Page 9 of 56

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Environmental Science & Technology

time has no effect on the charcoal yield because pyrolysis kinetics is primarily governed by
temperature. This assumption is consistent with experimental data reported in several research
studies.4445 Concerning the effect of the particle size on the charcoal yield, it seems reasonable to
assume that an increase in particle size leads to higher charcoal yields. As the particle size is greater,
the rate of diffusion of the volatiles through the char decreases and, consequently, the formation of
additional char by means of secondary reactions should be expected.9,10,23
Table 1 reports several experimental data from the literature for slow pyrolysis of different
biomass feedstocks under various operating conditions.24,27,46,47 A qualitative examination of the
data reported in Table 1 appears to confirm the effects of some variables (peak temperature,
pressure, and both moisture and lignin contents) on the charcoal yield. In this sense, and in
agreement with the considerations previously mentioned in this section, the charcoal yield is favored
by increasing pressure and/or decreasing peak temperature. Moreover, greater charcoal yields were
obtained, under identical operating conditions, for biomass samples with high moisture and/or lignin
contents.
To reach a more conclusive interpretation of the experimental data showed in Table 1, normalized
principal components analysis (NPCA) was applied to the same data using the Rcmdr package in R
(version 2.14.2). NPCA is a robust statistical technique, the purpose of which is to reduce the
complexity of the multivariate data into the principal components space and then choose the first
principal components that explain most of the variation in the original variables.10,42,48 The
following variables, the values of which are listed in Table 1, were selected for the principal
component analysis: pressure, peak temperature, heating rate, soaking time, lignin content, moisture
content, and ash content. Figure 1 shows both score and loading plots obtained for the three first
principal components (derived from the correlation matrix), which explained 67% of the total
variance. From the analysis of the results displayed in Figure 1, some considerations can be
9

ACS Paragon Plus Environment

Environmental Science & Technology

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Page 10 of 56

outlined: (a), the first principal component is partially due to pressure (data points with a large xcoordinate correspond to experiments performed under elevated pressure); (b); pressure and char
yield are highly correlated, as can be seen from the loading plot (Figure 1a) (c), it seems that the
peak temperature is associated with the second principal component because the value of the ycoordinate increases as the peak temperature raises; (d), the second principal component is also
partially explained by soaking time, but no correlation between this variable and char yield is
observed; (e) the third principal component seems to be related to the intrinsic properties of the
biomass feedstock (in this case, lignin and moisture contents).
As a preliminary conclusion, it can be stated that predicting the charcoal yield as a function of
both operating conditions and biomass properties is still difficult despite the large number of studies
published to date. It seems clear that increasing pressure and decreasing peak temperature enhances
the yield of charcoal. Nevertheless, the effect of the intrinsic properties (holocellulose and lignin
contents, moisture, amount and composition of the mineral matter) of the biomass feedstock on
the charcoal yield (as well as on the chemical and textural characteristics of produced charcoals) is
critical and needs to be experimentally determined.
Alternative processes
Fast Pyrolysis
Fast pyrolysis uses high heating rate (above 200 K min1) and short vapor residence time (around
2 s). The peak temperature is usually set between 500 and 550 C in order to obtain the highest biooil yield.4952 These operating conditions particularly favor the formation of liquid products (biooil), but inhibit the formation of charcoal.18 Duman and co-workers,53 in a recent study, compare the
charcoal yields from both cherry seeds and cherry seed shells obtained using two pyrolysis
processes: a fixed bed reactor heated at 5 K min1 and a fluidized bed reactor (fast pyrolysis). At a
constant final temperature of 500 C, the charcoal yield decreased from 27% to 18% for cherry seeds
10

ACS Paragon Plus Environment

Page 11 of 56

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Environmental Science & Technology

(from 28% to 17% for cherry seed shells) when using the fluidized bed reactor instead of the fixed
bed one.
Concerning the physical properties of charcoals obtained by fast pyrolysis, Boateng,54 Brewer and
co-workers,11 and Mullen and co-workers55 reported relatively low BET surfaces areas (3.1021.6
m2 g1) for chars formed from switch grass and corn stover in a fluidized bed reactor. This result is
expected because of the short residence time of solid particles.56 In addition to this, Brewer and coworkers11 also observed a very small particle size for charcoals obtained from fast pyrolysis of
switch grass. This fact is mainly due to the small particle size of the feedstock (averaging around 1
mm) usually required in fast pyrolysis systems, and, probably, to the hypothesis that fast
devolatilization might create very fragmented char structures.57 From a chemical composition point
of view, charcoals obtained at high heating rates are characterized by high oxygen content50 and low
calorific value,53 probably as a result of the relatively short particle residence time.
In the last years, increasing attention has been focused on upgrading the composition and qualities
of the bio-oil product by means of the addition of a catalyst (in-situ upgrading).58,59 Taking into
account the catalytic effect of alkaline and alkaline earth metals on the pyrolysis of biomass, several
researchers measured the effect of impregnating a given biomass feedstock with a potassium,
sodium or magnesium salt on both product distribution and composition.6065 Di Blasi and coworkers63 observed a substantial increase of the char yield (from 19% to 30% in weight basis)
during the fast pyrolysis, at a peak temperature of 527 C, of fir wood previously impregnated with
an aqueous solution of KOH. A similar finding (an increase of the char yield of around 10%) was
reported by Wang and co-workers65 during the pyrolysis of pine wood particles physically mixed
with potassium carbonate. In both cases, the increase of the charcoal yield occurred at the expense of
the liquid-phase organic products, as a consequence of the catalytic enhancement of the secondary
charring reactions of primary volatiles.
11

ACS Paragon Plus Environment

Environmental Science & Technology

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Page 12 of 56

Nevertheless, the in-situ upgrading of the bio-oil product is usually performed using catalysts
based on mesoporous aluminosilicate materials (Al-MCM-41) or microporous HZSM-5 zeolites.66
69

Zhang and co-workers68 reported a decrease of the char yield (from 23.2% to 20.1%) when

corncob samples were catalytically pyrolyzed using a HZSM-5 zeolite. The bio-oil yield also
decreased in the presence of catalyst (from 33.9% to 13.7%). However, the collected liquid after
catalytic fast pyrolysis exhibited interesting properties for its use as transport oil (lower oxygen
content and higher heating value than the bio-oil collected without catalyst).
Flash carbonization
The flash carbonization (FC) process has been developed at the University of Hawaii (UH) under
the leadership of Professor Michael J. Antal. This process is a novel procedure by which biomass
can be converted to charcoal in a more efficient way than conventional carbonization or slow
pyrolysis.7072 A canister containing a packed bed of a given biomass feedstock is placed within a
pressure vessel. Air is used to pressurize the vessel to an initial pressure of 12 MPa, and a flash fire
is ignited at the bottom of the packed bed. After a few minutes, air is delivered to the top of the
packed bed and biomass is converted to charcoal. The total reaction time is less than 30 min and the
temperature profile of the packed bed is conditioned by several factors: biomass feedstock, moisture
content of the feedstock, heating time, and the total amount of air delivered.72 In any case, the flame
front moves up the packed bed, causing the middle and top temperatures to successively increase,
until reaching values near 600 C.
Using the FC process, Nunuora and co-workers70 reported high fixed-carbon yields (in the range
28%32%) for two types of biomass feedstocks (corncob and macadamia nut shells). The fixed
carbon yield (yFC) (which is a better index than the charcoal yield, because the yFC takes into account
the chemical composition of the produced charcoal) is defined as follows:

12

ACS Paragon Plus Environment

Page 13 of 56

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Environmental Science & Technology

m
yFC char
mbio

% FC

100 %ash

(1)

where %FC and %ash denote the percentage of fixed-carbon contained in the charcoal and the
percentage of ash in the feedstock, respectively.27 The ratio between mchar (dry mass of produced
charcoal) and mbio (dry mass of feedstock) correspond to the charcoal yield.
The thermochemical equilibrium value of the fixed-carbon content can be useful for comparison
purposes. These theoretical values can be calculated using the STANJAN software,73 as a function
of the elemental biomass composition, final temperature, and pressure. Figure 2 shows a comparison
of the attainment of the theoretical yield of fixed carbon for four biomass feedstocks (leucaena
wood, oak wood, macadamia nut shells, and corncob), which were pyrolyzed using two different
processes: slow pyrolysis at 1.0 MPa (heating at 6 K min1 up to 450 C with no soaking time)27 and
flash carbonization at 1.01.5 MPa.72 The results obtained for leucaena wood, oak wood, and
macadamia nut shells were similar in terms of fixed-carbon retained in the charcoal. However, a
substantial increase of the yFC value (reaching 100% of its theoretical maximum value) is deduced
for the corncob samples when they were carbonized using the FC process. In other words, it is
possible to retain, in the charcoal, the maximum amount of fixed carbon from corncob samples
using the FC process. Taking into account the elemental analysis of corncob samples (43% of C)72
and the yFC value obtained using the flash carbonisation process (28%),72 it can be determined that
65% of the carbon initially contained in the feedstock was transformed to carbon in the charcoal. In
addition to this, the FC process seems to be a very interesting option, because the reaction times are
very short (< 30 min) compared to the slow pyrolysis process.
Gasification
Gasification is a thermochemical process by which a carbonaceous feedstock (coal, biomass, or a
mixture of both) is converted into a non-condensable gas at high temperatures (> 800 C).18,7476

13

ACS Paragon Plus Environment

Environmental Science & Technology

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Page 14 of 56

When air is employed as the gasification agent (the most common case when biomass wastes are
processed), the process involves partial combustion of the fuel to generate a combustible gas with a
low heating value of 3.510.0 MJ Nm3,75 which can be used as a fuel for boiler, gas turbine or gas
engine. The quality of the producer gas is improved when other oxidizing agents are used (i.e.,
steam, carbon dioxide, or a mixture of oxygen and steam). The producer gas obtained in this way is
rich in carbon monoxide and hydrogen and its field of applicability is wide: chemical synthesis, fuel
cell feed, hydrogen production, etc.
Charcoal yield from gasification is very low (510%)11,76 because of the high operating
temperature and the partial oxidizing atmosphere. In addition, the produced char from gasification
systems can exhibit a high concentration of metals and minerals depending of the ash content and
composition of the feedstock. This fact may imply potential safety concerns with regard to the
application of this kind of biochars to soil.77
For all of the reasons mentioned above, it seems clear that conventional gasification systems,
whose purpose is to maximize the gas product fraction, are not the best option to generate biochar
for soil amendment. However, a partial and controlled gasification process using air, steam or CO2
as the oxidizing agent; may be an interesting way to improve the textural properties of a given
charcoal, which has previously been obtained by a pyrolysis process.30,78 The gasification step is
also known as physical activation process and is widely used to produce activated carbons from
biomass feedstocks for adsorption and catalysis purposes.7984 The percentage of burn-off (usually
ranged from 30% to 55%)80 and the activation temperature (which is kept constant at 700850 C)80
are the key operating parameters during the gasification step.
Table 2 lists the activation conditions and the textural properties of several activated charcoals,
which have been collected from some published works.80,8387 In all of these previous studies, the
surface area (SBET) was calculated using the BET equation from the N2 adsorption data; the total
14

ACS Paragon Plus Environment

Page 15 of 56

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Environmental Science & Technology

pore volume (V) was estimated by converting the amount of N2 gas adsorbed at a relative pressure of
0.99 to liquid volume of nitrogen; and the micropore volume (V0) was determined according to the
DubininRadushkevich method88 (see detailed explanation of textural characterization in section 5).
Although a direct comparison of the results reported in Table 2 is difficult, because of the
different experimental conditions (for both pyrolysis and activation processes) and biomass
feedstocks used in each study, it is pertinent to note that both the surface area and the micropore
volume are affected by the conditions of activation. Nevertheless, the choice of the best set of
activation conditions will depend on the experimental results obtained for a specific biomass
charcoal. In other words and as has been previously noted for the pyrolysis charcoal yield, the nature
of the precursor (biomass) has a great influence on the properties of the resulted activated carbons.
Despite the positive benefits linked to the production of valuable porous materials from biomass
feedstocks, it must be kept in mind that as the conversion of the fixed-carbon (during the
gasification or activation step) becomes greater; the carbon sequestration potential of the biochar
becomes smaller. For this reason, it will be interesting to see if a compromise between the textural
properties and the fixed-carbon yield can be reached for biochar purposes.

4. EFFECTS OF BIOCHAR ON SOIL QUALITY


In this section, the effects of the addition of biochar on soil properties, processes, and functions
are reviewed.

Soil properties
The addition of biochar into soil can alter soil physical properties such as structure, pore size
distribution, bulk density, and texture. This fact brings important implications for soil aeration,
water holding capacity, plant growth, and soil workability.89,90 It is reported that biochar application
into soil could increase the overall surface area of the soil91 and consequently, could improve soil
water retention89 and soil aeration.92 Laird and co-workers93 reported that the specific surface area of
15

ACS Paragon Plus Environment

Environmental Science & Technology

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Page 16 of 56

a typical very deep loam soil (Clarion soil map unit 1138B, from Iowa State, USA) increased from
130 to 153 m2 g1 as the biochar concentration increased from 0 to 20 g kg1. An increased surface
area might also benefit the overall sorption capacity and the native microbial communities of soils.89
However, experimental evidence of such mechanisms is very scarce at present.
The soil water retention is determined by the distribution and connectivity of pores in the soilmatrix, which is largely regulated by soil particle size (texture), combined with structural
characteristics (aggregation) and the soil organic matter (SOM) content.90 The soil aggregation can
be improved by addition of biochar, as a consequence of the related interactions with SOM,
minerals, polymers from micro activity, clay, and microbiological activity. All of these reasons may
explain the data of Glaser et al.,5 who reported an increase in the water hold capacity of 18% for
anthrosols (man-made tropical soils) rich in biochar in comparison to adjacent soils in which biochar
was absent.
Certain studies have reported high cation exchange capacity (CEC) values for biochar2,94
(consistently higher than that of whole soil, clay or soil organic matter2), probably due to its
negative surface charges94. This fact can enable biochar to act as a binding agent. Nevertheless, the
addition to a given soil of a biochar with a high CEC does not necessarily imply an increase in the
cation exchange capacity of the soil, which results in an enhancement of the ability of the soil to
adsorb and retain cations (e.g.; Mg2+, Ca2+, K+, and NH4+). The CEC of a given soil indicates how
well some nutrients (cations) can be bound to the soil, and, therefore; available for plant uptake. An
increase in nutrient retention also results in decreased leaching losses below to the effective rooting
zone. Leaching of nutrients from soils decreases soil fertility, promotes soil acidification and
negatively affects the quality of surface and groundwater.95
Experimental results obtained from previous studies concerning the effects of biochar addition on
the CEC of soil are given in Table 3. In the most of cases, the incorporation of biochar into soil
16

ACS Paragon Plus Environment

Page 17 of 56

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Environmental Science & Technology

increased both pH and CEC. However, this improvement in soil quality was not observed for a high
pH soil tested by Van Zwieten and co-workers96 (a loamy calcisol from a vineyard in Victoria,
Australia). In addition, Van Zwieten and co-workers observed differences in soil properties when
these soils were amended with different biochar samples. More in detail, these researchers observed
that biochar samples with higher liming values (measured in percentage of CaCO3 equivalent)
showed a better ability to increase pH of an acid soil (ferralsol). Partially in line with this, Yuan and
co-workers94 also tested the effect of biochar incorporation on the pH of an acidic soil (ultisol from
Anhui province, China). In that study, authors reported that the liming effects of the biochars
produced from several crop straws (canola, corn, soybean, and peanut) increased with the rise of the
peak pyrolysis temperature, the value of which seems to affect both the alkalinity and the form of
alkalis of a given biochar, as recently suggested by Hossain and co-workers.97

Soil processes
Biochar stability on the environment
The stability of biochar in soil is a key parameter in order to evaluate the potential of using
biochar as a CO2 sequestration tool. Current evaluations of the age of black carbon particles from
anthropogenic activity (and from natural fire events) indicate great stability of (at least) a significant
component of biochar, ranging from several thousands to hundreds of years.90,98,99 Nevertheless,
freshly-made biochar is not an inert material and can be oxidized in the short term by contact with
strong chemical oxidants at high temperatures.100,101 In soil, biochar can be degraded by both
photochemical and microbiological processes, as reported in a relatively small number of short-term
incubation studies.101,102 From these experimental results, it was also deduced that biological
decomposition was negligible compared to abiotic degradation.101
Fresh biochar surfaces are commonly hydrophobic and have negative surface changes.90,103
Nevertheless, biochar in the soil environment can probably be oxidized over time resulting in a
17

ACS Paragon Plus Environment

Environmental Science & Technology

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Page 18 of 56

probable accumulation of carboxylic functionalities in the surface of biochar particles.104 This fact,
which depends at least on the biochar characteristics and the environmental conditions, could
enhance the interactions between biochar and other soil components, such as organic and mineral
matter.101,104
The influence of biochar properties (i.e., particle size, pore size distribution and surface
chemistry) on the both short-term and long-term carbon loss (mineralization) of biochar remains still
unclear. Hamer and co-workers102 reported that biochar obtained from corn stover and rye was
mineralized more rapidly than that produced from wood, indicating a certain role of the biomass
type in the stability of biochar (influence of H/C, O/C, and C/N ratios). Baldock and Smernik105
observed, for biochar produced from red pine wood, an inversely proportional relationship between
the pyrolysis peak temperature and the carbon loss by mineralization. Recently, Nguyen and coworkers106 found that increasing the peak temperature from 350 to 600 C (during slow pyrolysis of
corn residues and oak wood) produced a decrease in the carbon loss for mixtures of biochar and pure
sand incubated for 1 year. An increase of the pyrolysis peak temperature results in a greater degree
of aromaticity of the biochar and, consequently, in a greater chemical recalcitrance. Furthermore,
these researchers also observed that the remaining carbon for mixtures of biochar from corn residue
and sand (at a given peak temperature) was lower than that of mixtures composed by biochar from
oak wood. Besides the biomass feedstock and the pyrolysis peak temperature, the stability of
biochar also depends on environmental conditions and soil type. Nguyen and co-workers16,106
reported strong influences of both water regime (saturated or unsaturated conditions) and
temperature on the mineralization of biochar; whereas Qayyum and co-workers observed different
carbon mineralization rates of a wheat straw-derived biochar for three types of soils (ferralsol.
topsoil lixisol, and subsoil lixisol). The last authors reported the lowest mineralization rate for the
ferralsol soil type.107
18

ACS Paragon Plus Environment

Page 19 of 56

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Environmental Science & Technology

Greenhouse gas (N2O and CH4) emissions


Under anaerobic conditions N2O is emitted from soil through denitrification, a microbially
facilitated process of NO3 reduction that may ultimately produce N2. In addition, nitrifying bacteria
generally involved in conversion of N2 to ammonium can simultaneously promote denitrification.2
The addition of biochar in a given soil can decrease the availability of N for denitrification and,
consequently, reduce the total N2O emissions. Yanai and co-workers108 showed, in a short-term
laboratory chamber experiment, a significant decrease in N2O emissions from a wetted volcanic ash
soil (hapludand) when biochar derived from municipal biowaste was applied (at a rate of 180 t ha1).
In line with this, Zhang and co-workers109 showed that the total N2O emissions from a hydroagric
stagnic anthrosol were decreased by 40%51% and by 2128% when biochar (produced by slow
pyrolysis of wheat straw at 350550 C) was added at a rate of 40 t ha1 compared to the control
treatments with and without N-fertilizer, respectively. Similar findings were also reported by
Sarkhot and co-workers110 for a dairy manure-derived biochar (26% reduction in cumulative N2O
flux).
On the other hand, wide variations in the rates of CH4 emissions from soils amended with biochar
have been reported in the literature. Xiong and co-workers111 observed that CH4 emissions depended
on the properties of soil. These authors measured the CH4 emissions during the flooded season for
two Chinese anthrosols with different CEC and organic carbon content. After analyzing
experimental results (in which the highest CH4 emissions was measured for the soil type with a
highest organic C content) and as a preliminary conclusion, Xiong and co-workers stated that soil
organic carbon is more important than CEC as driving factor controlling CH4 production.
Rondon and co-workers112 reported that application of wood-derived biochar at a rate of 20 t ha1
remarkably increased the annual methane sink in an acidic tropical soil. In contrast to this finding,
Zhang and co-workers109 reported that CH4 emissions were increased by 34% and 41% in a
19

ACS Paragon Plus Environment

Environmental Science & Technology

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Page 20 of 56

hydroagric stagnic anthrosol amended with wheat straw-derived biochar at 40 t ha1 compared to the
treatments with and without N-fertilizer, respectively. As suggested by Zhang and co-workers, labile
components of biochar could be decomposed and become the predominant source of methanogenic
substrates, thus promoting CH4 production. In line with this and according to Van Zwieten and coworkers113, the source and chemical properties of biochar might also have an influence on CH4
yield. In addition, and as has mentioned before, wide variation of soil CH4 emission has been
reported for soils with different chemical and physical properties111 and under different water
regimes. In any case, the precise mechanism behind the soil CH4 emissions still remains unclear.
Sorption of hydrophobic organic compounds (HOCs)
Biochar incorporation into soil can enhance the sorption capacity of soils towards hydrophobic
organic compounds (such as PAHs and pesticides).90 Previous studies have indicated that this
negative effect can be a function of the chemical and structural properties of the contaminant (e.g.,
molecular weight and hydrophobicity),114116 as well as of the surface area, pore size distribution,
and functionality on surface of the biochar.114,115
The influence of the textural properties of biochar on sorption capacity was analyzed in previous
studies,116,118 in which researchers observed a strong (and expected) effect of the pyrolysis peak
temperature on sorption capacity for biochars obtained from wood and wheat residues. As the
pyrolysis peak temperature increases, produced biochars exhibits a greater surface area (and a
greater micropore volume) and a lower oxygen content (lower O/C ratio). Taking into account that
the O/C ratio of a given biochar is a potential indicative of both its hydrophilicity and polarity, an
increase of the pyrolysis peak temperature probably causes a decrease in polar surface groups, and
consequently, a reduction of the biochar affinity for water molecules. As a consequence of both the
increase of pore surface area and the decrease of water affinity, the sorption capacity of biochars is
expected to increase with the pyrolysis final temperature, as observed by Chun and co-workers118
20

ACS Paragon Plus Environment

Page 21 of 56

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Environmental Science & Technology

and Wang and co-workers.116 In contrast to these results, Kinney and co-workers119 observed that
biochars obtained from several feedstocks (magnolia leaves, apple wood chips, and corn stover) by
slow pyrolysis (at a peak temperature ranged from 400 to 600 C) exhibited a very low
hydrophobicity. Kinney and co-workers also reported a statistically significant effect of the presence
of surface alkyl functionalities (CH), which were detected when biochars were pyrolyzed at a peak
temperature below 400 C, on the hydrophobicity of the analyzed biochars. These apparently
contradictory results suggest that further studies focused on analyzing the hydrophobicity of
biochars are required.
Other sources of soil contamination
This section is focused on the potential for soil contamination linked to some component of
biochar, such as heavy metals and PAHs. Despite the fact that this type of contamination can lead to
severe public health problems, relatively little attention has been focused on this issue.90
Biochar produced from pyrolysis of some organic wastes, such as sewage sludge and tannery
residue, generally retains high levels of heavy metals (e.g., chromium, cooper, nickel, and zinc).97,120
However, McHenry17 suggested that high levels of biochar addition (> 250 t ha1) are needed to
potentially contaminate soil, surface water and crops. Obviously, this topic needs further assessment
in future studies.
Otherwise, it seems clear that biomass pyrolysis at peak temperatures above 700 C could
generate heavily condensed PAHs.121,122 Brown and co-workers123 reported that several biochar
products, which were obtained by slow pyrolysis at different peak temperatures (ranging from 450
to 1000 C) from pitch pine wood, exhibited PAHs concentrations ranged from 3 to 16 g g1 (with
the highest value at the highest peak temperature). Brown and co-workers also analyzed the PAHs
content of a natural biochar (charred pine from a prescribed burn area), showing that this value (28
g g1) was slightly higher than that measured for synthetic biochars. This preliminary finding could
21

ACS Paragon Plus Environment

Environmental Science & Technology

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Page 22 of 56

suggest that PAHs levels in biochar can be often comparable (or even lower) than those found in
some soils.90

Soil productivity
An increase in soil fertility is the most frequently reported benefit linked to adding biochar to
soils. Most of the published studies to date have been conducted for tropical soils.90 In tropical or
sub-tropical environments, soil fertility tends to be poor due to rapid mineralization of soil organic
matter, the low cation exchange capacity (CEC) of the tropical soils (which is usually due to their
clay content and mineralogy), and the low nutrient contents.5 Moreover, the use of inorganic
fertilizers in these types of soils has certain drawbacks, the most important of which are the high
cost of continuous applications of fertilizers and their low efficiency in highly weathered soils.124,125
Some previous studies reported that biochar addition in several tropical soils resulted in an increase
of soil nutrient availability.5,126129 In the short-term basis, the direct nutrient additions with the
added biochar (e.g., K, P, and Ca, which are present in the inorganic fraction of biochar) seems to be
responsible for short-term enhancement of soil fertility.126 Regarding the long-term effect of biochar
on nutrient availability, it depends on an appropriate increase of both CEC and surface
oxidation.125,130 Previous investigations, which have been focused on analyzing both CEC and pH
evolution over time, reported an increase in both variables as biochar addition time increased.101
Currently published studies considering the effect of biochar addition on crop yield were
generally performed on small scale and sometimes, without considering the environmental
conditions, under which a decrease of the biochar content in the soil can occur through
decomposition (when temperature raises), leaching, or erosion.98 Glaser and co-workers131 reviewed
a substantial number of earlier studies, which were conducted for tropical soils during the 1980s and
1990s. These studies reported positive impacts of biochar additions at a low application rate of 0.5
Mg ha1 on several plant species. However, biochar addition at higher rates (> 100 Mg ha1) seemed
22

ACS Paragon Plus Environment

Page 23 of 56

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Environmental Science & Technology

to inhibit plant growth. On the other hand, a later study conducted by Steiner and co-workers132
showed that biochar application (at a rate of 11 Mg ha1) significantly improved plant growth for a
highly weathered Central Amazonian upland soil fertilized with NPK (in comparison to the effect of
the same rate of NPK-fertilizer without biochar).
Table 4 presents examples of experimental studies focused on investigating the response of crops
to biochar application. As can be deduced from the data reported in Table 4, the effect of biochar
depends on several factors including the soil type, the addition rate, and the kind of crop. Moreover,
an interaction between biochar and fertilizer addition is generally observed. In this sense, and as
argued above, the fertility of tropical and sub-tropical soils (such as acidic ferralsols and nitisols)
seems to substantially improve by biochar treatment,5,96,125,132,133 especially when biochar was
applied together to inorganic fertilizers.96,134,135 However, Van Zwieten and co-workers96 reported
significant decreases in wheat and radish biomass production for a high pH calcisol. Negative
impacts on crop yield were also observed by Haefele and co-workers133 for rice growth in a gleysol
(which had a high CEC and base saturation and high N, P, and K availability). A mechanism to
explain the negative effect of biochar for these soil types was proposed by Lehmann and coworkers:126 the available nutrients applied with biochar in this type of soils are not limiting, the CEC
is very high already, and water stress does not occur; nevertheless, the high C/N ratio of biomass
probably limits N availability (from both soil and inorganic fertilizer), causing a decrease of grain
yield.
From Table 4, it is also important to highlight the promising results reported by Vaccari and coworkers136 regarding the yield in durum wheat for a silt loam soil (with a pH of 5.2) under the
Mediterranean climate conditions. These preliminary results, which should be confirmed in further
studies, could indicate that the positive effect of biochar addition on soil production is also possible
for other soils than ferralsols in tropical environments.
23

ACS Paragon Plus Environment

Environmental Science & Technology

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Page 24 of 56

Unfortunately, very little information is available in the literature with respect to the influence of
both biochar properties and pyrolysis conditions on plant growth. Nevertheless, a recent study
conducted by Peng and co-workers135 reveals some interesting findings. These authors analyzed the
effect of both the pyrolysis peak temperature and the soaking time at this temperature for rice strawderived biochar on soil properties and production function. Regarding the biochar characteristics,
Peng and co-workers observed that increasing both peak temperature (from 250 C to 450 C) and
soaking time (from 2 to 8 hours) obviously decreased the biochar yield and volatile matter content
but increased the C, K, and P contents. In addition, volatile matter, O, H, and aliphatic functional
groups decreased at the expense of aromatic C as peak temperature and soaking time increased. As a
result, the biochar stability and its liming effect increased with pyrolysis peak temperature and
residence time. Nevertheless, and interestingly, no significant effects of pyrolysis conditions on the
CEC of the tested soil (a highly weathered ultisol from southern China) and the maize yield were
observed by Peng and co-workers.
In another interesting study, Deenik and co-workers at the University of Hawaii137 showed that
partially carbonized biochar containing a relatively high volatile matter (VM) content produced
lower yielding plants in biochar-amended soil compared with soil not treated with biochar. The poor
yield in the high-VM biochar amended soils could be due to an inhibition of N availability (the
authors attributed this effect to the presence of phenolic compounds in the volatile matter, which
stimulated microbial activity leading to a reduction of inorganic N). In contrast, more fully
carbonized biochar with low-VM content did not produce a negative effect on plant growth, and
when it was combined with N fertilizer, there was a significant improvement in crop yield compared
with the fertilized control. Both biochars were obtained from macadamia nut shells by means of a
flash carbonization process at different peak temperatures: 430 C for the high (225 g kg1) VM
biochar and 650 C for the low (63 g kg1) VM biochar. Deenik and co-workers,137 who conducted a
24

ACS Paragon Plus Environment

Page 25 of 56

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Environmental Science & Technology

series of short-term (46 weeks) greenhouse experiments and laboratory incubations, observed the
above-mentioned effects for two types of Hawaiian soils: an andosol (a volcanic soil) and an
uncultivated, highly weathered and extremely acid ultisol. The results obtained for the tested ultisol
are clearly in disagreement with many other studies,5,96,125,132,135 in which positive effects on plant
growth of biochar addition to acid tropical soils are reported. Nevertheless, it should be noted that
not all biochars will exhibit the same effects for a given soil type. In other words, the negative
results reported by Deenik and co-workers137 only suggest that the quality of biochar is at least as
important as the soil type.

5. BIOCHAR CHARACTERIZATION REQUIREMENTS


Taking into account that the form of carbon (aromatic or nonaromatic C) present in biochar is
believed to be related to the stability of this material on soil, a key aspect of determining the
potential of a given charcoal for biochar purpose may be the ability to characterize its surface
chemistry.11 However, additional properties should be considered in order to preliminary evaluate
the potential of a given biochar. These properties can be physical (e.g.; specific surface area and
morphology) or chemical (such as proximate and elemental analysis and mineral content). Recently,
the International Biochar Initiative has published guidelines138 to provide standardized information
regarding the characterization of biochar materials and to assist in achieving more consistent levels
of product quality. These Biochar Guidelines identify three categories of tests for biochar: test A for
basic utility properties, test B for toxicant assessment, and test C for advanced analysis and soil
enhancement properties. In the next sections, information concerning the analytical methods used to
measure biochar properties is given.

Proximate and elemental analysis


The proximate analysis yields the weight fractions of moisture, volatile matter (VM), ash, and
fixed carbon (FC). There are standardized methods for performing a proximate analysis (ASTM,
25

ACS Paragon Plus Environment

Environmental Science & Technology

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Page 26 of 56

ISO, DIN and BS). These standards are very similar in nature except for slight differences in the
operating conditions (temperature and soaking time) used to quantify the volatile matter content. As
was mentioned in previous sections, the volatile matter content is negatively correlated with peak
temperature and, according to earlier studies,137 a high value of this parameter could indicate a low
potential of a given biochar for soil amelioration purposes.
Regarding the elemental analysis, the weight percentage of carbon, hydrogen, nitrogen and sulfur
are usually determined using analytical devices, the operation of which is based on the complete
combustion with a pure oxygen atmosphere.139

Inorganic fraction characterization


Two techniques are generally applied to isolate the inorganic fraction of carbonaceous
materials:139 the low-temperature ashing (LTA) in an oxygen plasma at 100150 C and the
medium-temperature ashing (MTA) in air a 600 C. Surez-Garca and co-workers140 suggested the
use of both isolation techniques to securely identify the inorganic constituents of a given sample.
Once the inorganic fraction has been isolated, several analytical techniques can be applied to
characterize the inorganic species: Inductively Coupled Plasma Atomic Emission Spectroscopy
(ICP-AES), X-ray fluorescence (XRF), and X-ray diffraction (XRD). ICP-AES is able to determine
the absolute concentration of inorganic elements (Al, Ca, Fe, K, P, Mg, Si...).141 XRF spectrometry
is useful to determine the ash compositions in terms of weight fraction of oxides140 and XRD can be
used to identify the crystalline minerals in ash.141
Both exchangeable K and P are important parameters that can partially establish the capability of
biochar to supply nutrients to soil in a short-term basis. These contents (exchangeable K and
exchangeable P) in biochar were found to range widely as a function of the feedstock, with values
between 1.058.0 and 2.7480 g kg1, respectively.142 These ranges are somewhat wider than those
reported in the literature for typical organic fertilizers.90 Nevertheless and according to Joseph and
26

ACS Paragon Plus Environment

Page 27 of 56

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Environmental Science & Technology

co-workers,143 the role of high-ash biochars is still unknown and experimental data are needed in
order to determine the effect of the ash on soil properties on the medium and long-term basis.

Textural characterization and morphology


As has already been mentioned in the earlier sections, both the specific surface area and pore size
distribution depend mainly on two factors: the nature of the biomass feedstock and the pyrolysis
operating conditions (especially, peak temperature). To experimentally determine the textural
parameters of a biochar sample, adsorption of N2 at 77 K and adsorption of CO2 at 273 K are
typically used. From the results corresponding to the N2 adsorption isotherms, the specific surface
area based on the equation of Brunauer, Emmett and Teller (SBET); can be determined.88 From the
same adsorption isotherms and adopting the DubininRadushkevich method, the micropore volume
(V0) can be calculated.88 Furthermore, the volume of mesopores (Vme) can be estimated from the
isotherm as the difference between the volume of N2 adsorbed at a relatively pressure of 0.95 and
the value of V0.144 On the other hand, the narrow micropore volume (W0; pore width below 0.7 nm)
can be estimated from the CO2 adsorption isotherms assuming the DubininRadushkevich
method.145
Regarding the morphological characterization, Scanning Electron Microscopy (SEM) is
commonly used to analyze the char particle structure and surface topography.2,11

Surface functionality
Surface functionality can be investigated by means of Fourier Transform Infrared (FTIR)
spectroscopy. The FTIR spectra of both biomass feedstock and biochars obtained at different
pyrolysis peak temperatures are useful to analyze the gradual loss of lignocellulosic functional
groups (change in the OH stretch peak around 3400 cm1, which dominates the feedstocks
spectrum).11 Assignment of other spectral peaks of interest for biochar samples, including the
aliphatic CH stretch at 30002860 cm1, the aromatic CH stretch around 3060 cm1, and the
27

ACS Paragon Plus Environment

Environmental Science & Technology

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Page 28 of 56

various aromatic ring modes at 1590 and 1515 cm1, was proposed by Sharma and co-workers.146
The peaks characteristic of the carbonyl groups should appear in the range 16601725 cm1. The
exact position of the peaks depends on whether the carbonyl groups are in conjunction with the
aromatic ring (position below 1700 cm1) or not (position above 1700 cm1).146
X-ray photoelectron spectroscopy (XPS) can also be used for surface analysis.16,101,147 The XPS
wide-scan spectrum usually shows the presence of two main peaks in C (C1s) at around 285 eV and
O (O1s) at around 530 eV. The spectra of high resolution XPS of C1s and O1s are used to quantify
the carbon and oxygen forms on the biochar surface. For the C1s spectrum, different binding
energies are assigned to CC, C=C, CH, CO, C=O, and COO stretches; whereas for the O1s
spectrum, signal peaks at different binding energies can be attributed to O=C and OC stretches.101

Aromatic character
Solid-state

13

Magic Angle Spinning (MAS) Nuclear Magnetic Resonance (NMR) is

commonly used for making quantitative comparisons without recurring to the procedure of taking
peak ratios. Rather, each resonance peak can be quantified in relation to the total resonance
intensity, giving therefore the relative abundance of individual molecular groups.146 As mentioned
before, the aromatic character of the produced biochar seems to be directly correlated to the value of
the pyrolysis peak temperature: as peak temperature increases it is expected to show a higher
aromatic structure. Freitas and co-workes148 reported

13

C Cross Polarization (CP) MAS NMR

spectra for biochars obtained by pyrolysis of rice hulls at different peak temperatures. For the
biochars obtained at a peak temperature of 300 C, these authors observed two main resonance lines,
around 130 ppm (broader) and 148 ppm, associated with non-oxygenated and oxygenated aromatic
carbons, respectively. Simultaneously and for the same biochar samples, Freitas and co-workers147
observed broad resonance around 31 ppm, probably associated with aliphatic chains, and the
development of a small signal near 208 ppm, ascribed to ketone groups. Regarding the biochars
28

ACS Paragon Plus Environment

Page 29 of 56

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Environmental Science & Technology

produced at higher peak temperatures (390605 C), the authors showed a progressive development
of a well-defined aromatic resonance, centered at 125 ppm, which occurs simultaneously with the
attenuation of the signals corresponding to oxygenated aromatic carbons (around 150 ppm) and
aliphatic groups (broad line around 30 ppm).
Recently, McBeath and co-workers149 analyzed both cross polarization (CP) and direct
polarization (DP) spectra for chestnut wood-derived biochars pyrolyzed at different peak
temperatures. Results from the work of McBeath and co-workers indicated that aromaticity of
biochar rapidly increases when peak temperature is above 400 C. In addition, the authors also
reported that proportion of aromatic C detected was similar for both CP and DP techniques for all
charcoals.

6. CONCLUSIONS
The present review highlights the need for greater collaboration among researchers working in
different fields of study: production and characterization of biochar on one hand, and on the other,
measurement of both environmental and agronomical benefits linked to the addition of biochar to
agricultural soils. In this sense, when experimental results concerning the effect of the addition of
biochar to a given soil on crop yields and/or soil properties are published, details about the
properties of the used biochar should be well reported. These details include the biomass feedstock
and its composition (elemental and proximate analysis, holocellulose and lignin contents, and
mineral matter characterization), the process chosen for the biochar production and the detailed
operating conditions of which (peak temperature, soaking time, heating rate...), and information
concerning the properties of the used biochar (ultimate and proximate analysis, specific surface area,
pore size distribution, organic character...). The inclusion of this valuable information seems to be
essential in order to establish the appropriate process conditions to produce a biochar with more
suitable characteristics.
29

ACS Paragon Plus Environment

Environmental Science & Technology

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Page 30 of 56

In addition to the general consideration outlined above, several research gaps and issues have
been identified through this literature review. These research priorities are listed below:

Among the operating conditions of the slow pyrolysis process, the peak temperature
seems to be the most important parameter affecting the characteristics of biochar
product. An increase of peak temperature seems to lead to the generation of biochars
with higher aromatic character and fixed carbon and higher porosity. At the current state
of the art, this fact seems to be positive regarding the stability of the carbon in the
biochar and the enhancement of nutrient retention of a given biochar-amended soil.
Further studies analyzing the effect of pyrolysis peak temperature on both biochar
stability and nutrient retention (CEC) are required to confirm this preliminary trend.

Although slow pyrolysis or carbonization is the process commonly used to produce


biochar, because of the high charcoal yields obtained, other technologies cannot be
underestimated. In this sense, in-situ catalytic fast pyrolysis can be an interesting option
to simultaneously produce a bio-oil with enhanced properties and a biochar at an
acceptable yield. On the other hand, developing innovative process, such as the flash
carbonization process, would be a key priority for the research community in order to
improve both the productivity and the quality (fixed carbon yield) of the produced
biochar.

The specific surface area and the micropore volume of a given biochar obtained after
pyrolysis can be substantially enhanced through an activation step. This secondary
activation process can be a gasification step (physical activation by using an oxidizing
agent at a final temperature of 700850 C) or an additional carbonization step (under an
inert atmosphere at a temperature of 8501000 C). In both cases (but especially in the
physical activation process), the benefit of improving biochar porosity (and,
30

ACS Paragon Plus Environment

Page 31 of 56

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Environmental Science & Technology

consequently, the potential of the biochar to improve the soil water retention and soil
aeration) is accompanied by a loss of carbon retention and sequestration capacity. For
this reason, further investigations would be required to reach a compromise between the
desired textural properties and the carbon sequestration potential for a biochar obtained
from a given biomass feedstock.

Despite the fact of that the form of carbon (aromatic or nonaromatic C) present in
biochar is believed to be related to the stability of this material on soil, the influence of
additional properties (physical and chemical) on the stability of the biochar placed in
soil remains still unclear and further studies, in which the effect of environmental
conditions (i.e., water regime) on biochar stability can be measured, are needed. As
mentioned before, the properties of the tested biochars must be reported in these studies.

Very little information is now available regarding the influence of both biochar
properties and pyrolysis conditions on plant yield. Consequently, further research
studies, at the field scale, focused on analyzing the effect of a given biochar, obtained
under a given set of operating conditions, on the biomass yield of a given plant in a
given type of soil will be crucial to gain knowledge on this topic.

ACKNOWLEDGEMENTS
The author would like to thank Prof. Clara Mart for useful remarks and comments in the field of
soil science. The author also wishes to thank reviewer #2 for his detailed comments and helpful
suggestions aimed at improving the paper.

31

ACS Paragon Plus Environment

Environmental Science & Technology

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Page 32 of 56

REFERENCES
[1] Lehmann, J.; Joseph, S. Biochar for environmental management: an introduction. In Biochar for
environmental management: science and technology; Lehmann J, Joseph S, Eds.; Earthscan:
London 2009; pp 110.
[2] Sohi, S.; et al. Biochar, climate change and soil: a review to guide future research; CSIRO Land
and Water Science Report 05/09: 156; 2009; http://www.csiro.au/files/files/poei.pdf.
[3] Yoshida, T.; Antal, M. J. Sewage sludge carbonization for terra preta applications. Energy Fuels

2009, 23, 54545459.


[4] Lehmann, J. A handful of carbon. Nature 2007, 447, 143144.
[5] Glaser, B.; Lehmann, J.; Zech, W. Ameliorating physical and chemical properties of highly
weathered soils in the tropics with charcoal a review. Biol. Fertil Soils 2002, 35, 219230.
[6] Laird, A. D. The charcoal vision: a win-win-win scenario for simultaneously producing
bioenergy, permanently sequestering carbon, while improving soil and water quality. Agron J. 2008,
100, 178181.
[7] Fowles, M. Black carbon sequestration as an alternative to bioenergy. Biomass Bioenergy 2007,
31, 426432.
[8] Gaunt, J. L.; Lehmann, J. Energy balance and emissions associated with biochar sequestration
and pyrolysis bioenergy production. Environ. Sci. Technol. 2008, 42, 41524158.
[9] Di Blasi, C.; Signorelli, G.; Di Russo, C.; Rea, G. Product distribution from pyrolysis of wood
and agricultural residues. Ind. Eng. Chem. Res. 1999, 38, 22162224.
[10] Many, J. J.; Ruiz, J.; Arauzo, J. Some peculiarities of conventional pyrolysis of several
agricultural residues in a packed bed reactor. Ind. Eng. Chem. Res. 2007, 46, 90619070.
[11] Brewer, C. E.; Schmidt-Rohr, K.; Satrio, J. A.; Brown, R. C. Characterization of biochar from
fast pyrolysis and gasification systems. Environ. Prog. Sustainable Energy 2009, 28, 386396.
32

ACS Paragon Plus Environment

Page 33 of 56

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Environmental Science & Technology

[12] Snchez, M. E.; Lindao, E.; Margaleff, D.; Martnez, O.; Morn, A. Pyrolysis of agricultural
residues from rape and sunflowers: Production and characterization of bio-fuels and biochar soil
management. J. Anal. Appl. Pyrolysis 2009, 85, 142144.
[13] Keiluweit, M.; Nico, P. S.; Johnson, M. G.; Kleber, M. Dynamic molecular structure of plant
biomass-derived black carbon (biochar). Environ. Sci. Technol. 2010, 44, 12471253.
[14] Hammes, K.; Smernik, R. J.; Skjemstad, J. O.; Schmidt, M. W. I. Characterisation and
evaluation of reference materials for black carbon analysis using elemental composition, colour,
BET surface area and 13C NMR spectroscopy. Appl. Geochem. 2008, 23, 21132122.
[15] Swift, R. S. Sequestration of carbon by soil. Soil Sci. 2001, 166, 858871.
[16] Nguyen, B. T.; Lehmann, J.; Kinyangi, J.; Smernik, R.; Riha, S.; Engelhard, M. H. Long-term
black carbon dynamics in cultivated soil. Biogeochemistry 2009, 92, 163176.
[17] McHenry, M. P. Agricultural bio-char production, renewable energy generation and farm
carbon sequestration in Western Australia: certainty, uncertainty and risk. Agric. Ecosyst. Environ.

2009, 129, 17.


[18] Zhang, L.; Chunbao, X.; Champagne, P. Overview of recent advances in thermo-chemical
conversion of biomass. Energy Conv. Manage. 2010, 51, 969982.
[19] Antal, M. J.; Gronli, M. The art, science, and technology of charcoal production. Ind. Eng.
Chem. Res. 2003, 42, 16191640.
[20] Antal, M. J.; Croiset, E.; Dai, X.; DeAlmeida, C.; Mok, W. S.; Niclas, N.; et al. High-yield
biomass charcoal. Energy Fuels 1996, 10, 652658.
[21] Antal, M. J.; Mok, W. S.; Varhegyi, G.; Szekely, T. Review of methods for improving the yield
of charcoal from biomass. Energy Fuels 1990, 4, 221225.
[22] Raveendran, K.; Ganesh, A.; Khilar, K. C. Pyrolysis characteristics of biomass and biomass
components. Fuel 1996, 75, 987997.
33

ACS Paragon Plus Environment

Environmental Science & Technology

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Page 34 of 56

[23] Varhegyi, G.; Szabo, P.; Till, F.; Zelei, B.; Antal, M. J.; Dai, X. TG, TG-MS, and FTIR
characterization of high-yield biomass charcoals. Energy Fuels 1998, 12, 969974.
[24] Demirbas, A. Carbonization ranking of selected biomass for charcoal, liquid and gaseous
products. Energy Conv. Manage. 2001, 42, 12291238.
[25] Demirbas, A. Effects of temperature and particle size on bio-char yield from pyrolysis of
agricultural residues. J. Anal. Appl. Pyrolysis 2004, 72, 243248.
[26] Schenkel, Y.; Bertaux, P.; Vanwijnbserghe, S.; Carre, J. An evaluation of the mound kiln
carbonization technique. Biomass Bioenergy 1998, 14, 505516.
[27] Antal, M. J.; Allen, S. G.; Dai, X.; Shimizu, B.; Tam, M. S.; Gronli, M. Attainment of the
theoretical yield of carbon from biomass. Ind. Eng. Chem. Res. 2000, 39, 40244031.
[28] Khalil, L. B. Porosity characteristics of chars derived from different lignocellulosic materials.
Adsorpt. Sci. Technol. 1999, 17, 729739.
[29] MacKay, D. M.; Roberts, P. V. The influence of pyrolysis conditions on yield and
microporosity of lignocellulosic chars. Carbon 1982, 20, 95104.
[30] Dai, X.; Antal, M. J. Synthesis of a high-yield activated carbon by air gasification of
macadamia nut shell charcoal. Ind. Eng. Chem. Res. 1999, 38, 33863395.
[31] Shim, H. S.; Hurt, R. H. Thermal annealing of chars from diverse organic precursors under
combustion-like conditions. Energy Fuels 2000, 14, 340348.
[32] Stenseng, M.; Jensen, A.; Dam-Johansen, K. Investigation of biomass pyrolysis by
thermogravimetric analysis and differential scanning calorimetry. J. Anal. Appl. Pyrolysis 2001, 58
59, 765780.
[33] Cetin, E.; Moghtaderi, B.; Gupta, R.; Wall, T. F. Influence of pyrolysis conditions on the
structure and gasification reactivity of biomass chars. Fuel 2004, 83, 21392150.

34

ACS Paragon Plus Environment

Page 35 of 56

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Environmental Science & Technology

[34] Melligan, F.; Auccaise, R.; Novotny, E. H.; Leahy, J. J.; Hayes, M. H. B.; Kwapinski, W.
Pressurised pyrolysis of Miscanthus using a fixed bed reactor. Bioresour. Technol. 2011, 102, 3466
3470.
[35] Vrhegyi, G.; Szab, P.; Mok, W. S.; Antal, M. J. Kinetics of the thermal decomposition of
cellulose in sealed vessels at elevated pressures. Effects of the presence of water on the reaction
mechanism. J. Anal. Appl. Pyrolysis 1993, 26, 159174.
[36] Mok, W. S.; Antal, M. J.; Szabo, P.; Varhegyi, G.; Zelei, B. Formation of charcoal from
biomass in a sealed reactor. Ind. Eng. Chem. Res. 1992, 31, 11621166.
[37] Di Blasi, C.; Branca, C.; Santoro, A.; Hernandez, E. G. Pyrolytic behavior and products of
some wood varieties. Combust. Flame 2001, 124, 165177.
[38] Yaman, S. Pyrolysis of biomass to produce fuels and chemical feedstocks. Energy Conv.
Manage. 2004, 45, 651671.
[39] Teng, H.; Wei, Y. C. Thermogravimetric studies on the kinetics of rice hull pyrolysis and the
influence of water treatment. Ind. Eng. Chem. Res. 1998, 37, 38063811.
[40] Jensen, P. A.; Frandsen, F. J.; Dam-Johansen, K.; Sander, B. Experimental investigation of the
transformation and release to gas phase of potassium and chlorine during straw pyrolysis. Energy
Fuels 2000, 14, 12801285.
[41] Many, J. J.; Velo, E.; Puigjaner, L. Kinetics of biomass pyrolysis: A reformulated threeparallel-reactions model. Ind. Eng. Chem. Res. 2003, 42, 434441.
[42] Meszaros, E.; Jakab, E.; Varhegyi, G.; Szepesvary, P.; Marosvolgyi, B. Comparative study of
the thermal behavior of wood and bark of young shoots obtained from an energy plantation. J. Anal.
Appl. Pyrolysis 2004, 72, 317328.
[43] Gomez, C. J.; Many, J. J.; Velo, E.; Puigjaner, L. Further applications of a revisited
summative model for kinetics of biomass pyrolysis. Ind. Eng. Chem. Res. 2004, 43. 901906.
35

ACS Paragon Plus Environment

Environmental Science & Technology

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Page 36 of 56

[44] Tsai, W. T.; Chang, C. Y.; Lee, S. L. Preparation and characterization of activated carbons from
corn cob. Carbon 1997, 35, 11981200.
[45] Tsai, W. T.; Chang, C. Y.; Lee, S. L. A low cost adsorbent from agricultural waste corn cob by
zinc chloride activation. Bioresour. Technol. 1998, 64, 211217.
[46] Antal, M. J.; Croiset, E.; Dai, X.; DeAlmeida, C.; Mok, W. S.; Norberg, N. High-yield biomass
charcoal. Energy Fuels 1996, 10, 652658.
[47] Karaosmanoglu, F.; Isigigur-Ergundenler, A.; Sever, A. Biochar from the straw-stalk of
rapeseed plant. Energy Fuels 2000, 14, 336339.
[48] Hardle, W.; Simar, L. Applied multivariate statistical analysis; statistical series; Springer: New
York, 2003.
[49] Maschio, G.; Koufopanos, C.; Lucchesi, A. Pyrolysis, a promising route for biomass utilization.
Bioresour. Technol. 1992, 42, 219231.
[50] Yanik, J.; Kornmayer, C.; Saglam, M.; Yksel, M. Fast pyrolysis of agricultural wastes:
characterization of pyrolysis products. Fuel Process. Technol. 2007, 88, 942947.
[51] Onay, .; Beis, S. H.; Kokar, . M. Fast pyrolysis of rape seed in a well-swept fixed bed
reactor. J. Anal. Appl. Pyrolysis 2001, 5859, 9951007.
[52] Uzun, B. B.; Ptn, A. E.; Ptn, E. Composition of products obtained via fast pyrolysis of
olive-oil residue: effect of pyrolysis temperature. J. Anal. Appl. Pyrolysis 2007, 79, 147153.
[53] Duman, G.; Okutuku, C.; Ucar, S.; Stahl, R.; Yanik, J. The slow and fast pyrolysis of cherry
seed. Bioresour. Technol. 2011, 102, 18691878.
[54] Boateng, A. A. Characterization and thermal conversion of charcoal derived from fluidized-bed
fast pyrolysis oil production of switchgrass. Ind. Eng. Chem. Res. 2007, 46, 88578862.

36

ACS Paragon Plus Environment

Page 37 of 56

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Environmental Science & Technology

[55] Mullen, C. A.; Boateng, A. A.; Goldberg, N. M.; Lima, I. M.; Laird, D. A.; Hicks, K. B. Bio-oil
and bio-char production from corn cobs and stover by fast pyrolysis. Biomass Bioenergy 2010, 34,
6774.
[56] Day, D.; Evans, R. J.; Lee, J. W.; Reicosky, D. Economical CO2, SO2, and NOx capture from
fossil-fuel utilization with combined renewable hydrogen production and large-scale carbon
sequestration. Energy 2005, 30, 25582579.
[57] Scala, F.; Chirone, R.; Salatino, P. Combustion and attrition of biomass chars in a fluidized bed.
Energy Fuels 2006, 20, 91102.
[58] Czernik, S.; Bridgwater, A. V. Overview of applications of biomass fast pyrolysis oil. Energy
Fuels 2004, 18, 590598.
[59] Mohan, D.; Pittman, C. U.; Steele, P. H. Pyrolysis of wood/biomass for bio-oil: a critical
review. Energy Fuels 2006, 20, 848889.
[60] Nowakowski, D. J.; Jones, J. M.; Brydson, R. M. D.; Ross, A. B. Potassium catalysis in the
pyrolysis behaviour of short rotation willow coppice. Fuel 2007, 86, 23892402.
[61] Nowakowski, D. J.; Jones, J. M. Uncatalysed and potassium-catalysed pyrolysis of the cell-wall
constituents of biomass and their model compounds. J. Anal. Appl. Pyrolysis 2008, 83, 1225.
[62] Raveendran, K.; Ganesh, A.; Khilar, K. C. Influence of mineral matter on biomass pyrolysis
characteristics. Fuel 1995, 74, 18121822.
[63] Di Blasi, C.; Galgano, A.; Branca, C. Effects of potassium hydroxide impregnation on wood
pyrolysis. Energy Fuels 2009, 23, 10451054.
[64] Shimada, N.; Kawamoto, H.; Saka, S. Different action of alkali/alkaline earth metal chlorides
on cellulose pyrolysis. J. Anal. Appl. Pyrolysis 2008, 81, 8087.

37

ACS Paragon Plus Environment

Environmental Science & Technology

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Page 38 of 56

[65] Wang, Z.; Wang, F.; Cao, J; Wang, J. Pyrolysis of pine wood in a slowly heating fixed-bed
reactor: Potassium carbonate versus calcium hydroxide as a catalyst. Fuel Process. Technol. 2010,
91, 942950.
[66] Adam, J.; Antonakou, E.; Lappas, A.; Stcker, M.; Nilsen, M. H.; Bouzga, A.; et al. In situ
catalytic upgrading of biomass derived fast pyrolysis vapours in a fixed bed reactor using
mesoporous materials. Microporous Mesoporous Mater. 2006, 96, 93101.
[67] Torri, C.; Reinikainen, M.; Lindfors, C.; Fabbri, D.; Oasmaa, A.; Kuoppala, E. Investigation on
catalytic pyrolysis of pine sawdust: catalyst screening by Py-GC-MIP-AED. J. Anal. Appl. Pyrolysis

2010, 88, 713.


[68] Zhang, H.; Xiao, R.; Huang, H.; Xiao, G. Comparison of non-catalytic and catalytic fast
pyrolysis of corncob in a fluidized bed reactor. Bioresour. Technol. 2009, 100, 14281434.
[69] Pattiya, A.; Titiloye, J. O.; Bridgwater, A. V. Fast pyrolysis of cassava rhizome in the presence
of catalysts. J. Anal. Appl. Pyrolysis 2008, 81, 7279.
[70] Nunoura, T.; Wade, S. R.; Bourke, J.; Antal, M. J. Studies of the flash carbonization process. 1.
propagation of the flaming pyrolysis reaction and performance of a catalytic afterburner. Ind. Eng.
Chem. Res. 2006, 45, 585599.
[71] Wade, S. R.; Nunoura, T.; Antal, M. J. Studies of the flash carbonization process. 2. violent
ignition behavior of pressurized packed beds of biomass: a factorial study. Ind. Eng. Chem. Res.

2006, 45, 35123519.


[72] Antal, M. J.; Mochidzuki, K.; Paredes, L. S. Flash carbonization of biomass. Ind. Eng. Chem.
Res. 2003, 42, 36903699.
[73] Reynolds, W. C. The element potential method for chemical equilibrium analysis:
implementation in the interactive program STANJAN version 3. Stanford University, 1989;

38

ACS Paragon Plus Environment

Page 39 of 56

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Environmental Science & Technology

http://www.stanford.edu/~cantwell/AA283_Course_Material/STANJAN_writeup_by_Bill_Reynolds.pdf.
[74] Pan, Y. G.; Velo, E.; Roca, X.; Many, J. J.; Puigjaner, L. Fluidized-bed co-gasification of
residual biomass. Fuel 2000, 79, 13171326.
[75] Rezaiyan, J.; Cheremisinoff, N. P. Gasification technologies a primer for engineers and
scientists; CRC Press Taylor & Francis Groups: Boca Raton, FL, 2005.
[76] IEA bioenergy annual report; International Energy Agency, 2006.
[77] Fernandes, M. B.; Brooks, P. Characterization of carbonaceous combustion residues: II.
nonpolar organic compounds. Chemosphere 2003, 53, 447458.
[78] Ioannidou, O.; Zabaniotou, A. Agricultural residues as precursors for activated carbon
productionA review. Renewable Sustainable Energy Rev. 2007, 11, 19662005.
[79] Haykiri-Acma, H.; Yaman, S.; Kucukbayrak, S. Gasification of biomass chars in steam
nitrogen mixture. Energy Convers. Manage. 2006, 47, 10041013.
[80] Zhang, T.; Walawender, W. P.; Fan, L. T.; Fan, M.; Daugaard, D.; Brown, R. C. Preparation of
activated carbon from forest and agricultural residues through CO2 activation. Chem. Eng. J. 2004,
105, 5359.
[81] Lua, A. C.; Yang, T.; Guo, J. Effects of pyrolysis conditions on the properties of activated
carbons prepared from pistachio-nut shells. J. Anal. Appl. Pyrolysis 2004, 72, 279287.
[82] Baaoui, A.; Yaacoubi, A.; Dahbi, A.; Bennouna, C.; Phan Tan Luu, R.; Maldonado-Hodar, F.
J.; et al. Optimization of conditions for the preparation of activated carbons from olive-waste cakes.
Carbon 2001, 39, 425432.
[83] Nowicki, P.; Pietrzak, R. Carbonaceous adsorbents prepared by physical activation of pine
sawdust and their application for removal of NO2 in dry and wet conditions. Bioresour. Technol.

2010, 101, 58025807.


39

ACS Paragon Plus Environment

Environmental Science & Technology

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Page 40 of 56

[84] Plaza, M. G.; Pevida, C.; Arias, B.; Fermoso, J.; Casal, M. D.; Martn, C. F.; et al. Development
of low-cost biomass-based adsorbents for postcombustion CO2 capture. Fuel 2009, 88, 24422447.
[85] Skodras, G.; Diamantopoulou, I.; Zabaniotou, A.; Stavropoulos, G.; Sakellaropoulos, G. P.
Enhanced mercury adsorption in activated carbons from biomass materials and waste tires. Fuel
Process. Technol. 2007, 88, 749758.
[86] Demiral, H.; Demiral, .; Tmsek, F.; Karabacakolu, B. Adsorption of chromium (VI) from
aqueous solution by activated carbon derived from olive bagasse and applicability of different
adsorption models. Chem. Eng. J. 2008, 144, 188196.
[87] Lua, A. C.; Jia, Q. Adsorption of phenol by oilpalm-shell activated carbons in a fixed bed.
Chem. Eng. J. 2009, 150, 455461.
[88] Gregg, S. J.; Sing, K. S. W. Adsoption, surface area and porosity; Academic Press: London,
U.K., 1982.
[89] Downie, A.; Crosky, A.; Munroe, P. Physical properties of biochar. In Biochar for
Environmental Management: Science and Technology; Lehmann, J., Joseph, S., Eds.; Earthscan:
London, U.K., 2009; pp 1329.
[90] Verheijen, F.; Jeffery, S.; Bastos, A. C.; van der Velde, M.; Diafas, I. Biochar application to
soils. A critical scientific review of effects on soil properties, processes and functions; JRC
Scientific

and

Technical

Reports;

European

Commission:

Luxembourg,

2010;

http://publications.jrc.ec.europa.eu/repository/bitstream/111111111/13558/1/jrc_biochar_soils.pdf.
[91] Chan, K. Y.; Van Zwieten, L.; Meszaros, I.; Downie, A.; Joseph. S. Agronomic values of
greenwaste biochar as a soil amendment. Aust. J. Soil Res. 2007, 45, 629634.
[92] Kolb, S. E.; Fermanich, K. J.; Dornbush, M. E. Effect of charcoal quantity on microbial
biomass and activity in temperate soils. Soil Sci. Soc. Am. J. 2007, 73, 11731181.

40

ACS Paragon Plus Environment

Page 41 of 56

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Environmental Science & Technology

[93] Laird, D. A.; Fleming, P.; Davis, D. D.; Horton, R.; Wang, B.; Karlen, D. L. Impact of biochar
amendments on the quality of a typical Midwestern agricultural soil. Geoderma 2010, 158, 443449.
[94] Yuan, J.; Xu, R.; Zhang, H. The forms of alkalis in the biochar produced from crop residues at
different temperatures. Bioresour. Technol. 2011, 102, 34883497.
[95] Laird, D.; Flora, G.; Wang, B.; Horton, R.; Karlen, D. Biochar impact on nutrient leaching from
a Midwestern agricultural soil. Geoderma 2010, 158, 436442.
[96] Van Zwieten, L.; Kimber, S.; Morris, S.; Chan, K. Y.; Downie, A.; Rust, J.; et al. Effects of
biochar from slow pyrolysis of papermill waste on agronomic performance and soil fertility. Plant
Soil 2010, 327, 235246.
[97] Hossain, M. K.; Strezov, V.; Chan, K. Y.; Ziolkowski, A.; Nelson, P. F. Influence of pyrolysis
temperature on production and nutrient properties of wastewater sludge biochar. J. Environ.
Manage. 2011, 92, 223228.
[98] Lehmann, J.; Czimczik, C.; Laird, D.; Sohi, S. Stability of biochar in the soil. In Biochar for
Environmental Management: Science and Technology; Lehmann, J., Joseph, S., Eds.; Earthscan:
London, U.K., 2009, pp 183206.
[99] Preston, C.M.; Schmidt, M. W. I. Black (pyrogenic) carbon in boreal forests: a synthesis of
current knowledge and uncertainties. Biogeosci. Discuss 2006, 3, 211271.
[100] Kawamoto, K.; Ishimaru, K.; Imamura, Y. Reactivity of wood charcoal with ozone. J. Wood
Sci. 2005, 51, 6672.
[101] Cheng, C. H.; Lehmann, J.; Thies, J.; Burton, S. D.; Engelhard, M. H. Oxidation of black
carbon by biotic and abiotic processes. Org. Geochem. 2006, 37, 14771488.
[102] Hamer, U.; Marschner, B.; Brodowski, S.; Amelung, W. Interactive priming of black carbon
and glucose mineralisation. Org. Geochem. 2004, 35, 823830.

41

ACS Paragon Plus Environment

Environmental Science & Technology

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Page 42 of 56

[103] Lehmann, J.; Lan, Z.; Hyland, C.; Sato, S.; Solomon, D.; Ketterings, Q. M. Long term
dynamics of phosphorus and retention in manure amended soils. Environ. Sci. Technol. 2005, 39,
66726680.
[104] Brodowski, S.; Amelung, W.; Haumaier, L.; Abetz, C.; Zech, W. Morphological and chemical
properties of black carbon in physical soil fractions as revealed by scanning electron microscopy and
energy-dispersive X-ray spectroscopy. Geoderma 2005, 128, 116129.
[105] Baldock, J. A.; Smernik, R. J. Chemical composition and bioavailability of thermally altered
Pinus resinosa (Red pine) wood. Org. Geochem. 2002, 33, 10931109.
[106] Nguyen, B. T.; Lehmann, J.; Hockaday, W. C., Joseph, S., Masiello, C. A. Temperature
sensitivity of black carbon decomposition and oxidation. Environ. Sci. Technol. 2010, 44, 3324
3331.
[107] Qayyum, M. F.; Steffens, D.; Reisenauer, H. P.; Schubert, S. Kinetics of carbon
mineralization of biochars compared with wheat straw in three soils. J. Environ. Qual. 2011, in
press; DOI 10.2134/jeq2011.0058.
[108] Yanai, Y.; Toyota, K.; Okazaki, M. Effect of charcoal addition on N2O emissions from soil
resulting from rewetting air-dried soil in short-term laboratory experiments. Soil Sci. Plant Nutr.

2007, 53, 181188.


[109] Zhang, A.; Cui, L.; Pan, G.; Li, L.; Hussain, Q.; Zhang, X.; et al. Effect of biochar amendment
on yield and methane and nitrous oxide emissions from a rice paddy from Tai Lake plain, China.
Agric. Ecosyst. Environ. 2010, 139, 469475.
[110] Sarkhot, D. V.; Berhe, A. A.; Ghezzehei, T. A. Impact of biochar enriched with dairy manure
effluent on carbon and nitrogen dynamics. J. Environ. Qual. 2011, in press; DOI
10.2134/jeq2011.0123.

42

ACS Paragon Plus Environment

Page 43 of 56

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Environmental Science & Technology

[111] Xiong, Z.; Xing, G.; Zhu, Z. Nitrous oxide and methane emissions as affected by water, soil
and nitrogen. Pedosphere 2007, 17, 146155.
[112] Rondon, M. A.; Molina, D.; Hurtado, M.; Ramirez, J.; Lehmann, J.; Major, J., et al. Enhancing
the productivity of crops and grasses while reducing greenhouse gas emissions through biochar
amendments to unfertile tropical soils. In Proceedings of the 18th World Congress of Soil Science;
Philadelphia, PA, 2006; pp 138168.
[113] Van Zwieten, L.; Singh, B.; Joseph, S.; Kimber, S.; Cowie, A.; Chan, K. Y. Biochar and
emissions of non-CO2 greenhouse gases from soil. In Biochar for Environmental Management
Science and Technology; Earthscan Press: London, U.K., 2009; pp 227249.
[114] Cornelissen, G.; Gustafsson, .; Bucheli, T. D.; Jonker, M. T. O.; Koelmans, A. A.; van
Noort, P. C. M. Extensive sorption of organic compounds to black carbon, coal and kerogen in
sediments and soils: mechanisms and consequences for distribution, bioaccumulation and
biodegradation. Environ. Sci. Technol. 2005, 39, 68816895.
[115] Zhu, D.; Kwon, S.; Pignatello, J. J. Adsorption of single-ring organic compounds to wood
charcoals prepared to under different thermochemical conditions. Environ. Sci. Technol. 2005, 39,
39903398.
[116] Wang, X.; Sato, T.; Xing, B. Competitive sorption of pyrene on wood chars. Environ. Sci.
Technol. 2006, 40, 32673272.
[117] Chen, J.; Zhu, D.; Sun, C. Effect of heavy metals on the sorption of hydrophobic organic
compounds to wood charcoal. Environ. Sci. Technol. 2007, 41, 25362541.
[118] Chun, Y.; Sheng, G. Y.; Chiou, C. T.; Xing B. Compositions and sorptive properties of crop
residue-derived chars. Environ. Sci. Technol. 2004, 38, 46494655.

43

ACS Paragon Plus Environment

Environmental Science & Technology

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Page 44 of 56

[119] Kinney, T. J.; Masiello, C. A.; Dugan, B.; Hockaday, W. C.; Dean, M. R.; Zygourakis, K.;
Barnes, R. T. Hydrologic properties of biochars produced at different temperatures. Biomass
Bioenergy 2012, 41, 3443.
[120] Fytili, D.; Zabaniotou, A. Utilization of sewage sludge in EU application of old and new
methodsa review. Renewable Sustainable Energy Rev. 2008, 12, 116140.
[121] Ledesma, E. B.; Marsh, N. D.; Sandrowitz, A. K.; Wornat, M. J. Global kinetic rate
parameters for the formation of polycyclic aromatic hydrocarbons from the pyrolyis of
catechol, a model compound representative of solid fuel moieties. Energy Fuels 2002, 16, 1331
1336.
[122] Conesa, J. A.; Font, R.; Fullana, A.; Martin-Gullon, I.; Aracil, I.; Galvez, A.; et al.
Comparison between emissions from the pyrolysis and combustion of different wastes. J. Anal.
Appl. Pyrolysis 2009, 84, 95102.
[123] Brown, R. A.; Kercher, A. K.; Nguyen, T. H.; Nagle, D. C.; Ball, W. P. Production and
characterization of synthetic wood chars for use as surrogates for natural sorbents. Org. Geochem.

2006, 37, 321333.


[124] Garrity, D. P. Agroforestry and the achievement of the millenium development goals.
Agroforest. Syst. 2004, 61, 517.
[125] Major, J.; Rondon, M.; Molina, D.; Riha, S. J., Lehmann, J. Maize yield and nutrition during 4
years after biochar application to a Colombian savanna oxisol. Plant Soil 2010, 333, 117128.
[126] Lehmann, J.; Da Silva, J. P.; Steiner, C.; Nehls, T.; Zech, W.; Glaser, B. Nutrient availability
and leaching in an archaeological Anthrosol and a Ferralsol of the Central Amazon basin: fertilizer,
manure and charcoal amendments. Plant Soil 2003, 249, 343357.

44

ACS Paragon Plus Environment

Page 45 of 56

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Environmental Science & Technology

[127] Hass, A.; Gonzalez, J. M.; Lima, I. M.; Godwin, H. W.; Halvorson, J. J.; Boyer, D. G.
Chicken manure biochar as liming and nutrient source for acid Appalachian soil. J. Environ. Qual.
2011, in press; DOI 10.2134/jeq2011.0124.
[128] Rondon, M.; Lehmann, J.; Ramirez, J.; Hurtado, M. Biological nitrogen fixation by common
beans (Phaseolus vulgaris L.) increases with bio-char additions. Biol. Fertil. Soils 2007, 43, 699
708.
[129] Steiner, C.; Glaser, B.; Teixeira, W. G.; Lehmann, J.; Blum, W. E. H.; Zech, W. Nitrogen
retention and plant uptake on a highly weathered central Amazonian Ferralsol amended with
compost and charcoal. J. Plant Nutr. Soil Sci. 2008, 171, 893899.
[130] Liang, B.; Lehmann, J.; Solomon, D.; Sohi, S.; Thies, J. E.; Skjemstad, J. O.; et al. Stability of
biomass-derived black carbon in soils. Geochim. Cosmochim. Acta 2008, 72, 60696078.
[131] Glaser, B.; Haumaier, L.; Guggenberger, G.; Zech, W. The 'Terra Preta' phenomenon: a model
for sustainable agriculture in the humid tropics. Naturwissenschaften 2001, 88, 3741.
[132] Steiner, C.; Teixeira, W. G.; Lehmann, J.; Nehls, T.; MacDo, J. L. V.; Blum, W. E. H.; et al.
Long term effects of manure, charcoal and mineral fertilization on crop production and fertility on a
highly weathered Central Amazonian upland soil. Plant Soil 2007, 291, 275290.
[133] Haefele, S. M.; Konboon, Y.; Wongboon, W.; Amarante, S.; Maarifat, A. A.; Pfeiffer, E. M.;
et al. Effects and fate of biochar from rice residues in rice-based systems. Field Crops Res. 2011,
121, 430440.
[134] Hossain, M. K.; Strezov, V.; Yin Chan, K.; Nelson, P. F. Agronomic properties of wastewater
sludge biochar and bioavailability of metals in production of cherry tomato (Lycopersicon
esculentum). Chemosphere 2010, 78, 11671171.

45

ACS Paragon Plus Environment

Environmental Science & Technology

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Page 46 of 56

[135] Peng, X.; Ye, L.; Wang, C. H.; Zhou, H.; Sun, B. Temperature- and duration-dependent rice
straw-derived biochar: Characteristics and its effects on soil properties of an Ultisol in southern
China. Soil Tillage Res. 2011, 112, 159166.
[136] Vaccari, F. P.; Baronti, S.; Lugato, E.; Genesio, L.; Castaldi, S.; Fornasier, F.; et al. Biochar as
a strategy to sequester carbon and increase yield in durum wheat. Eur. J. Agron. 2011, 34, 231238.
[137] Deenik, J. L.; McClellan, T.; Uehara, G.; Antal, M. J.; Campbell, S. Charcoal volatile matter
content influences plant growth and soil nitrogen transformations. Soil Sci. Soc. Am. J. 2010, 74,
12591270.
[138] Standardized product definition and product testing guidelines for biochar that is used in soil;
International Biochar Initiative, 2012; http://www.biocharinternational.org/sites/default/files/Guidelines_for_Biochar_That_Is_Used_in_Soil_Final.pdf.
[139] Bahng, M.; Mukarakate, C.; Robichaud, D. J.; Nimlos, M. R. Current technologies for
analysis of biomass thermochemical processing: a review. Anal. Chim. Acta 2009, 651, 117138.
[140] Surez-Garca, F.; Martnez-Alonso, A.; Fernndez-Llorente, M.; Tascn, J. M. D. Inorganic
matter characterization in vegetable biomass feedstocks. Fuel 2002, 81, 11611169.
[141] Liao, C.; Wu, C.; Yan, Y. The characteristics of inorganic elements in ashes from a 1 MW
CFB biomass gasification power generation plant. Fuel Process. Technol. 2007, 88, 149156.
[142] Chan, K. Y.; Xu, Z. Biochar: nutrient properties and their enhancement. In Biochar for
environmental management: science and technology; Lehmann, J., Joseph, S., Eds.; Earthscan:
London, U.K., 2009.
[143] Joseph, S.; Peacock, C.; Lehmann, J.; Munroe, P. Developing a biochar classification and test
methods. In Biochar for environmental management: science and technology; Lehmann, J., Joseph,
S., Eds.; Earthscan: London, U.K., 2009.

46

ACS Paragon Plus Environment

Page 47 of 56

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Environmental Science & Technology

[144] Gonzlez, J. F.; Romn, S.; Encinar, J. M.; Martnez, G. Pyrolysis of various biomass residues
and char utilization for the production of activated carbons. J. Anal. Appl. Pyrolysis 2009, 85, 134
141.
[145] Plaza, M. G.; Pevida, C.; Martn, C. F.; Fermoso, J.; Pis, J. J.; Rubiera, F. Developing almond
shell-derived activated carbons as CO2 adsorbents. Sep. Purif. Technol. 2010, 71, 102106.
[146] Sharma, R. K.; Wooten, J. B.; Baliga, V. L.; Lin, X.; Geoffrey Chan, W.; Hajaligol, M. R.
Characterization of chars from pyrolysis of lignin. Fuel 2004, 83, 14691482.
[147] Gao, Y.; Wang, X. H.; Yang, H. P.; Chen, H. P. Characterization of products from
hydrothermal treatments of cellulose. Energy 2012, 42, 457465.
[148] Freitas, J. C. C.; Bonagamba, T. J.; Emmerich, F. G. Investigation of biomass and polymerbased carbon materials using 13C high-resolution solid-state NMR. Carbon 2001, 39, 535545.
[149] McBeath, A. V.; Smernik, R. J.; Schneider, M. P. W.; Schmidt, M. W. I.; Plant, E. L.
Determination of the aromaticity and the degree of aromatic condensation of a thermosequence of
wood charcoal using NMR. Org. Geochem. 2011, 42, 11941202.
[150] Lehmann, J.; Silva, J. P.; Rondon, M.; Silva, C. M.; Greenwood, J.; Nehls, T.; et al. Slashand-char a feasible alternative for soil fertility management in the central Amazon?. In Proceedings
of 17th World Congress of Soil Science; Bangkok, Thailand, 2002.
[151] World Reference Base for Soil Resources 2006; F.A.O., Rome, Italy, 2006;
http://www.fao.org/ag/Agl/agll/wrb/doc/wrb2006final.pdf.

47

ACS Paragon Plus Environment

Environmental Science & Technology

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Page 48 of 56

Table 1. Charcoal yields obtained from slow pyrolysis of several biomass feedstocks

Reference

Biomass

Alder wood

Composition

Dried with air


Lignin: 24.8%
Ash: 1.1%

Pressure
(MPa)

0.1

1.0

Birch wood

Dried with air


Lignin: 19.3%
Ash: 1.1%

1.0

Antal et al.
(2000)27
Oak wood

Pine wood

Spruce wood

Macadamia nut
shells
Antal et al.
(1996)46

0.1

Eucalyptus
wood

Olive husk

Dried with air


Lignin: 28.0%
Ash: 1.4%
Dried with air
Lignin: 25.0%
Ash: 0.9%
Dried with air
Lignin: 28.0%
Ash: 0.7%
Dried with air
Lignin: 40.0%
Ash: 0.8%
Moisture: 16%
Lignin: 18.1%
Ash: 1.3%
Moisture: 29%
Moisture: 56.9%

0.1
1.0
0.1
1.0
0.1
1.0
0.4
1.0
3.3
1.0

Dried with air


Lignin: 48.4%
Ash: 1.4%

Demirbas (2001)24

0.1
Beech wood

Dried with air


Lignin: 21.9%
Ash: 0.4%

Heating conditions
2 K min1 up to 450 C
followed by 60 min of
soaking at 450 C
2 K min1 up to 450 C
followed by 240 min of
soaking at 450 C
6 K min1 up to 450 C
with no soaking time
2 K min1 up to 450 C
followed by 60 min of
soaking at 450 C
2 K min1 up to 450 C
followed by 240 min of
soaking at 450 C
6 K min1 up to 450 C
with no soaking time
2 K min1 up to 450 C
followed by 60 min of
soaking at 450 C
6 K min1 up to 450 C
with no soaking time
2 K min1 up to 450 C
followed by 60 min of
soaking at 450 C
6 K min1 up to 450 C
with no soaking time
2 K min1 up to 450 C
followed by 60 min of
soaking at 450 C
6 K min1 up to 450 C
with no soaking time

10 K min1 up to 450 C
followed by 15 min of
soaking at 450 C

10 K min1 up to 377 C
with no soaking time
10 K min1 up to 477 C
with no soaking time
10 K min1 up to 577 C
with no soaking time
10 K min1 up to 377 C
with no soaking time
10 K min1 up to 477 C
with no soaking time
10 K min1 up to 577 C
with no soaking time

Charcoal
yield
(% w/w)

Number
in Fig. 3

30.5

29.8

35.9

28.8

32.1

34.6

31.2

39.8

32.1

35.2

10

32.2

11

37.5

12

40.5
44.4
51.0

13
14
15

41.8

16

42.2
46.1

17
18

39.7

19

36.8

20

33.3

21

29.7

22

26.2

23

24.7

24

48

ACS Paragon Plus Environment

Page 49 of 56

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Environmental Science & Technology

Table 1 (continued)

Reference

Demirbas (2001)24

Karaosmanoglu et
al. (2000)47

Biomass

Corncob

Straw and stalks


of rapeseed plant

Composition

Dried with air


Lignin: 15.0%
Ash: 1.0%

Dried with air


Lignin: 19.3%
Ash: 5.9%

Pressure
(MPa)

0.1

0.1

Heating conditions
10 K min1 up to 377 C
with no soaking time
10 K min1 up to 477 C
with no soaking time
10 K min1 up to 577 C
with no soaking time
5 K min1 up to 400 C
with no soaking time
5 K min1 up to 500 C
with no soaking time
5 K min1 up to 600 C
with no soaking time
5 K min1 up to 700 C
with no soaking time

Charcoal
yield
(% w/w)
26.0

Number
in Fig. 3
25

23.2

26

21.5

27

39.4

28

35.6

29

32.6

30

29.6

31

49

ACS Paragon Plus Environment

Environmental Science & Technology

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Page 50 of 56

Table 2. Activation conditions and textural properties of several activated charcoals

Reference

Biomass
feedstock

Charcoal
production

Oxidizing
agent

Oak
wood

Zhang et al.
(2004)80

Corn
hulls

Fast
pyrolysis
at 500 C

CO2

Corn
stover

Plaza et al.
(2009)84

Olive
stones

Skodras et al.
(2007)85

Pine
wood
Oak
wood
Olive
seed
waste

Demiral et al.
(2008)86

Olive
bagasse

Lua and Gia


(2009)87

Oil palm
shell

Nowicki and
Pietrzak
(2010)83

Pine
sawdust
pellets

Slow
pyrolysis
up to
600 C

Fast
pyrolysis
at 800 C
Slow
pyrolysis
up to
500 C
Slow
pyrolysis
up to
400 C

Slow
pyrolysis
up to
800 C

CO2

Activation
temperature
(C)
700
700
800
800
700
700
800
800
700
700
800
800
Not
activated
800
800
800
900

H2OCO2
mixture

Burn-off
(%)

SBET
(m2 g1)

V0
(cm3 g1)

V0/V

31.8
41.8
43.0
51.4
32.3
37.0
44.4
45.2
41.5
41.7
42.6
50.2

642
644
845
985
977
902
1010
975
660
432
712
616

0.270
0.245
0.321
0.379
0.335
0.328
0.435
0.379
0.282
0.182
0.285
0.234

0.657
0.606
0.534
0.592
0.376
0.399
0.521
0.557
0.577
0.546
0.519
0.555

43

n. a.

n. a.

20.0
40.0
50.0

613
909
1079

0.242
0.364
0.436

0.840
0.833
0.868

n. a.
(activation
time: 2.5 h)

897

0.340

0.557

684

0.220

0.489

800

n. a.
(activation
time: 3 h)

1690

0.700

0.778

Steam

850

n. a.
(activation
time: 0.5 h)

718

0.315

0.851

Steam

900

n. a.
(activation
time: 1 h)

1183

0.310

0.449

140

0.092

0.948

176

0.090

0.783

800

n. a.
(activation
time: 0.5 h)
n. a.
(activation
time: 1 h)
n. a.
(activation
time: 1.5 h)
n. a.
(activation
time: 2 h)

269

0.129

0.872

352

0.178

0.918

CO2

50

ACS Paragon Plus Environment

Page 51 of 56

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Environmental Science & Technology

Table 3. Cation exchange capacity of soils amended with biochar


Biochar
(g kg1)

CEC (cmol kg1)

Available K

Available Ca

Available Mg

(cmol kg1)

(cmol kg1)

(cmol kg1)

pH

Soil: terra preta from Amazonia; biochar: from incomplete combustion of organic material5
1.4

3.20

0.239

2.58

0.330

Not reported

11.9

28.9

0.026

15.4

1.71

Not reported

Soil: ferralsol from Amazonia; biochar: from incomplete combustion of black locust wood150
0

5.40

2.81

1.48

0.88

Not reported

135

29.0

25.8

1.7

1.0

Not reported

Soil: ferralsol from Amazonia; biochar: from incomplete combustion of black locust wood126
0

5.40

2.81

1.48

0.88

5.14

100

28.5

25.8

1.71

0.97

5.89

Soil: red ferrosol (ferralsol151) from New South Gales (Australia); biochar: from slow pyrolysis of
paper mill wastes96
0

4.03

0.11

1.23

0.30

4.20

20

10.5

0.66

8.87

0.67

5.93

Soil: calcarosol (carcisol144) from Victoria (Australia); biochar: from slow pyrolysis of paper mill
wastes96
0

31.0

2.07

21.7

6.23

7.67

15

29.0

2.23

20.3

6.10

7.67

51

ACS Paragon Plus Environment

Environmental Science & Technology

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Page 52 of 56

Table 4. Summary of studies assessing the impact of biochar on crop yield


Reference
Glaser et al.
(2002)5

Steiner et al.
(2007)132

Van Zwieten at al.


(2010)96

Soil type

Biochar

Crop

A weathered
xanthic
ferralsol from
Central
Amazonia
(Brazil)
A weathered
xanthic
ferralsol from
Central
Amazonia
(Brazil)
A ferralsol
(acidic soil)
from New
South Gales
(Australia)

From secondary
forestry wood.
Addition rate:
67.2 and 135.2
Mg ha1.

Rice and cowpea

From secondary
forestry wood.
Addition rate:
11.0 Mg ha1.

Rice and sorghum

From mixtures of
paper mill wastes
and wood chips
pyrolyzed at 550
C.
Addition rate:
10.0 Mg ha1.

Wheat, soybean
and radish

Van Zwieten at al.


(2010)96

A carcisol
(alkaline soil)
from Victoria
(Australia)

From mixtures of
paper mill wastes
and wood chips
pyrolyzed at 550
C.
Addition rate:
10.0 Mg ha1.

Wheat, soybean
and radish

Major et al.
(2010)125

A clay-loam
ferralsol from
the oriental
savanna of
Colombia

Commercial
wood charcoal.
Addition rate:
20.0 Mg ha1.

Maize

Response
At a rate of 67.2 Mg ha1 biomass
increased by 20% (rice) and 50%
(cowpea) compared to control (no
biochar).
At 135.2 Mg ha1 biomass
cowpea increased by 100%.
Stover and grain yields increased
by 29% and 73%, respectively;
compared to control treatment
(only inorganic fertilizer).
Wheat: biochar improved yield
by a factor of 1.3. When biochar
was applied together to inorganic
fertilizer, yields increased by a
factor of 2.4 compared to using
fertilizer alone.
Soybean: biochar improved yield
by a factor of 1.4 in the presence
of fertilizer.
Radish: dry biomass production
was significantly increased by a
factor of 1.52 both in the
presence and absence of fertilizer.
Biochar significantly increased
both pH and CEC and reduced Al
availability.
Wheat: yields were reduced by a
factor of 2 both in the presence
and absence of inorganic
fertilizer.
Soybean: biochar improved yield
by a factor of 1.3 in the presence
of fertilizer.
Radish: biochar addition in the
absence of fertilizer increased
biomass production by a factor of
1.5. However, in the presence of
fertilizer, yields were reduced by
a factor of 2.
Effect for 4 years (20032006).
Maize grain yield did not
significantly increase in the first
year, but increases over the
control were 28, 30 and 140% for
2004, 2005 and 2006,
respectively.

52

ACS Paragon Plus Environment

Page 53 of 56

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Environmental Science & Technology

Table 4 (continued)
Reference

Soil type

Biochar

Crop

Hossain et al.
(2010)134

A luvisol from
Sydney
(Australia)

From sewage
sludge pyrolysed
at 550 C.
Addition rate:
10.0 Mg ha1.

Cherry tomato

Haefele et al.
(2011)133

An anthraquic
gleysol from
Laguna
(Philippines)

From rice husk


partially burned
in a combustion
chamber
Addition rate:
0.413 Mg ha1.

Rice

Haefele et al.
(2011)133

A humic
nitisol from
Siniloan
(Philippines)

From rice husk


partially burned
in a combustion
chamber
Addition rate:
0.413 Mg ha1.

Rice

Haefele et al.
(2011)133

A gleyic
acrisol from
Ubon
(Thailand)

From rice husk


partially burned
in a combustion
chamber
Addition rate:
0.413 Mg ha1.

Rice

Peng et al.
(2011)135

A typical
ultisol from
southern China

From rice straw


pyrolyzed at low
heating rate and
at peak
temperatures
below 450 C.
Addition rate:
240 Mg ha1.

Maize

Vaccari et al.
(2011)136

A silt loam
soil (with a
sub-acid pH of
5,2) from the
region of
Tuscany
(Italy)

A commercial
charcoal
obtained from
coppiced
woodlands
(beech, hazel,
oak, and birch).
Addition rate:
3060 Mg ha1.

Durum wheat

Response
Application of biochar improves
the production of cherry tomatoes
by 64% above the control soil
conditions.
The yield of production was
found to be at its maximum when
biochar was applied in
combination with an inorganic
fertilizer.
Effect for 4 years (20052008).
Application of carbonized rice
husks increased total organic
carbon, total soil N, the C/N ratio,
and available P and K.
Biochar application decreased
rice yields, especially in the first
few seasons after application. For
the entire period evaluated, rice
yield decreased by 2.5% relative
to control conditions.
Effect for 4 years (20052008).
Application of carbonized rice
husks increased total organic
carbon, total soil N, the C/N ratio,
and available P and K.
For the entire period evaluated,
rice yield increased by 8.9%.
Effect for 4 years (20052008).
Application of carbonized rice
husks increased total organic
carbon, total soil N, the C/N ratio,
and available P and K.
For the entire period evaluated,
rice yield increased by 8.7%.
The effected of biochar
amendment application on the
40-day maize dry matter
production was marked: in the
absence of NPK fertilizer,
biomass production increased by
64%. In the presence of fertilizer,
biomass production increased by
a factor of 3 compared to using
fertilizer alone.
Effect for 2 years (20092010).
Biochar addition significantly
increased grain production with
respect to the control. On
average, the grain yield increase
ranged from 28% to 39%. No
significant differences were
observed between biochar rate
treatments of 30 and 60 Mg ha1.

53

ACS Paragon Plus Environment

Environmental Science & Technology

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Page 54 of 56

Caption for figures


Figure 1. Score and loading plots for the three first principal components: (a), second versus first
principal component; (b); third versus second principal component.

Figure 2. Percentage of the theoretical value of yFC for four biomass feedstocks: leucaena wood
(LW), oak wood (OW), macadamia nut shells (MS), and corncob (CC). The white bars correspond
to the results obtained using a slow pyrolysis process, whereas the gray bars indicate the results
obtained when a FC process was used.

54

ACS Paragon Plus Environment

Page 55 of 56

Coefficient PC1

(a)

-0.6

-0.4

-0.2

0.0

0.2

0.4

0.6

4
0.6

soaking time

0.4

0.2

PC 2

char yield

0.0

pressure

-0.2

moisture

ash
-2

Coefficient PC2

lignin

-0.4

heating rate

-0.6

peak temp.

-4
-2

PC 1

(b)

Coefficient PC2
-0.6

-0.4

-0.2

0.0

0.2

0.4

0.6
0.6

lignin

heating rate

0.2

peak temp.
0

0.0

pressure

-0.2

-2

ash

soaking time

char yield

Coefficient PC3

0.4

PC 3

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Environmental Science & Technology

-0.4

moisture

-0.6

-4
-4

-2

PC 2

Figure 1. Score and loading plots for the three first principal components: (a), second versus first
principal component; (b); third versus second principal component.
55

ACS Paragon Plus Environment

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

Percentage of the theoretical value of yFC

Environmental Science & Technology

Page 56 of 56

100

80

60

40

20

LW

OW

MS

CC

Figure 2. Percentage of the theoretical value of yFC for four biomass feedstocks: leucaena wood
(LW), oak wood (OW), macadamia nut shells (MS), and corncob (CC). The white bars correspond
to the results obtained using a slow pyrolysis process, whereas the gray bars indicate the results
obtained when a FC process was used.

56

ACS Paragon Plus Environment

You might also like