Soil Biology & Biochemistry

Download as pdf or txt
Download as pdf or txt
You are on page 1of 25

Soil Biology & Biochemistry 43 (2011) 1812e1836

Contents lists available at ScienceDirect

Soil Biology & Biochemistry


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

Review

Biochar effects on soil biota e A review


Johannes Lehmann a, *, Matthias C. Rillig b, Janice Thies a, Caroline A. Masiello c,
William C. Hockaday d, David Crowley e
a
Department of Crop and Soil Science, Cornell University, Ithaca, NY 14853, USA
b
Institute for Biology, Free University of Berlin, 14195 Berlin, Germany
c
Department of Earth Science, Rice University, Houston, TX 77005, USA
d
Department of Geology, Baylor University, Waco, TX 76798, USA
e
Department of Environmental Sciences, University of California, Riverside, CA 92521, USA

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

Article history: Soil amendment with biochar is evaluated globally as a means to improve soil fertility and to mitigate
Received 14 December 2010 climate change. However, the effects of biochar on soil biota have received much less attention than its
Received in revised form effects on soil chemical properties. A review of the literature reveals a significant number of early studies
6 April 2011
on biochar-type materials as soil amendments either for managing pathogens, as inoculant carriers or for
Accepted 28 April 2011
Available online 18 May 2011
manipulative experiments to sorb signaling compounds or toxins. However, no studies exist in the soil
biology literature that recognize the observed large variations of biochar physico-chemical properties. This
shortcoming has hampered insight into mechanisms by which biochar influences soil microorganisms,
Keywords:
Activated carbon
fauna and plant roots. Additional factors limiting meaningful interpretation of many datasets are the
Biochar clearly demonstrated sorption properties that interfere with standard extraction procedures for soil
Black carbon microbial biomass or enzyme assays, and the confounding effects of varying amounts of minerals. In most
Charcoal studies, microbial biomass has been found to increase as a result of biochar additions, with significant
Fauna changes in microbial community composition and enzyme activities that may explain biogeochemical
Microorganisms effects of biochar on element cycles, plant pathogens, and crop growth. Yet, very little is known about the
Roots mechanisms through which biochar affects microbial abundance and community composition. The effects
of biochar on soil fauna are even less understood than its effects on microorganisms, apart from several
notable studies on earthworms. It is clear, however, that sorption phenomena, pH and physical properties
of biochars such as pore structure, surface area and mineral matter play important roles in determining
how different biochars affect soil biota. Observations on microbial dynamics lead to the conclusion of
a possible improved resource use due to co-location of various resources in and around biochars. Sorption
and thereby inactivation of growth-inhibiting substances likely plays a role for increased abundance of soil
biota. No evidence exists so far for direct negative effects of biochars on plant roots. Occasionally observed
decreases in abundance of mycorrhizal fungi are likely caused by concomitant increases in nutrient
availability, reducing the need for symbionts. In the short term, the release of a variety of organic molecules
from fresh biochar may in some cases be responsible for increases or decreases in abundance and activity
of soil biota. A road map for future biochar research must include a systematic appreciation of different
biochar-types and basic manipulative experiments that unambiguously identify the interactions between
biochar and soil biota.
Ó 2011 Elsevier Ltd. All rights reserved.

1. Introduction soil fertility as well as other ecosystem services and sequester


carbon (C) to mitigate climate change (Lehmann et al., 2006;
Biochar is the product of thermal degradation of organic mate- Lehmann, 2007a; Laird, 2008; Sohi et al., 2010). The observed
rials in the absence of air (pyrolysis), and is distinguished from effects on soil fertility have been explained mainly by a pH increase
charcoal by its use as a soil amendment (Lehmann and Joseph, in acid soils (Van Zwieten et al., 2010a) or improved nutrient
2009). Biochar has been described as a possible means to improve retention through cation adsorption (Liang et al., 2006). However,
biochar has also been shown to change soil biological community
* Corresponding author. Tel.: þ1 607 254 1236. composition and abundance (Pietikäinen et al., 2000; Yin et al.,
E-mail address: [email protected] (J. Lehmann). 2000; Kim et al., 2007; O’Neill et al., 2009; Liang et al., 2010;

0038-0717/$ e see front matter Ó 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.soilbio.2011.04.022
J. Lehmann et al. / Soil Biology & Biochemistry 43 (2011) 1812e1836 1813

Grossman et al., 2010; Jin, 2010). Such changes may well have effects their organo-chemical (Czimczik et al., 2002; Nguyen et al., 2010)
on nutrient cycles (Steiner et al., 2008b) or soil structure (Rillig and and physical properties (Downie et al., 2009). The role of biochar in
Mummey, 2006) and, thereby, indirectly affect plant growth soil biological processes therefore represents a frontier in soil
(Warnock et al., 2007). Rhizosphere bacteria and fungi may also science research, with many unexplained phenomena awaiting
promote plant growth directly (Schwartz et al., 2006; Compant exploration. Recent advances in our understanding of biochar
et al., 2010). The possible connections between biochar properties warrant an evaluation of the relationship between its properties
and the soil biota, and their implications for soil processes have not and its impact on the soil biota.
yet been systematically described. In this paper, we critically examine the state of knowledge on soil
The effectiveness of using biochar as an approach to mitigate populations of archaeans, bacteria, fungi, and fauna as well as plant
climate change rests on its relative recalcitrance against microbial root behavior as a result of biochar additions to soil. We develop
decay and thus on its slower return of terrestrial organic C as carbon concepts for a process-level understanding of the connection
dioxide (CO2) to the atmosphere (Lehmann, 2007b). Both the between biochar properties and biological responses, discuss the
composition of the decomposer community as well as metabolic ramifications of such changes for biogeochemical processes in soil,
processes of a variety of soil organismal groups may be important in and develop a road map for future research.
determining to what extent biochar is stable in soils, as is known for
wood decay (Fukami et al., 2010). Changes in microbial community 2. Modification of the soil habitat by biochar
composition or activity induced by biochar may not only affect
nutrient cycles and plant growth, but also the cycling of soil organic The material properties of biochar are very different from those
matter (Wardle et al., 2008; Kuzyakov et al., 2009; Liang et al., 2010). of uncharred organic matter in soil (Schmidt and Noack, 2000), and
In addition, biochar may change emissions of other greenhouse are known to change over time due to weathering processes,
gases from soil such as nitrous oxide (N2O) or methane (CH4) interactions with soil mineral and organic matter and oxidation by
(Rondon et al., 2005; Yanai et al., 2007; Spokas and Reicosky, 2009; microorganisms in soil (Lehmann et al., 2005; Cheng et al., 2008;
Clough et al., 2010; Singh et al., 2010; Zhang et al., 2010; Cheng and Lehmann, 2009; Nguyen et al., 2010). However, the
Taghizadeh-Toosi et al., 2011). Such changes may either reduce or relationships between biochar chemical and physical properties
accelerate climate forcing. The driving processes are still poorly and their effects on soil biota and potential concomitant effects on
identified (Van Zwieten et al., 2009). A more rapid mineralization of soil processes are poorly understood. This section gives a brief
indigenous soil C or greater emission of other greenhouse gases as overview of the unique properties of biochars compared to other
a result of biochar additions may counteract the benefits of reduced compounds in soil as a background to the following sections that
emissions elsewhere in the life cycle of a biochar system. A discuss the effects of biochar on soil biota.
systematic examination of the ways in which different microbial
and faunal populations may play a role in these biogeochemical 2.1. Basic properties: organic and inorganic composition
processes is still lacking.
Biochar may pose a direct risk for soil fauna and flora, but could Biochar composition can be crudely divided into relatively
also enhance soil health. Biochar addition may affect the soil bio- recalcitrant C, labile or leachable C and ash. The greatest chemical
logical community composition as demonstrated for the biochar- difference between biochar and other organic matter is the much
rich Terra preta soils in the Amazon (Yin et al., 2000; Kim et al., larger proportion of aromatic C and, specifically, the occurrence of
2007; O’Neill et al., 2009; Grossman et al., 2010), and has been fused aromatic C structures (Table 1), in contrast to other aromatic
shown to increase soil microbial biomass (Liang et al., 2010; O’Neill structures of soil organic matter such as lignin (Schmidt and Noack,
et al., 2009; Jin, 2010). Whether the abundance of microorganisms 2000). This fused aromatic structure of biochars in itself can have
increases or not, as discussed for mycorrhizal fungi (Warnock et al., varying forms, including amorphous C, which is dominant at lower
2007), is likely connected to the intrinsic properties of both biochar pyrolysis temperatures, and turbostratic C, which forms at higher
and the soil. Biochar properties vary widely and profoundly; not temperatures (Keiluweit et al., 2010; Nguyen et al., 2010). It is clear
only in their nutrient contents and pH (Lehmann, 2007a), but also in that the nature of these C structures is the chief reason for the high

Table 1
Physical and chemical properties of contrasting biochars relevant to biological processes in soil (Nguyen and Lehmann, 2009; Nguyen et al., 2010; Enders, Hanley and Lehmann,
unpubl. data; Hockaday, unpubl. data).

Feedstock Temperature pH pH CECa CECa C (%) C/N Total P Ashb Volatilesb Fixed H/C O/C Aromatic Aromatic SSAe
( C) (KCl) (H2O) (mmolc (molc m2) ratio (mg kg1) (%) (%) Cb (%) ratioc ratioc Cd (% of clusters (m2 g1)
kg1) total)
Oak wood 60 3.16 3.73 182.1 NDf 47.1 444 5 0.3 88.6 11.1 1.48 0.72 ND ND ND
350 5.18 4.80 294.2 0.65 74.9 455 12 1.1 60.8 38.1 0.55 0.20 82.8 18 450
600 7.90 6.38 75.7 0.12 87.5 489 29 1.3 27.5 71.2 0.33 0.07 86.6 37 642

Corn stover 60 6.33 6.70 269.4 ND 42.6 83 526 8.8 85.2 6.0 1.56 0.74 2.0 6 ND
350 9.39 9.39 419.3 1.43 60.4 51 1889 11.4 48.8 39.8 0.75 0.29 76.9 19 293
600 9.42 9.42 252.1 0.48 70.6 66 2114 16.7 23.5 59.8 0.39 0.10 88.2 40 527

Poultry litter 60 7.53 7.53 363.0 ND 24.6 13 16,685 36.4 60.5 3.1 1.51 1.03 ND ND ND
350 9.65 9.65 121.3 2.58 29.3 15 21,256 51.2 47.2 1.6 0.57 0.41 ND ND 47
600 10.33 10.33 58.7 0.63 23.6 25 23,596 55.8 44.1 0.1 0.18 0.62 ND ND 94
a
Cation exchange capacity, determined at pH 7 using buffered ammonium acetate (Nguyen and Lehmann, 2009).
b
Mass % w/w analyzed using ASTM D1762-84.
c
Molar ratios.
d
In rings, determined by direct polarization 13C nuclear magnetic resonance spectroscopy.
e
Specific surface area, CO2 as sorbent (courtesy A. Zimmerman).
f
Not determined.
1814 J. Lehmann et al. / Soil Biology & Biochemistry 43 (2011) 1812e1836

stability of biochars (Nguyen et al., 2010) (Fig. 1). It is less clear what pyrolysis temperatures above 600  C (Freitas et al., 2002). Grasses
precise mechanisms directly or indirectly confer stability to the and a number of common feedstocks (rice hulls, nut shells, sewage
aromatic C structures in soil. sludge, etc.) also contain substantial quantities of amorphous silica
The chemical stability of a large fraction of a given biochar (>2 wt %). Analyses by 29Si (silicon) NMR and X-ray diffraction
material means that microorganisms will not be able to readily shows the formation of silicon carbide (SieC) at pyrolysis temper-
utilize the C as an energy source or the N and possibly other atures above 1200  C (e.g., Lee and Cutler, 1975; Freitas et al., 2000),
nutrients contained in the C structure. However, depending on the temperatures that are commonly reached during biomass gasifi-
type of biochar, a fraction may be readily leached and therefore cation. The SieC bonds likely take part in cross-links between
mineralizable (Lehmann et al., 2009) and in some cases has been aromatic domains or crystallites (Freitas et al., 2000). At tempera-
shown to stimulate microbial activity and increase abundance tures of 400e600  C, pyrolysis alters the chemical structure of bio-
(Steiner et al., 2008a). At present, such fractions may be quantified silicates, with a progressive increase in SiO4 relative to SiO2e3 with
by incubation studies and are frequently referred to as “volatile increasing heat treatment temperature (Freitas et al., 2000). Sili-
matter” or the labile fraction (Table 1). Volatile matter refers to an cates can occupy a substantial portion (>14% for corn cobs and 88%
ASTM standard methodology that was developed to evaluate the for rice hulls) of the biochar pore space (Bourke et al., 2007; Freitas
quality of coals as fuels, and is only beginning to be evaluated as et al., 2000). Nevertheless, the influence of silicates and the effects
a material property with explanatory value for biochar stability of changes in silica crystal structure on biochar structure and
(Deenik et al., 2010; Zimmerman, 2010). However, such quantified function have not been investigated. The bioavailability of Fe and Si
volatile matter (5e37% of C in the study by Zimmerman, 2010) is in biochar is unknown, but treatments with aqueous acid solutions
typically much larger than the corresponding mineralization used in common soil tests are effective in extracting a portion of Si,
(2e18% of C over one year). This may indicate that the mineralizable Fe, S, P, K, Mg, and Ca from biochar (Freitas et al., 2002; Bourke et al.,
fraction is imperfectly captured by volatile matter despite often 2007; Major et al., 2010), suggesting that some fraction of these
acceptable correlation results. Further improvement to capture the nutrients may be accessible to plants and microorganisms.
fraction that is potentially bioavailable may be required.
The third major component is comprised of minerals that are
present as ash inclusions in biochar. These minerals include several 2.2. Biochar surface properties and sorption
essential macro- and micro-nutrients for biological uptake and,
therefore, represent valuable resources in the soil food web. Addi- Fresh biochars can have net positive or net negative surface
tionally, the presence of these elements during pyrolysis plays a role charge, but typically have initially low cation exchange capacities
in the biochar chemical structure to the extent that they are (CEC) compared to soil organic matter on a mass basis (Table 1;
incorporated into the aromatic structure or that organo-metal Lehmann, 2007a; Chan and Xu, 2009). Notable is the initially
reactions are thermodynamically favorable at high temperatures. measurable anion exchange capacity which disappears over time in
For instance, N may substitute one or two C atoms in aromatic soil (Cheng et al., 2008), and in some cases strong interaction with
compounds (Leinweber et al., 2007) with largely unknown effects phosphates (Beaton et al., 1960). High-ash biomass generates bio-
on biochar behavior in soil. Iron (Fe)-rich biochars made from peat chars with slightly greater CEC and charge density upon normali-
and investigated by 57Fe Mössbauer spectroscopy show the zation of CEC to surface area (Table 1). On the other hand, greater
formation of Fe3C bonds and small ferromagnetic iron clusters at pyrolysis temperatures cause a decrease in CEC, especially in charge
density as a result of the greater surface area produced at high
temperatures of up to 600  C and loss of volatile matter (Table 1),
which may contain a substantial portion of the negative charge and
CEC as organic acids. Sorption of comparatively polar organic matter
such as catechol or humic acid extracts to wood biochars increased
Recalcitrance in the range from 400 to 650  C, likely due to greater nanopore
Fraction of surface area (Kasozi et al., 2010). At even higher temperatures of up
labile OM to 1000  C (typically to produce so-called “activated carbon”, but
Energy
availability also gasification biochars), carbons are mainly hydrophobic and do
not sorb appreciable amounts of nutrients or polar organic
substances, such as sugars (Yam et al., 1990). The same has also been
Fungi Amount and observed for naturally occurring black C (Cornelissen et al., 2005).
composition
Bacteria of minerals
Rather, such high-temperature carbons sorb mainly non-polar or
weakly polar organic solutes, notably those bearing aromatic
Surface
area
Fauna (nutrients,
structures, as have been described in a large number of studies in
pH salts, heavy
Roots the scientific literature (Moreno-Castilla, 2004). These also include
metals)
enzymes (Cho and Bailey, 1978) and other substances important for
Sorption
microbial processes in soil. This body of information is not fully
transferable to biochars that are produced at much lower temper-
Pore structure atures, and without the addition of chemicals that are commonly
and size Surface chemical used to increase the surface area of activated carbons. Nonetheless,
distribution properties (charge, prior research on activated carbon may be informative for
hydrophobicity) describing the properties of biochars produced at comparatively
high temperatures such as during some gasification procedures and
may even apply for understanding some surface properties of low-
temperature biochars. In soil, biochars (those produced at or below
Fig. 1. Schematic overview of the connection between primary biochar properties
(outer circle), the soil processes they may influence (intermediate circle) and the soil
600e700  C) seem to oxidize rapidly and attain greater amounts of
biota (inner circle) (shorter distance give a qualitative estimate of the strength of the CEC (Cheng et al., 2008; Nguyen et al., 2010), but initially still retain
connection). White arrows indicate influence between biochar properties. a significant proportion of non-polar surfaces (Smernik, 2009).
J. Lehmann et al. / Soil Biology & Biochemistry 43 (2011) 1812e1836 1815

Given that different feedstock properties range from mineral- 2.4. The biochar-aggregate analogy
poor woody materials to mineral-rich manures or crop residues,
such as rice hulls, the resulting pH is highly variable from below pH In contrast to other organic matter in soil, biochars remain
4 to above pH 12, even for the same biomass type (Lehmann, particulate over long periods of time (Skjemstad et al., 1996;
2007a). Typically, biochars with high mineral ash content have Lehmann et al., 2005, 2008b), even though particle sizes may
greater pH values than those with lower ash contents (Table 1). For decrease on a decadal time scale (Nguyen et al., 2008). Although
all feedstocks, pH increases with greater pyrolysis temperature. biochar apparently has a monolithic structure on the millimeter
Over time, the pH of biochars may change and either decrease or scale, it can be viewed, on the micro- and nanometer scale, as
increase depending on type of feedstock. Nguyen and Lehmann a disordered mixture of C clusters and mineral elements (i.e., ash
(2009) observed a pH decrease with mineral-poor oak wood bio- inclusions). In addition, biochar particles have large internal surface
char from pH 4.9 to 4.7, but an increase with mineral-rich corn areas and pores that may be important for biological processes. In
stover biochar from pH 6.7 to 8.1 over the course of a one year this sense, the biochar particle can be compared to a soil aggregate.
incubation. The driving force behind a pH decrease is oxidation of C Biochar “aggregates” may provide similar functions such as protec-
to form acidic carboxyl groups (Cheng et al., 2006), whereas the tion of organic matter, habitat for soil biota, or retention of soil
increase in pH is likely related to the dissolution of alkaline moisture and nutrients as described for aggregates made from
minerals. minerals and organic matter (Tisdall and Oades, 1982).
Biochar properties such as total surface area and pore size
2.3. Biochar physical properties distribution are known to vary with feedstock properties and
pyrolysis temperatures (Downie et al., 2009; Table 1). In addition,
The effects of biochar on soil biota may be driven as much by its surface area and pore volume may change upon contact with soil by
physical properties as by its chemical properties. The differences in pore clogging from sorbed organic (Pignatello et al., 2006) and
physical structure between biochar and soils lead to altered soil mineral material (Joseph et al., 2010) or, conversely, possibly by
tensile strength, hydrodynamics, and gas transport in a soilebiochar mineralization of volatile matter that may be blocking pores. These
mixture; all of which can be expected to have major impacts on properties have shown to change sorption behavior of mineral
soil biota. The extent of these effects will depend on the biochar (Liang et al., 2006) and organic matter (Kasozi et al., 2010) which in
production conditions and feedstock, which together control the turn may influence energy and pore space available to soil biota
macro-and micro-structure of biochar particles (Downie et al., 2009). (Fig. 1).
Whether these effects are merely a result of a mixture of two very Many soil microorganisms are specialists living in microhabitats
different materials (soil and biochar) or whether biochar has that provide resources for their specific metabolic needs. For
a distinct effect on soil properties on a fine spatial scale has not been instance, aerobic microbes live at the surface of soil aggregates,
focus of experimentation. while denitrifiers and semi-aquatic species dwell within the moist
When the tensile strength of biochar is less than that of soils interior of soil peds (Sexstone et al., 1985). Organic matter decom-
(e.g., for clay-rich soils), biochar addition can reduce the overall position rates are higher at the surface of soil aggregates than in the
tensile strength of the soil. In a pot trial with a hard-setting Chro- core of aggregates due to higher influx of resources at the surface
misol (an Alfisol in USDA nomenclature), Chan et al. (2007) found (organic matter, moisture, and O2). This is evident from depleted C
a decrease in soil tensile strength from an initial, biochar-free value concentrations and C-to-N ratios, as well as the oxidation of lignin
of 64.4 kPae31 kPa at an amendment rate of 50 t biochar ha1; the phenols and the accumulation of microbial polysaccharides at the
tensile strength was again reduced to 18 kPa at 100 t biochar ha1 aggregate surface relative to the aggregate core (Amelung and Zech,
(Chan et al., 2007). Mechanical impedance is one of the main 1996). Similarly, the exterior surfaces of biochar particles in the soil
factors determining root elongation and proliferation in soil are significantly more oxidized than the particle interior or core
(Bengough and Mullins, 1990). Reductions in soil tensile strength (Lehmann et al., 2005; Liang et al., 2006; Cheng et al., 2008). This is
may therefore make root and mycorrhizal nutrient mining more due to sorption of organic matter on the biochar surface and the
effective, as well as allow seeds to germinate more easily. Reduced oxidation of the biochar C itself (Liang et al., 2006), both biotically
soil tensile strength may also make it physically easier for inver- and abiotically mediated via reactions with O2 (Cheng et al., 2006,
tebrates to move through the soil, altering predator/prey dynamics. 2008). Similar to soil aggregates, the preferential oxidation of the
Because a reduction in tensile strength could facilitate both biochar particle surface relative to the particle interior implies
increased root growth and increased root predation, it is not clear a limited diffusion of O2 to the interior of biochar particles. Such
what the net effect of a reduction in tensile strength would be on differential redox conditions not only influence organic matter
root systems. oxidation but also metal transformation.
Biochar application can also change soil bulk density (e.g., Major
et al., 2010); with possible effects on soil water relations, rooting 3. Responses of the soil biota to biochar
patterns and soil fauna. This occurs both because the density of
biochar is lower than that of some minerals, and because biochar The application of biochar as a targeted strategy for managing
contains macro- and micropores (Downie et al., 2009), which can soil biota is a topic of growing interest, and inadvertent changes of
hold air or water, greatly reducing the bulk density of the entire soil biota as a result of biochar application are of equally strong
biochar particle. Surprisingly little bulk density data have been concern. This line of research is an important one, as the health and
published for biochar or natural char samples. Density measure- diversity of soil microbial populations are critical to soil function
ments for biochar should distinguish between the true, solid and ecosystem services, which, in turn have implications for soil
particle density and the bulk density of the biochar particles plus structure and stability, nutrient cycling, aeration, water use effi-
their pore space. Published true biochar densities are high, between ciency, disease resistance, and C storage capacity (e.g., Brussaard,
1.5 and 2.1 g cm3 for a range of feedstocks (Brewer et al., 2009), 1997). Brussaard et al. (2007) suggest that organic amendments
whereas, bulk densities typically lie between 0.09 and 0.5 g cm3 are perhaps the most important means of managing biodiversity in
(Karaosmanoglu et al., 2000; Özçimen and Karaosmanolu, 2004; soils. It is well-known that the quantity, quality, and distribution of
Bird et al., 2008; Spokas et al., 2009), values much lower than organic amendments each affect the trophic structure of the soil
those of soils. food web (Moore et al., 2004). Therefore, all three of these aspects
1816 J. Lehmann et al. / Soil Biology & Biochemistry 43 (2011) 1812e1836

should be considered in the use of biochar as a soil management (Makoto et al., 2010). Likewise, AM colonization of wheat roots was
tool. Early research has focused on the effects of biochar properties found to increase to 20e40% two years after Eucalyptus wood
at the level of primary decomposers (bacteria and fungi). Other biochar additions of 0.6e6 t ha1, in comparison to a colonization
functional groups, including secondary decomposers, predators, rate of 5e20% in unamended controls (Solaiman et al., 2010). It is
and soil animals, also play important roles in nutrient and energy far less clear how the soil-borne phase of the fungus, the extra-
cycling. In the following sections we consider how biochar affects radical mycelium, is affected by biochar. Direct interactions with
soil biota on several trophic levels, including root dynamics, and biochar particles could be important. For example, the internal pore
discuss the reasons behind observed changes with respect to systems of biochar particles may protect the extraradical mycelium
different biochar properties. from grazers (Warnock et al., 2007; Table 2 for a summary of
mechanisms; Fig. 2). Sorption of signaling compounds, detoxifica-
3.1. Abundance of microorganisms tion of allelochemicals, soil physico-chemical properties or indirect
effects through alterations of other soil microbial populations have
Microbial abundance has been determined in biochar-amended also been discussed (Warnock et al., 2007; Elmer and Pignatello,
soil by various methods including, total genomic DNA extracted 2011). Recently, Rillig et al. (2010) found that biochar produced
(O’Neill, 2007; Grossman et al., 2010; Jin, 2010), culturing and plate via hydrothermal carbonization (here called “hydrochar” due to its
counting (Jackson, 1958; O’Neill et al., 2009), substrate-induced properties that differ greatly from pyrolysis chars) could stimulate
respiration (Zackrisson et al., 1996; Steiner et al., 2004, 2009; Wardle spore germination of AM fungi, which is another mechanism
et al., 2008; Kolb et al., 2009), fumigationeextraction (Jin, 2010; potentially leading to increased populations of these symbionts in
Liang et al., 2010), phospholipid fatty acid (PLFA) extraction (Birk soil.
et al., 2009), staining and direct observation of individual biochar Decreases in AM abundance or relative proportion have also
particles (Jackson, 1958; Pietikäinen et al., 2000; Warnock et al., been observed after additions of biochar (Gaur and Adholeya, 2000;
2007; Jin, 2010; Fig. 2). The microbial reproduction rate has also Birk et al., 2009; Warnock et al., 2010). The reasons for such
been shown to increase in some biochar-amended soils (Pietikäinen decreases are not entirely clear but could stem from: (i) a reduced
et al., 2000; Steiner et al., 2004), and in waste water (Koch et al., requirement for mycorrhizal symbiosis due to increased nutrient
1991). Similarly, in biodigesters used to generate methane (CH4) as and water availability to plants; decreases in mycorrhizal abun-
an energy source, additions of biochar (commercial wood charcoal) dance have been observed, for example, with greater P availability
led to an increase in anaerobic and cellulose-hydrolyzing bacterial in soil (Corbin et al., 2003; Covacevich et al., 2006; Gryndler et al.,
abundance (Kumar et al., 1987). 2006); (ii) changes in soil conditions, e.g., due to modifications of
The reasons for changes in microbial abundance may differ for pH or water relations (discussed below); or (iii) direct negative
different groups of microorganisms. The two most commonly effects from high contents of mineral elements or organic
occurring types of mycorrhizal fungi (arbuscular [AM] and ecto- compounds detrimental to the fungi, such as high salt or heavy
mycorrhizal [EM]), are often positively affected by biochar pres- metal contents (Killham and Firestone, 1984; Killham, 1985). (iv)
ence, as reviewed in Warnock et al. (2007). Mycorrhizal response in Sorption of organic C and organically-bound nutrients may influ-
the host plant is most commonly assessed by measuring root ence their availability (Pietikäinen et al., 2000; Chan and Xu, 2009).
colonization; that is, the abundance of fungal tissue in the host. The first two scenarios are specific to a certain soil environment,
Both formation rate and tip number of EM infection of larch whereas the third is primarily a result of biochar properties which
seedling roots was increased by 19e157% with biochar additions can significantly vary.

Fig. 2. Visual observation of spatial association and colonization of biochar by microorganisms. (a) fresh biochar showing fungal hyphae (Lehmann and Joseph, 2009; with
permission); (b) fresh corn biochar showing microorganisms in pores (arrows) (Jin, 2010; with permission). (c) 100-year-old char from a forest fire isolated from a frigid entic
Haplorthod (Hockaday et al., 2007; with permission); (d) 350-year-old char from a forest fire in a Boreal forest soil (Zackrisson et al., 1996; with permission).
J. Lehmann et al. / Soil Biology & Biochemistry 43 (2011) 1812e1836 1817

Table 2
Summary of possible mechanisms by which microbial abundance is affected by biochar additions to soil; (þ) indicates that relative abundance may increase (not necessarily
better growth conditions), () indicates that relative abundance decreases.

Mechanism Rhizobia Other bacteria Mycorrhizal fungi Other fungi


Protection from grazers nc (þ) (þ) (þ)
Improved hydration þ þ ? ? or 
Greater P, Ca, Mg, K availability þ þ e e
Greater micronutrient availability þ þ e ?
Higher pH þ þ nc nc
Lower pH e e nc nc or e
Sorption of signaling compounds ? or e ? ? ?
Greater N availability (also through sorption of phenolics and increased nitrification) e þ or e nc nc
Sorption of microorganisms nc þ nc nc
Biofilm formation þ þ ? ?
Sorption of inhibitory compounds ? þ ? ?
Sorption of dissolved OM as an energy source for microorganisms ? ? nc ?

nc, no change.
?, reaction not known.
Parentheses, weak circumstantial evidence.

3.1.1. Influence of nutrient and carbon availability production and to decrease the time between inoculation and first
on microbial abundance appearance of nodules. Gibson and Nutman (1960) attributed
Nutrient additions by fertilizers reduced the enhancing effect of similar findings to nitrate adsorption, which is unlikely to be the
biochars on microbial reproduction rates (Steiner et al., 2009). case given the low amount of anion exchange capacity typically
Similarly, Blackwell et al. (2010) found significant increases in the found in biochars (Cheng et al., 2008).
proportion of root colonization of wheat with AM in biochar- The available research appears to indicate that nutrient and C
amended soils at no or low fertilizer additions, but no significant availability changes may both increase or decrease microbial
increases when large amounts of nutrients were applied. This effect biomass, depending on (i) the existing nutrient and C availability in
depends on the type of fertilizer applied and the particular soil; (ii) the magnitude of change; and (iii) the microorganism
microorganism group. Mycorrhizal infection was reduced by P- group. This response may prove to be predictable with further
containing fertilizers despite the presence of biochar, whereas this experimental evidence.
was not observed with fertilizers that only contained N. The reverse
was observed for nodule formation by rhizobia (Ogawa and 3.1.2. Influence of pH on microbial abundance
Okimori, 2010). This can be explained by the different need of the Microbial biomass increases with rising pH values have been
plant to form symbiotic relationships with microorganisms under shown for a gradient from pH 3.7 to 8.3 under otherwise identical
changing nutrient limitations. With N fertilizer additions, the plant environmental conditions by Aciego Pietry and Brookes (2008).
may not need to rely on biological N2 fixation as much as under N However, fungal and bacterial populations react differently to
limitation. Similar explanations may hold for the effect of what is changes in pH. Bacteria are likely to increase in abundance with
likely increased C supply by exudation or root turnover in the rising pH up to values around 7, whereas, fungi may show no
rhizosphere and C as energy sources for heterotrophic microor- change in total biomass (Rousk et al., 2010), or potentially
ganisms. Consequently, Jin (2010) found greater enhancement of dramatically reduce their growth at higher pH (Rousk et al., 2009).
microbial abundance by biochar additions in the rhizosphere than This may also apply to rhizobial infection of legumes (Angelini et al.,
in bulk soil, whereas Graber et al. (2010) reported the opposite. 2003).
On the other hand, microbial abundance, especially that of non- After biochar additions, the pH of soils may increase or decrease,
symbiotic microorganisms under nutrient limiting conditions, may depending on the pH and liming value of the biochar. Biochars can
be increased by slightly greater nutrient availability (Lochhead and have pH values of below 4 or above 12, depending on feedstock
Chase, 1943; Taylor, 1951), either due to biochar-driven improve- type, pyrolysis temperature (Lehmann, 2007a; Chan and Xu, 2009)
ments in nutrient retention or due to nutrients that are released by and degree of oxidation (Cheng et al., 2006), which generates very
the biochar. However, the abundance of symbionts such as different living conditions for microorganisms in biochar pore
mycorrhizal fungi has been shown to improve by alleviating N and spaces. Given the often observed spatial proximity of microorgan-
P limitations of the fungi themselves (Treseder and Allen, 2002). In isms and biochar surfaces (Fig. 2), the pH of biochars may therefore
most cases, though, reductions in abundance have been found to have a very important influence on total microbial abundance.
occur for rhizobia and mycorrhizal fungi as a result of significantly Rillig et al. (2010), working with hydrochars (which had an
greater N and P availability, respectively. Little direct evidence is acidic pH of 4.10) observed a microbial activity-dependent increase
available for nutrient-related effects of biochar on microorganisms in soil pH, due to some unidentified microbially-mediated reduc-
(Warnock et al., 2010). tions of substrates or electron acceptors. This illustrates that
Increased micronutrient concentrations, namely of molyb- secondary processes may also be responsible for soil pH changes
denum (Mo) and boron (B), were thought to be responsible for following addition of biochar. Microbial biomass assessed by
enhanced biological N2 fixation (BNF) by rhizobia in legumes grown substrate-induced respiration (SIR) correlated positively with pH in
in biochar (Rondon et al., 2007), but the same may not hold for an acid Xanthic Ferralsol of Brazil after biochar applications (Steiner
abundance of rhizobia. For example, Vantsis and Bond (1950) did et al., 2004). This likely indicates a stimulation of microbial repro-
not find that the Mo in biochar improved nodule dry weight and duction when the pH of these acid soils was increased by adding
activity; instead, they favored an explanation whereby biochar biochar. Similar to nutrient and C changes, the effects of pH changes
enhanced BNF by sorbing toxic substances, which was confirmed induced by biochar will largely depend on the pre-existing soil pH,
by Nutman (1952). Similarly, Turner (1955) identified sorption of the direction and magnitude of change. This may be predicted from
inhibitory compounds to wood biochar to increase nodule available literature but must recognize that pH measurements of
1818 J. Lehmann et al. / Soil Biology & Biochemistry 43 (2011) 1812e1836

Fig. 3. Processes for possible attachment of viable microbial cells to surfaces (Cassidy et al., 1996; with permission).

bulk soils does not reflect pH values experienced by microorgan- volatile content that are highly variable (Table 1). Formation of
isms located around biochar particles. surfactants by microorganisms (Ron and Rosenberg, 2001) may
additionally facilitate adhesion to biochars.
3.1.3. Influence of bacterial adhesion to biochar The mere increase of colonizable surfaces through biochar may
on microbial abundance increase the microbial biomass as shown for sediments (Yamamoto
Bacteria may sorb to biochar surfaces, rendering them less and Lopez, 1985). The specific surface area of coarse-textured soil
susceptible to leaching in soil (Pietikäinen et al., 2000). This would may be increased by additions of those biochars that have large
increase bacterial abundance but likely has no effect on fungal surface areas (Table 1). In how far the shapes of biochar surfaces
abundance, as fungi will be less mobile owing to their hyphal plays a role (Yamamoto and Lopez, 1985) is not clear.
network. Microbial sorption (called “immobilization” in the relevant
literature) has been studied and used in industrial and scientific 3.1.4. Biochar protection of microorganisms from other biota
applications, but much less is known about environmental applica- Both bacteria and fungi are hypothesized to be better protected
tions (Cassidy et al., 1996). The main processes leading to attachment against grazers or competitors by exploring pore habitats in bio-
are (1) flocculation, (2) adsorption on surfaces, (3) covalent bonding chars (Ogawa, 1994; Ezawa et al., 2002; Saito and Marumoto, 2002;
to carriers, (4) cross-linking of cells, (5) encapsulation in a polymer- Thies and Rillig, 2009). No quantitative evidence is currently
gel, and (6) entrapment in a matrix (Fig. 3). available for microbial protection in biochar pores, but pore size
Adsorption to biochar may occur via different processes distribution of microorganisms and biochars, as well as visual
including hydrophobic attraction or electrostatic forces. At the investigations (Fig. 2), provide justification for this hypothesis
iso-electric point, adsorption of Escherichia coli to demineralized (Thies and Rillig, 2009). As pointed out above, some quantitative
activated carbon was negligible and increased with increasing
hydrophobicity (Rivera-Utrilla et al., 2001). In the presence of
minerals, the adsorption further increased. The iso-electric point of
biochars, however, has been found to be low (pH < 4; Cheng et al.,
2008). Adsorption may be facilitated by precipitates forming
on carbon surfaces under an electric current (Fig. 4; George and
Davies, 1988). The effect of electric currents on microbial adhe-
sion and activity is only beginning to be explored (Bond, 2010).
Since some biochars may contain large proportions of minerals,
such as those produced from crop residues or animal manures, such
processes might be relevant.
Adhesion may also depend on pores sizes (Rivera-Utrilla et al.,
2001). Pore sizes for optimum adhesion may need to be 2e5
times larger than cell size if microorganisms are to enter the pores,
or about 2e4 mm for Bacillus mucilaginosus and Acinetobacter sp.
(Samonin and Elikova, 2004). Adhesion may be diminished in
larger and smaller pores either because pore curvature is too large
to enhance adhesion or microorganisms do not fit into the pores,
respectively (Samonin and Elikova, 2004). It is therefore likely that Fig. 4. Adhesion of Escherichia coli (white arrows) on activated carbon with the
the ability of biochars to retain bacteria will vary greatly depending occurrence of precipitates (black arrows) from added Mg2þ under application of
on the biochar properties including the ash content, pore size, and a negative potential (George and Davies, 1988; with permission).
J. Lehmann et al. / Soil Biology & Biochemistry 43 (2011) 1812e1836 1819

evidence exists for the importance of pore sizes for the retention of 3.1.7. Methodological challenges and recommendations
microorganisms (Cassidy et al., 1996; Rivera-Utrilla et al., 2001; Several important analytical challenges arise as a result of strong
Samonin and Elikova, 2004). Targeted experimentation is needed to sorption of lysed cells, cell contents, and microbial exoenzymes to
test whether the physical location within pores or also the biochar. This sorption occurs during bead-beating for DNA extrac-
adsorption to surfaces confers protection against predators. tion, during fumigationeextraction when measuring microbial
biomass, and when using fluorometric methods to detect enzyme
3.1.5. Influence of sorption of toxins and chemical activity. For fumigation methods, a correction factor needs to be
signals on microbial abundance applied to account for the stronger sorption of lysed cell constitu-
Sorption of compounds to biochar that would otherwise inhibit ents to the biochar than to soil alone. Liang et al. (2010) calculated
microbial growth may increase microbial abundance. Compounds an extraction factor based on recovery of 13C-labeled microbial
such as catechol that are toxic to microorganisms (Chen et al., 2009) biomass added to soil. The recovery of microbial biomass was found
were found to be strongly sorbed to comparatively high- to be 21e41% lower in biochar-rich Terra preta soils than in biochar-
temperature biochars produced from ash-rich corn stover (Kasozi poor adjacent soils. For temperate soils with additions of corn
et al., 2010). Application of fast-pyrolysis biochar from wood stover biochar, Jin (2010) used adsorption isotherms to correct for
powder increased AM colonization of asparagus in soils that the sorption of dissolved organic C to biochar. After correction for
received aromatic acids (e.g., cinnamic, cumaric, and ferulic acids) sorption, estimated microbial biomass increased by over 70% at an
known to have allelopathic effects. Graphite and activated carbon application rate of 30 t biochar ha1. The need to correct for sorp-
(which may have properties similar to some high-temperature tion significantly increased with greater biochar application rates
biochars with the caveats discussed earlier) were shown to increase (Jin, 2010).
colonization and germination of Bacillus strains under high salt Durenkamp et al. (2010) observed no effect on extraction effi-
contents on agar media (Matsuhashi et al., 1995). For gonococci and ciency when biochars were mixed briefly with soil prior to extract-
meningococci, growth changes varied strongly depending on the ing soluble C. This indicates that either short-term exposure to
type of carbons used and the temperature to which it was heated biochars may have limited effects on extractability, or, lysed cells
(Glass and Kennett, 1939). Compounds were desorbed from biochar- after fumigation are more susceptible to sorption than total soil
type substances used for preparation of an agar growth medium organic matter. In the same experiment, activated carbon produced
which proved toxic to the bacillus Bordetella pertussis (Pollock, at high temperature resulted in significantly decreased extraction
1947), indicating that growth-inhibiting substances were retained efficiency, suggesting that biochars with greater surface area exac-
by biochars. Cell counts were several fold greater already with 0.05% erbate the sorption.
(w/v) of biochar-type material mixed into the agar and did not Similarly, DNA extracted from biochar-amended soils is reduced
increase with greater additions (Ensminger et al., 1953). Microbial by sorption of DNA released from microbial cells to biochar and
cultures were in general shown to grow more profusely with the requires suitable selection of extraction kits to overcome these
addition of biochar-type materials (Mishulow et al., 1953); however, limitations. O’Neill (2007) found that bacterial abundance was
the origin and properties of the charcoals used were not sufficiently higher in four different Terra preta soils than in adjacent soils when
described to allow further conclusions to be drawn. The authors measured by plate counting, whereas, DNA extractions from the
speculated that sorption of growth-regulating compounds or same soils indicated the reverse. Indeed, when biochar is added to
bacterial cells may have played a role. soil, DNA yield typically decreases (Fig. 5; Jin, 2010). This has been
DeLuca et al. (2009) speculated whether also adsorption of described for various charcoals and activated carbons in the past
signaling compounds from legumes, such as flavonoids, to biochar (Rapaport et al., 1981; Gani et al., 1999). Appropriate selection of an
surfaces may render them ineffective for inducing nodule formation extraction kit with the greatest extraction efficiency as well as
(Jain and Nainawatee, 2002). Interference may also occur with purity of the extracted DNA (Fig. 5) is able to address this constraint
flavonoid signaling of AM fungi, wherein flavonoids were shown to be (O’Neill, 2007; Jin, 2010). Use of different DNA extraction methods
sorbed to activated carbon (Akiyama et al., 2005). Very little is known did not, however, affect the characterization of the soil microbial
about how biochar might interference with or possibly enhance the community as assessed by terminal restriction fragment length
activity of signaling compounds. Given the strong sorption of organic polymorphism (T-RFLP) profiling of both the bacterial and fungal
matter to biochar surfaces (Smernik, 2009), interferences with communities (Jin, 2010), which affirmed the robustness of the
signaling between roots and rhizosphere microorganisms is very results for these community analyses despite the challenges of
likely. Its quantitative importance has not been explored. extraction.
It is insufficiently quantified how biochar may interfere with
3.1.6. Influence of protection against desiccation other microbial measurements based on extraction of biomole-
on microbial abundance cules, such as in phospholipid fatty acid (PLFA) analyses, or activity
Periodic drying of soil leads to stress and, ultimately, to measures, such as determined with dyes during enzyme assays, etc.
dormancy or mortality of microorganisms, with important differ- Enzyme interactions with biochars have been studied by Bailey
ences between gram negative and gram positive bacteria, as well as et al. (2010) who recommended the use of fluorescence studies
between bacteria and fungi (Schimel et al., 2007). Given the large due to possible sorption making color reactions less reliable. They
surface area of biochars (Liang et al., 2006; Downie et al., 2009) and tested whether the enzyme or the substrate or both were likely to
greater water holding capacity after addition to light-textured soil become sorbed to biochar particles and found sorption varied for
(Kishimoto and Sugiura, 1985; Glaser et al., 2002), biochars may a range of both enzymes and substrates, making it difficult to draw
retain moist pore spaces that allow continued hydration of micro- general conclusions. It is highly likely that the nature of the biochar
organisms in a drying soil. Malik (1990) found that survival capacity tested will also alter these results. In contrast to Bailey et al. (2010),
and reactivation of several oxygen-sensitive bacteria during freeze- Jin (2010) showed that fluorogenic molecules such as 4-methyl-
drying was greatly increased in the presence of activated carbon, umbelliferone (MUF) and 7-amino-4-methylcoumarin (MCA) are
and attributed the effect partly to reduced surface tension, in sorbed strongly to biochar, especially within the first 30 min of
addition to pH buffering and sorption of radicals. Given the greatly incubation with a MUF- or MCA-labeled substrate. Such interfer-
varying pore structures of different biochars, biochar properties ences should be considered more rigorously in the future for all
will largely determine whether such processes will occur. other extraction or staining methods, as most will require
1820 J. Lehmann et al. / Soil Biology & Biochemistry 43 (2011) 1812e1836

2.5 Instead, the altered soil environment, either in terms of an altered


a resource base (e.g., available C, nutrients, water), shifts in abiotic
factors (e.g., pH, toxic elements), or different habitat as discussed
DNA purity (A260/A280)

2.0 above may cause some microbial groups to become competitively


dominant, leading to changes in community composition and
structure. Additionally, changes in trophic relationships (as a conse-
1.5 quence of changes in soil biota abundances higher up in the soil food
web; see Section 3.3) may cause top-down effects that constrain
certain microbial groups.
1.0 Consequently, studies on (i) Terra preta soils, (ii) soils rich in char
from vegetation fires, and (iii) soils amended with biochar have
shown significant changes in community composition and diversity
0.5
of both fungal, bacterial, and archaeal populations (Kim et al., 2007;
Otsuka et al., 2008; O’Neill et al., 2009; Graber et al., 2010; Grossman
3.0 et al., 2010; Jin, 2010; Taketani and Tsai, 2010; Khodadad et al., 2011).
Bacterial diversity was greater by as much as 25% in biochar-rich
b
Terra preta soils compared to unmodified soils in both culture-
2.5 independent (Kim et al., 2007) and culture-dependent (O’Neill
et al., 2009) studies, with high diversity reported at both the genus
DNA purity (A260/A230)

2.0 and species (Kim et al., 2007) and at the family (O’Neill et al., 2009)
taxonomic levels. However, bacterial diversity was found to be lower
in burned and unburned forest soils amended with oak or grass
1.5
biochar (Khodadad et al., 2011). Lower diversity of archaea (Taketani
and Tsai, 2010) and fungi (Jin, 2010) were found in Terra preta and
1.0 a biochar-amended temperate soil, respectively, compared with
unmodified soils, which indicates that different microbial groups
0.5 respond in different ways. Time since biochar incorporation also
differed between all of these studies, with Khodadad et al. (2011) and
Jin (2010) being short-duration studies of six months and 2.5 years,
respectively, whereas for Terra preta, biochar was incorporated
40
c Modified protocol hundreds to thousands of years ago.
PowersoilTM
UltracleanTM 3.2.1. Community composition
DNA yield (ng µL )
-1

30 Bacterial community composition in soils high in black C or


biochar differs significantly from that in unmodified soils with the
same mineralogy (Kim et al., 2007; O’Neill et al., 2009; Grossman
et al., 2010; Jin, 2010). Kim et al. (2007) compared a biochar-
20
enriched Terra preta soil with a pristine forest soil, both sampled
from the Western Amazon, by use of oligonucleotide fingerprint
grouping. A higher number of unique operational taxonomic units
10
(OTUs) was resolved for the Terra preta (396) as compared to the
forest soil (291), with the former displaying a 25% higher taxonomic
diversity based on several diversity indices. When the forest soil
0 was compared with Terra preta using overall community similari-
0 10 20 30 ties at different phylogenetic distances, all of the OTU diversity in
Biochar application rate (t ha )
-1 the forest soil was found to be represented in the Terra preta. In
contrast, when the Terra preta soil was compared to the adjacent
Fig. 5. Purity of DNA extracts with respect to co-extraction of (a) proteins or (b) humic forest soil, the Terra preta soil was found to contain additional
extracts, and (c) efficiency of extraction (yield) using three different protocols sequences that did not occur in the forest soil. The greatest differ-
(modified from LaMontagne et al., 2002; PowersoilÔ and UltraCleanÔ DNA extraction
ences between the communities in the two soils were found at an
kits from MoBio Laboratories, Carlsbad, CA, USA) as a function of biochar application
rates (n ¼ 3; means and standard deviation) (Jin, 2010). evolutionary distance of 5%, suggesting that these differences were
primarily at the level of genus and species.
In culture-based studies, O’Neill et al. (2009) observed a higher
methodological adaptations before their use in characterizing taxonomic diversity of organisms in the biochar-enriched Terra preta
biochar-amended soils and should include recognition of the vastly soils from four locations in the Central Amazon as compared to
different properties of different biochars. Failure to recognize that adjacent unmodified soils. The greatest taxonomic differences were
modifications to well-known methods are necessary will propagate at the family level. Grossman et al. (2010) showed that bacterial
erroneous results that will be difficult to rectify a posteriori and community composition was most similar among three Terra preta
may confound the literature in this research area. soils formed on Oxisols (divergence 40e70%), which diverged by
over 80% from populations in their respective, unmodified, adjacent
3.2. Community structure of microorganisms soils (Fig. 6). The soils differed in land use history, current land use,
years since formation and other characteristics. Yet, the historical
Given that biochar induces changes in microbial biomass, it is biochar enrichment of the Terra preta soils, hundreds to thousands
extremely unlikely that such overall changes in abundance are of years ago, remained the major driver of bacterial community
spread equally across different phylotypes or functional groups. composition, irrespective of current land use, soil texture, soil
J. Lehmann et al. / Soil Biology & Biochemistry 43 (2011) 1812e1836 1821

Fig. 6. Cluster analysis of T-RFLP fingerprints, based on PCR amplified bacterial 16S rRNA genes, derived from four paired biochar-enriched anthrosols and adjacent unmodified soils
of the same mineralogy (Grossman et al., 2010; with permission).

mineralogy, soil nutrient contents or pH with the notable large soils where no or low biochar was applied. The results suggest that
difference induced by the Spodosol (Fig. 6). While a few additional biochar additions change soil properties such that they support
OTUs were resolved in T-RFLP fingerprints from some of the adjacent communities similar in some respects to the rhizosphere commu-
soils, there was no clear evidence from this analysis that the biochar- nities examined where biochar was not applied.
enriched soils were significantly more or less diverse than their
respective adjacent soils; rather, it was the species composition of 3.2.2. Taxonomy
the communities in the biochar-enriched soils that changed Kim et al. (2007) found two possible new clades of the Acid-
dramatically. obacteria in Terra preta soils; whereas O’Neill et al. (2009) found
Jin (2010) demonstrated that increasing rates of biochar addi- isolates from Terra preta soils representing two possibly new clades
tion to a temperate soil led to increasing divergence in bacterial in the a-Proteobacteria. Sequences obtained in the aforementioned
community composition, in both corn rhizosphere and bulk soils studies reveal that the Acidobacteria were well-represented in both
(Fig. 7). The rhizosphere soils with high biochar application rates soil types (Kim et al., 2007; Grossman et al., 2010). Grossman et al.
(12 and 30 t ha1) were the most dissimilar to bulk soils with little (2010) reported that most of the sequences obtained from the
or no biochar application (0 and 1 t ha1); conversely, the bulk soils Brazilian soils sampled were novel and matched those in databases
receiving high rates of biochar were most similar to the rhizosphere at less than 98% similarity. Several sequences obtained only from
the biochar-enriched Terra preta soils grouped at 93% similarity
with the Verrucomicrobia, a genus commonly found in rice paddies
1.0 in the tropics but increasingly detected in a variety of soils. In this
[12 and 30 t/ha] study, however, sequences closely related to Proteobacteria and
0.8
Cyanobacteria sp. were recovered only from adjacent soil samples.
0.6 Sequences related to Pseudomonas, Acidobacteria, and Flexibacter
0.4 sp. were recovered from both Terra preta and adjacent soils.
IPCA2 (10.98%)

0.2 In a temperate soil with and without additions of biochar made


from corn stover, 70% of the sequences obtained were classified as
0.0
Ascomycota, Basidiomycota or Zygomycota (Jin, 2010). However, the
-0.2 relative gene frequency of the main phylotypes detected differed
Bulk 1 t/ha
-0.4 Bulk 1 t/ha between biochar-amended and unamended soils, with a less
-0.6
Bulk 12 t/ha genetically diverse community found in the biochar-amended soils.
Bulk 30 t/ha
Rhizosphere 0 t/ha
Similarly, Taketani and Tsai (2010) found a less diverse archaeal
-0.8
Rhizosphere 1 t/ha community in Terra preta soils, particularly of ammonia-oxidizing
-1.0 Rhizosphere 12 t/ha Chrenarcheota. To date, only preliminary information exists about
[0 and 1 t/ha] Rhizosphere 30 t/ha
-1.2 the shifts in population of ammonia-oxidizing bacteria in response
-1.0 -0.8 -0.6 -0.4 -0.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2 to biochar accumulation that may be connected to changes in pH
IPCA1 (32.80%) (Ball et al., 2010), which require further experimentation.
Biochar-amended soils had several fold more fungi classified as
Fig. 7. Multivariate analysis of bacterial 16S rRNA gene T-RFLP profiles using HhaI Zygomycota known as glucose and cellulose degraders and Glom-
restriction enzymes. Clear separation of profiles based on the rate of biochar applied eromycota being able to form mycorrhizae, while also having 31%
(separated by a dashed line; IPCA1 and 2) and bulk vs. rhizosphere (separated by
circles, IPCA1) is evident (redrawn after Jin, 2010). Biochar was produced from corn
lower abundance of Basidiomycota and 37% lower abundance of
stalks at 550  C and applied to a loamy Alfisol cropped to corn in May 2007. Soil Ascomycota, than unamended control soils (Jin, 2010). Some Asco-
samples were taken from 0 to 0.15 m depth in October 2008. mycota are known for their ability to degrade lignin but also include
1822 J. Lehmann et al. / Soil Biology & Biochemistry 43 (2011) 1812e1836

many sugar fungi that utilize simple substrates. Lack of available C fresh biochar mainly produced from dairy and bull manure,
in the biochar particles themselves may discourage colonization by increases in both total respiration and metabolic quotient were
these latter fungi, whereas dissolved organic C sorbed to the bio- observed (Kolb et al., 2009). Possible explanations for this behavior
char surface may selectively enrich for Zygomycota, apparently are the high nutrient contents of the manure-based biochar
finding enough easily degradable C sources. including N and P, and a significant proportion of labile organic C in
Similarly, bacterial community changed with biochar additions. the biochar as indicated by the low substrate-induced respiration at
In response to high-temperature biochars (oak wood and grass high biochar additions (Kolb et al., 2009). Deenik et al. (2010) and
pyrolyzed at 650  C) bacterial diversity increased overall diversity Zimmerman (2010) found a direct and positive relationship
and specific taxa, in contrast to results from biochars made at 250  C between the amount of volatile, and hence labile, organic matter in
(Khodadad et al., 2011). The relative abundance of Actinobacteria biochar and CO2 evolved in an incubation experiment. Both
and Gemmatimonadetes was reported to increase in biochar-treated processes mentioned above, i.e., increases in nutrients and labile C,
soils suggesting changes of the community composition in response will likely result from biochar additions to soil, and the net effect on
to the more recalcitrant biochar (Khodadad et al., 2011), consistent biochar mineralization will depend on the proportion of labile C and
with studies on Terra preta (O’Neill et al., 2009) and on char layers the nutrient contents in the biochar applied as well as inorganic
after forest fires (Bääth et al., 1995). This may suggest testing for nutrients available from the soil.
a differential response of fungal and bacterial taxa with respect to It is reasonable to expect an influence of an altered microbial
their preferred energy sources. Although it is too early to draw community structure on the stability of biochar, as well. An
definite conclusions, the available data provide ample grounds for observed shift to a greater abundance of fungi after biochar accu-
interesting hypotheses. mulation in soil may indicate the potential for greater mineraliza-
Very little is known about changes in abundance of specific tion of biochar itself. White rot fungi are known to degrade lignin in
microorganisms (Graber et al., 2010). Ball et al. (2010) found only woody biomass and coal (Willmann and Fakoussa, 1997; Hofrichter
weak evidence that certain ammonia-oxidizing bacteria are affected et al., 1999; Derenne and Largeau, 2001; Hofrichter, 2002). An
by biochar accumulation. Even less information is available on the adaptation of the microbial population to available energy sources
effects of different biochars, which have shown to result in changing is a sensible hypothesis. Interestingly, within the fungi a shift
abundance of different microbial taxa (Khodadad et al., 2011). toward taxa that prefer glucose as an energy source may be
hypothesized, whereas, the opposite was true for bacteria (Section
3.3. Functional ecology of microorganisms with biochar 3.2.2). It is not clear that the much greater recalcitrance of biochar
warrants an adaptation to this food source since more labile organic
Many soil processes may be affected by additions of biochar. matter (particulate organic matter, litter, etc.) is likely still abun-
Denitrification and methane oxidation (Yanai et al., 2007; Van dant in all soil environments. With this background it is under-
Zwieten et al., 2009), C mineralization (Kuzyakov et al., 2009; standable that mineralization of biochar did not increase as a result
Liang et al., 2010) and nutrient transformations (DeLuca et al., of labile C additions, but rather the mineralization of existing non-
2009) were all found to either increase or decrease in the pres- pyrolyzed C (Liang et al., 2010).
ence of biochar. The reasons for such responses may be numerous.
These include altered C sources or nutrient availability, sorption of 3.3.2. Effect of biochar on mineralization of other organic
inorganic and organic compounds including enzymes, different soil matter in soil
water retention and infiltration properties or changes in pore A change in microbial abundance and community structure may
architecture. Here, we consider changes that are mediated by affect not only biochar mineralization itself, but also mineralization
microorganisms in soil: alterations of soil processes as a result of (i) of other soil C. The commonly observed greater microbial biomass
a changing microbial population structure and abundance, and (ii) has been presented as a reason for a greater decomposition of soil C
a direct change in activity and metabolism induced by an altered (also called priming) in the presence of biochar (Wardle et al., 2008).
physical and chemical environment. In some cases, the distinction The fact that this has generally not been observed beyond an initial
between these two types of responses may be blurred or may even greater mineralization after fresh biochar additions (Hamer et al.,
influence each other. From the perspective of improving our 2004; Wardle et al., 2008; Zimmerman et al., 2011) suggests
understanding of the underlying processes, such a differentiated different explanations for the C loss observed in these studies that
view may be an appropriate starting point. may instead be related to physical export of C, changes in nutrient
contents or pH (Lehmann and Sohi, 2008). Also, labile substances in
3.3.1. Mineralization of biochar biochars (such as condensable volatiles as found in smoke) may
As discussed above, the microbial community may show signif- stimulate microbial activity shortly after biochar application to soil
icant responses to biochar additions. A greater microbial abundance (Fischer and Bienkowski, 1999; Uvarov, 2000; Das et al., 2008;
may potentially lead to greater mineralization or oxidation of bio- Steiner et al., 2008a), but these are mineralized within a relatively
char itself as shown for mineralization of non-pyrolyzed organic C short period of time (Cheng et al., 2006). Longer incubations
which is typically stimulated by a greater microbial biomass (Carney (beyond one year) and field trials have shown that biochars decrease
and Matson, 2005). In several reports, however, experimental mineralization of other soil C (Kuzyakov et al., 2009; Kimetu and
results have rather indicated the opposite and not only a lower Lehmann, 2010; Zimmerman et al., 2011). However, the conun-
metabolic quotient (the ratio of microbial activity as measured by drum of greater microbial biomass yet lower soil C respiration still
CO2 production to microbial biomass; Liang et al., 2010; Jin, 2010), warrants closer examination. Interestingly, similar observations of
but also a lower absolute amount of respired C or C turnover greater microbial biomass yet lower metabolism have been made in
(Murage et al., 2007; Kuzyakov et al., 2009; Spokas et al., 2009; Liang waste water treatment, where biofilms on sand showed greater
et al., 2010; Kimetu and Lehmann, 2010; Jin, 2010) or no change removal and mineralization rates of dissolved aromatic C than bio-
(Zackrisson et al., 1996; Haefele et al., 2009; Steiner et al., 2009; Van films on activated carbons (Koch et al., 1991) that typically have
Zwieten et al., 2010b; for compost, Steiner et al., 2010). This could large surface areas (Downie et al., 2009).
result from lower amounts of available C sources, either due to the It is possible that CO2 precipitates as carbonates on biochar
presence of stable biochar or due to the sorption of organic C that surfaces that have high pH and abundant alkaline metals, which
would otherwise be easily degraded. In contrast, after additions of would explain reduced detection of CO2 evolved, despite measured
J. Lehmann et al. / Soil Biology & Biochemistry 43 (2011) 1812e1836 1823

increases in microbial biomass. This is not further examined here, 2006) or toxins in medical applications (Levy, 1982). This demon-
since it is mainly an abiotic precipitation reaction. What is discussed strated co-location of substrate, nutrients (Section 2) and micro-
in more detail here is the possibility that changes in the microbial organisms may result in greater C use efficiency, and thus less
community composition or in enzyme activities are responsible for respired C, which is also supported by lower enzyme activity of C-
lower mineralization of soil C observed with biochar additions. The hydrolyzing enzymes along with higher activity of alkaline phos-
activity of two carbohydrate-mineralizing enzymes was shown to phatase on biochar surfaces (Fig. 9). Biochar particles seem to
decrease after biochar additions to soil (Jin, 2010; Fig. 8). Maximum generate a micro-location in the soil that optimizes resource use for
velocity of both glucosidase and cellobiosidase decreased to very microbial growth. Such a co-location is also well-described in waste
low levels with an application rate of 12 t biochar ha1 or greater. water treatment where organic chemicals had lower toxicity effects
Similar decreases in glucosidase activity were also observed with on microorganisms and were metabolized much quicker when
purified enzymes and fast-pyrolysis biochar produced from sorbed to activated carbons (Ehrhardt and Rehm, 1985). Similar
switchgrass (Bailey et al., 2010). Given the responses shown in Fig. 8, observations were made using trickling filters with activated
application rates between 1 and 12 t ha1 will likely show signifi- carbons which were shown to be more efficient in metabolizing
cant decreases in the activity of some C-mineralizing enzymes. One added compounds than the same microbial community in aqueous
explanation for such change and the associated decrease in respi- batch cultures (Kaplan and Kaplan, 1985). Whether the same
ration may be a co-location of C and microorganisms on biochar mechanism also applies to the soil environment and biochars is not
surfaces that may improve efficiency and reduce the need for certain. It is also possible that such a mechanism leads to oppor-
enzyme production. As seen from greater microbial biomass and the tunities for the well-described syntrophism observed in biofilms
visual assessments of microorganisms colonizing biochars (Fig. 2), (Schink, 1997), where end products from one group of microor-
soil biota are in close contact with biochar surfaces. A decrease in ganisms are readily utilized by another.
enzyme activity by mere sorption to biochar is less likely as shown Alternatively, the sorption of soil organic C by various processes
first by Nelson and Griffin (1916) for non-activated charcoal. For to biochar surfaces may be strong enough to reduce its availability
example, lipases have been shown to sorb well to activated carbon as speculated by Liang et al. (2010) and, hence, to decrease the
matrices with long life and high activity (Quirós et al., 2011). So- ability of exoenzymes to contact, assume proper spatial orientation
called “immobilization” of enzymes on materials such as biochar with and break down the sorbed C. The lower mineralization of
is by now used in many industrial processes that allow stable herbicides and pesticides sorbed to biochars (Yang et al., 2006; Yu
conditions for optimum enzyme activity (Novick and Rozzell, 2005). et al., 2009) and activated carbons (Yang et al., 2009) in soils may
In addition, biochars can sorb large amounts of soil organic C as
shown from batch experiments with microbial cells (Liang et al.,
2010), plant (Miura et al., 2007) or dissolved organic C (DOC)
extracts (Jin, 2010), leaching studies from forest organic horizons
watered with birch litter extracts (Pietikäinen et al., 2000), and
direct observations of biochar surfaces using high-resolution
NEXAFS spectroscopy (Lehmann et al., 2005). These findings are
consistent with the large number of observations often showing
strong sorption of organic compounds such as polyaromatic
hydrocarbons to a variety of black C substrates in soils or sediments
(Cornelissen et al., 2005; Koelmans et al., 2006; Smernik, 2009).
Such sorption may apply both to C from plant litter as well as to
microbial metabolites, may be kinetically limited and therefore
increase over time (Kasozi et al., 2010), but may be weaker than
that documented for polyaromatic compounds (Pignatello et al.,

350
Maxiumum velocity (nmol cm h )
-1

Alkaline Phosphatase
300
-3

250

200 ß-D-Cellobiosidase

150

100 L-Leucine Aminopeptidase

50
ß-D-Glucosidase
0
0 10 20 30
-1
Biochar application rate (t ha )
Fig. 8. Activity of different soil enzymes (0e0.15 m) one year after application of corn Fig. 9. Activity of the enzyme alkaline phosphatase on biochar particle surfaces (corn
stalk biochar (slow pyrolysis at 550  C) at rates of 0, 1, 12, 30 t ha1 to a loamy Alfisol stalk biochar from slow pyrolysis at 550  C), visualized by fluorescence microscopy; (a)
cropped to corn (drawn after Jin, 2010). soil with biochar, (b) soil without biochar (Jin, 2010).
1824 J. Lehmann et al. / Soil Biology & Biochemistry 43 (2011) 1812e1836

provide supporting information for a decreased substrate rhizosphere effects (Fig. 7) indeed suggests a broader effect than
bioavailability in general. Even a hardening of microbial biofilms on merely N and P limitation due to plant nutrient uptake as an
activated C surfaces may need to be considered, as a result of explanation for the greater enzyme activity.
microbial polysaccharides production (Andrews and Tien, 1981).
However, given the increase in microbial biomass after biochar 3.3.4. Nitrous oxide and methane production
additions to soil and the more recent evidence of co-location of The observed varying effects of biochar on N2O and CH4
enzymes, organic matter and microorganisms on biochar surfaces production (Rondon et al., 2005; Yanai et al., 2007; Spokas and
as discussed above, the microbiologically mediated process of Reicosky, 2009; Clough et al., 2010; Singh et al., 2010; Van Zwieten
improved resource use as discussed above, seems a more likely et al., 2010b; Zhang et al., 2010; Knoblauch et al., 2011; Scheer
explanation for observed decreases in CO2 evolution. et al., 2011; Taghizadeh-Toosi et al., 2011) could at least partially be
Substrate use patterns of microorganisms may also change explained by a changing microenvironment for the microbial pop-
through biochar additions to soil as shown with BiologÒ assays of ulation. Biochar may change water relations (Ayodele et al., 2009;
a forest organic horizon (Pietikäinen et al., 2000). Whether this is Hidetoshi et al., 2009; Busscher et al., 2010), which could conceiv-
due to changes in microbial populations or sorption of substrate or ably decrease or increase O2 availability, thereby modifying non-CO2
enzymes, is not entirely clear, but may additionally suggest an GHG emissions (Singh et al., 2010; Van Zwieten et al., 2010b; Zhang
effect of a changing population. A dominance of certain groups of et al., 2010). Also N availability may increase or decrease as discussed
microorganisms, such as coenocytic fungi degrading simple C above. Evidence for reduced nitrous oxide emissions from urin
compounds (e.g., Zygomycota) was observed when corn biochar patches was found by isotope tracing to be at least partially caused
was added to a temperate Alfisol, whereas, abundance of septate by lower N availability after biochar additions to a New Zealand
fungi (such as Basidiomycota (known lignin degraders) and Asco- pasture (Taghizadeh-Toosi et al., 2011).
mycota) decreased (Jin, 2010). An increase in fungi that metabolize Similarly, C availability for microorganisms may change, which
simpler sugars would be in accordance with greater microbial depends on the net effects of C sorption (Miura et al., 2007; Liang
biomass and sorption of labile C compounds on biochar surfaces, et al., 2010), litter production (Major et al., 2010) and the interac-
rather than the inaccessibility of sorbed organic matter. Possibly, tions between water and N availability. In addition to a range of
a lower abundance of degraders of more complex compounds could possible abiotic effects including catalytic reduction with minerals
result in lower decomposition of lignin or aromatic structures of or radicals (as discussed by Van Zwieten et al., 2009) and adsorp-
the biochar itself, increasing its stability, which should be tested in tion of NH3 (Asada et al., 2006), changes in the dominance of either
future experiments. bacterial or fungal communities may play a role in greenhouse gas
production, but no concrete evidence has yet been presented.
3.3.3. Effect of biochar on nutrient transformation In biodigesters, the abundance of anaerobic bacteria was shown
Biochar can have significant effects on microbially-mediated to increase and, as a result, enhance biogas formation when biochar
transformation of nutrients in soil. In forest soils, nitrification was (commercial charcoal) was added (Kumar et al., 1987). This result
increased by biochar additions to soil (DeLuca et al., 2002, 2006; was also obtained with activated carbon (Hunsicker and Almeida,
Berglund et al., 2004; Gundale and DeLuca, 2006; MacKenzie and 1976; Kumar et al., 1987), but not with graphite, carbon black
DeLuca, 2006; Ball et al., 2010) and explained by sorption of pheno- (fossil fuel soot) or petroleum coke (Kumar et al., 1987). The reason
lics that would otherwise inhibit nitrification (Zackrisson et al., 1996; for this increased abundance of anaerobic bacteria under anaerobic
Wardle et al., 1998; Wallstedt et al., 2002; DeLuca et al., 2006) and an conditions in slurries may be similar to the effects discussed for
increase in ammonia-oxidizing bacteria (Ball et al., 2010). Whether microbial abundance in general, including greater resource supply
the observed change in ammonia-oxidizing community composition of C substrates and nutrients, more stable physical conditions,
(Ball et al., 2010) played a role, it not clear. Changes in pH that may better pH buffering and possibly sorption and neutralization of
trigger similar responses in soil were not able to explain the observed harmful substances.
changes in nitrification (DeLuca et al., 2006). On the other hand, Recently, ethylene found to be generated by fresh biochars may
biochar additions to agricultural and grassland soils have shown no be linked to decreases in N2O production (Spokas et al., 2010).
changes or even decreases in net N mineralization (DeLuca et al., Ethylene is both part of the remaining non-aromatic compounds in
2006; Rondon et al., 2007) and lower N availability for plants fresh biochars and is produced by microorganisms in the presence
(Lehmann et al., 2003), likely as a result of N immobilization during of biochar, which may also partly explain observed decreases in CO2
mineralization of a labile fraction of the biochar bearing a high C/N production (Spokas et al., 2010). Non-woody biochar materials
ratio (Deenik et al., 2010). In fact, the greater the mineralizable produced at lower temperatures were found to generate ethylene at
fraction of biochar (often quantified and described as volatile matter), significantly greater rates than soil alone, whereas woody biochars
the greater the N immobilization with resultant decreases in N and activated carbons may have sorbed the generated ethylene. In
uptake and growth of crops (Deenik et al., 2010). A larger microbial the presence of microorganisms ethylene production was 215%
biomass observed with biochar additions will certainly contribute to greater than in sterilized soil (Spokas et al., 2010). Ethylene may
both effects. regulate a series of soil processes (Abeles et al., 1992; Frankenberger
In addition, activity of alkaline phosphatase, aminopeptidase and Arshad, 1995) which need to be investigated further.
and N-acetylglucosaminidase was found to increase with biochar
applications (Bailey et al., 2010; Jin, 2010). Alkaline phosphatase 3.3.5. Microorganisms, biochar and plant growth
increased by 615% and aminopeptidase by 15% with increasing rates Changes in microorganism occurrence and resulting direct effects
of corn biochar application to an Alfisol (Fig. 8; Jin, 2010). This is in on plant growth are only beginning to be explored. Graber et al.
contrast to the decreases in cellobiosidase and glucosidase dis- (2010) demonstrated through phylogenetic characterization of
cussed earlier. Possibly, plant uptake of N and P and growth of fine bacterial isolates based on 16S rRNA gene analysis that of the 20
roots and root hairs into biochar pores (as discussed below) stim- unique identified isolates from the biochar-amended growing media
ulate the production of organic N and P mineralizing enzymes. cropped to pepper and onion, 16 were affiliated with previously
However, N-acetylglucosaminidase activity was also decreased in described plant-growth-promoting and biocontrol agents. The
the absence of plant roots (Bailey et al., 2010). The observation that genus Trichoderma, known for including plant-growth-promoting
biochar induces changes in the bacterial community similar to species, was only isolated from the rhizosphere of pepper when
J. Lehmann et al. / Soil Biology & Biochemistry 43 (2011) 1812e1836 1825

biochar had been added. A possible explanation for the observed Australian Ferrosol that earthworms clearly preferred biochar-
greater crop growth observed by Graber et al. (2010) was therefore amended soil over the controls; however, this preference was not
the promotion of beneficial microorganisms in the rhizosphere. present in a different soil type (Calcarosol) included in the same
experiment. In contrast, Gomez-Eyles et al. (2011) observed a signif-
3.3.6. Electrochemical reactions and biochar icant weight loss of earthworms that had been given hardwood
Carbon materials that are mostly produced at high temperatures biochar in a soil contaminated with PAH relative to the same soil
above 1200  C can have a range of electrochemical properties without biochar. It is not obvious, however, whether the sorbed PAHs
(Portet et al., 2007). The extent to which these properties may be were the root cause of the negative effects of biochars on earth-
selected for by choice of different feedstocks and pyrolysis worms, since at the same time, PAH bioavailability was also reduced.
temperatures is largely unknown (Joseph et al., 2010). However, It is not clear what earthworms gain from ingesting biochar.
there is a wealth of knowledge on the electrochemical properties of Biochar may serve to grind organic matter in their gizzard similar to
carbon (McCreary, 1999). There is also considerable new work on what has been observed for sand (Marhan and Scheu, 2005). Geo-
carbon nanotubes and carbon-based nanomaterials that demon- phagous earthworms may feed on microbes and microbial metab-
strates the types of reactions that can be catalyzed on carbon olites (Lavelle, 1988) which are more abundant on biochar surfaces
surfaces. Liu et al. (2005) demonstrated direct electron transfer to as often shown for soil amended with biochar (discussed above).
glucose oxidase by carbon nanotubes. In another process of rele- Topoliantz and Ponge (2003) also proposed that its ingestion may
vance to extracellular redox reactions, studies with the bacterium favor microbes on which earthworms depend for enzymatic
Shewanella oneidensis have shown facilitated electron transfer rates digestion, or that they profit from detoxifying or pH-ameliorating
from cytochromes located in the bacterial outer membrane via effects of the material.
carbon nanotubes to extracellular electrodes (Peng et al., 2010). Irrespective of the advantage to the earthworm, bioturbation by
Similar processes could be envisioned with biochar. this group of organisms, perhaps mostly by anecic earthworms, is
Within individual particles of biochar, electrochemical proper- likely responsible for vertical mixing of biochar within the soil profile
ties could be expected to be highly variable across microsite loca- (Gouveia and Pessenda, 2000; Carcaillet, 2001). Major et al. (2010),
tions as the local surface properties within a biochar particle will while not quantifying contribution of earthworms to the downward
vary depending on the chemical structures that were pyrolyzed migration of applied biochar in experimental plots, made the
(Amonette and Joseph, 2009). The different functional groups, observation that earthworms were active in the sites, and that the
binding of metals, and metal oxide precipitates will further change inside of earthworm burrows was stained darker than surrounding
the electrical conductivity of carbon surfaces. When carefully soil. Eckmeier et al. (2007) observed natural char particles in earth-
controlled, carbon surfaces can be generated with different surface worm feces at 0.08 m depth 6.5 years after an experimental fire,
oxides and electrochemical behavior. Many properties such as clearly indicating that earthworms may contribute to the movement
those conveyed by different metal oxides may be largely irrevers- of biochar within the soil profile.
ible. These properties ultimately lead to electro-catalytic surfaces Inorganic N concentrations increased to a greater extent when
that can promote electron transfer (Joseph et al., 2010). The extent biochar and earthworms were added together to soil, than if either
to which this would affect various biologically driven redox earthworms or biochar were added alone (Noguera et al., 2010). At
processes on biochar surfaces is a promising research area. the same time, growth and yield of a rice crop also increased the most
if earthworms and biochar were used together. Possibly, microor-
3.4. Faunal population and biochar ganisms in the guts of earthworms are equally more abundant and N-
processing enzymes more active in the presence of biochar as was
The soil fauna are among the least well-studied components of discussed above for soil. A greater microbial biomass and enzyme
the soil biota with respect to biochar effects. This is unfortunate, activity would then increase N release from organic matter in the gut.
since soil fauna may be important in at least three ways. First, soil Alternatively, or in addition, substances inhibitory to N mineraliza-
animals are part of the fungal and bacterial energy channels in the tion and nitrification may have been sorbed to biochars similar to
soil food web (Cragg and Bardgett, 2001), and as such, they may observations made in forest soils (DeLuca et al., 2009).
provide top-down control that is important in order to understand
microbial responses to biochar additions. Second, geophagous 3.4.2. Nematodes
organisms, such as earthworms, could be important modifiers of Data on the response of nematodes to biochar is very limited.
microbial effects to biochar, could modify the biochar material Matlack (2001) carried out an observational study at the landscape
themselves, or could be agents of transport of biochar within the scale, but could not detect a relationship between nematode pop-
soil profile. Finally, soil fauna may react to potentially toxic ulations and charred material in the soil. Direct experimental
components of biochar in ways that are not reflected in the study of evidence for the effects of biochar on nematodes is not yet avail-
microorganisms. able. Soils exposed to smoke from charcoal production increased
density of soil nematodes, increased density and diversity of
3.4.1. Earthworms collembolans, and diversity of oribatid mites (unpublished data
The interaction of earthworms with biochar appears to be the cited in Uvarov, 2000), indicative of the effects of pyrolysis
best-studied among all soil fauna effects. Earthworms clearly ingest condensates present in biochars on soil fauna. Evidence for a posi-
biochar particles. Using a peregrine tropical endogeic earthworm tive relationship between nematodes and biochar is still very weak
species, Pontoscolex corethrurus, Topoliantz and Ponge (2003, 2005) and warrants further study.
demonstrated the ingestion of biochar particles in microcosm
experiments. The earthworms evidently could grind the material and 3.4.3. Microarthropods
mix it into the soil, in fact, preferring soil with biochar over soil alone. As for nematodes, there is a dearth of information on the
The authors even propose that populations of this species may be response of microarthropods to biochar in soil. Using a micromor-
adapted to consumption of charred material and point to the phological approach, Bunting and Lundberg (1987) and Phillips
potential to include this earthworm in management practices et al. (2000) provided evidence that fecal pellets from micro-
involving soils with charred material (Ponge et al., 2006). Using arthropods were deposited within a charcoal-rich layer in forest
a behavioral experiment, Van Zwieten et al. (2010a) showed for an soil, indicating that this material can be ingested and processed by
1826 J. Lehmann et al. / Soil Biology & Biochemistry 43 (2011) 1812e1836

these organisms. However, it is unclear if charcoal ingestion is


incidental or what the microarthropods may gain from its
consumption; perhaps they are consuming fungal hyphae colo-
nizing the biochar. Since microarthropods are part of the fungal
energy channel in the soil food web (Moore et al., 1988), one would
expect increased populations, following a stimulation of fungal
biomass. However, to our knowledge there is no evidence in
support of this.
Bioavailability of pollutants (such as polychlorinated biphenyls,
polyaromatic hydrocarbons) or organic agrochemicals (such as
herbicides and pesticides) to soil fauna may be reduced due to the
strong sorption of non-polar and semi-polar compounds to bio-
chars (Smernik, 2009). There is no information available from
biochar-enriched soils, but ample research with activated carbons
indicates that remediation of pollution in sediments is possible
even at large scale (Cho et al., 2009). Activated carbon has been
shown to decrease availability of pollutants to diverse sets of fauna
such as clams (McLeod et al., 2007), polychaetes (Neanthes arena-
ceodentata) and an amphipod (Leptocheirus plumulosus) (Millward
et al., 2005).

3.5. Biochar and plant roots

Biochar-type materials have been reported to stimulate root Fig. 10. Root system of wheat grown for 15 days in water culture made from various
soil extracts with and without filtering by soot material: (1) fertile clay soil, (2) fertile
growth for some time (Breazeale, 1906; Nutman, 1952). The very
clay soil filtered, (3) poor clay soil, (4) poor clay soil filtered (48 wheat plants grown in
different properties of biochar in comparison to surrounding soil in bottles with aqueous extract for 15 days; changed after Breazeale, 1906).
most known cases improved root growth (Table 3). In fact, roots
may even grow into biochar pores (Lehmann et al., 2003; Joseph
et al., 2010). Makoto et al. (2010) showed not only a significant nutrient or water availability, pH, or aeration will likely improve
increase in root biomass (47%) but also root tip number (64%) root growth. In several cases, not only root and shoot biomass
increased within a layer of char from a forest fire with larch twigs, increased after biochar additions, but the shoot-to-root ratio
birch twigs, and shoots of dwarf bamboo buried in a dystric Cam- increased, as well (Table 3). Such an increase in the shoot-to-root
bisol. The number of storage roots of asparagus also increased with ratio may indicate improved resource supply that requires fewer
coconut biochar additions to a tropical soil (Matsubara et al., 2002). roots to sustain the same above-ground biomass production
Also, root length of rice was shown to increase with biochar addi- (Wilson, 1988). Conversely, lower shoot-to-root ratios at lower
tions (Noguera et al., 2010). Germination and rooting of fir embryos growth rate may indicate lower resource supply. Given the effects
(Abies numidica) significantly increased from 10 to 20% without biochars can have on nutrient and water availability mentioned
additions to 32e80% of embryos when activated carbon was added before, changes in resource supply are likely to play a role in root
to various growth media (Vookova and Kormutak, 2001). Therefore, dynamics. However, decreasing shoot-to-root ratios have also been
not only abundance, but also growth behavior of roots may change reported at increased shoot growth (Table 3). These observations
in response to the presence of biochar. are likely unrelated to resource supply but may need to be
explained by neutralization of a mechanism inhibiting root growth.
3.5.1. Reasons for changes in root growth Already a century ago, Breazeale (1906) and Dachnowski (1908)
The reasons for changes in root growth are rarely well identified have explained the pronounced increase in root growth after
in existing studies, and will likely vary depending on biochar additions of carbon black (soot) to soil with sorption of allelopathic
properties and the conditions that restrict root and shoot growth in compounds that were phytotoxic (Fig. 10). Later experimentation
different soil environments. Biochar with properties that improve including biochar-type material found similar behavior (Skinner
the chemical and physical characteristics of a given soil such as and Beattie, 1916). The effects observed in the former experiment

Table 3
Changes in root mass and shoot-to-root ratios as a result of biochar additions to soil from all available studies (positive values indicating an increase, negative ones a decrease).

Crop Fertilization Soil type Type of biochar (feedstock/pyrolysis Root biomass Above-ground Shoot-to-root References
temperature/application rate) (% change biomass (% change ratio (% change
from control) from control) from control)
Pea PK Compost and peat Unidentified wood/NA/5% w/v 24 37 17 Devonald (1982)
Birch No fertilizer Organic horizons Empetrum hermaphroditum/450/3 t ha1 13 þ29 þ34 Wardle et al. (1998)b
Pine No fertilizer Organic horizons Empetrum hermaphroditum/450/3 t ha1 þ300 þ350 þ58 Wardle et al. (1998)b
Cowpea NPK þ lime or Oxisol Unidentified wood/NA/20% w/w þ17 to þ28 þ68 to þ83 þ44 Lehmann et al. (2003)
no fertilizer
Maize NPK or no NAa Acacia bark/260-350/10 L m-2 þ88 to þ92 þ28 to þ48 23 to 49 Yamato et al. (2006)
fertilizer
Common bean NP þ lime Oxisol Eucalyptus deglupta/350/9% w/w 9.9 to þ9.3 þ3.5 to þ77.4 þ29 to þ37 Rondon et al. (2007)
Rice No fertilizer Inceptisol Eucalyptus deglupta/350/2.6% w/w) þ1 to þ10 þ1 to þ152 þ2 to þ200 Noguera et al. (2010)
and Oxisol and unidentified wood/NA/4.6% w/w
Wheat NP fertilizer Sandy clay loam Eucalyptus/open pan method/1.6e6 t ha1 5 to þ110 25 to þ73 33 to þ58 Solaiman et al. (2010)
a
NA, not available.
b
only Ericaceous site.
J. Lehmann et al. / Soil Biology & Biochemistry 43 (2011) 1812e1836 1827

could not be explained by additions of nutrients, and such additions (Pintos et al., 2010) or sorghum (Nguyen et al., 2007). Fridborg et al.
of activated carbon have been used to neutralize phytotoxic (1978) showed that both embryogenesis and germination of wild
compounds up to now (Inderjit and Callaway, 2003). However, carrot (Daucus carota) and onion (Allium cepa) were significantly
these results have been criticized by Lau et al. (2008), who point out enhanced in the presence of 1% activated carbon. The authors
that they could be due to the creation of artifacts such as nutrients explained the increased growth with a sorption and inactivation of
leaching from the activated carbons and that the addition of benzoic and phenylacetic acid assumed to be excreted by growing
carbonaceous adsorbents may have multiple effects on soil. In cells. Sorption of hydroxymethylfurfural was also identified as
addition, many of the studies lacked a control of only activated a possible growth inhibitor that could be made largely ineffective
carbon additions. No studies have been published where shoot-to- through additions of activated carbon (Weatherhead et al., 1978).
root ratios increased while plant growth decreased, which would However, decreased growth of cultures of soybean (Glycine max)
indicate a direct toxic effect of biochars on plant roots through the and goldenweed (Haplopappus gracilis) (Fridborg and Eriksson,
presence of organic or inorganic (heavy metals) compounds. 1975), as well as root organogenesis of tobacco (Nicotiana taba-
cum) has also been observed (Constantin et al., 1977). Constantin
3.5.2. Interactions with phytotoxic compounds et al. (1977) showed that plant hormones were simultaneously
Phytotoxic compounds may originate from different sources. removed from the culture by the activated carbon. The detrimental
These may, for example, play a role in interactions between different effect of plant hormone sorption by activated carbons has since
root systems. Mahall and Callaway (1992) observed a greater rate of been noted numerous times in culture media (Weatherhead et al.,
root elongation in the presence of activated carbon, partially over- 1978). Still, a clear functional relationship appears to be absent in
coming intraspecific competition of creosote bush (Larrea tri- many of these studies, and its transferability to biochars in soil is
dentata) root systems (Fig. 11). The results were explained by unclear, yet merits further investigation.
a sorption of allelopathic compounds onto the activated carbon, Pan and Van Staden (1998) stated a need to match specific acti-
making these substances ineffective in suppressing neighboring vated carbons to plants or plant growth stages. But systematic
plants of the same or other species. Phytotoxicity may also result studies have rarely been conducted beyond comparisons of different
from phenolic compounds contained in leaf biomass used as brands of activated carbons without information about their
mulches. Some plants are particularly rich in such phenolics. The production conditions or properties. The state of knowledge on
evidence of a suppression of the phytotoxic effects from these biochars is even more disappointing in this respect. Similarly, bio-
compounds by activated carbon is weak, and both alleviation of chars may have multiple and different effects on root growth that
growth suppression as well as no significant effects were found may also occur simultaneously, and their outcome will depend on
(Rutto and Mizutani, 2006; Sampietro and Vattuone, 2006). Some biochar properties, soil type and crop species. Notably, there is
evidence can be gleaned from improved growth of birch (Betula a scarcity of studies that have investigated sorption properties as
pendula) after addition of fire-derived char to ericaceous-rich well as effects of biochars on microbial function in the rhizosphere. It
organic horizons (Wardle et al., 1998). It is not clear whether this is not clear whether single explanations of the observed phenomena
effect was caused by stimulation of nitrification or direct root effects. are sufficient to capture the complex rhizosphere effects of biochars.
Activated carbons have been used for in vitro tissue culture In cases where biochar is added in ecosystems harboring plant
(Klein and Bopp, 1971) and were found both to inhibit or promote communities (e.g., agricultural fields containing crops and weeds,
growth, which was variously attributed to (i) a darkened environ- in a restoration context, etc.), such additions may also trigger
ment; (ii) the sorption of undesirable or inhibitory substances; (iii) changes in the species composition of plant communities (see
sorption of growth regulators and other signaling compounds; or Section 4.1), and thus also of root biomass, with resulting changes
(iv) the release of growth-promoting substances present in or in overall rooting features (e.g., depth, length, architecture). It is
sorbed by activated carbon (Pan and Van Staden, 1998). Specifically, also worth noting that changes in roots could be important for
the improvement of orchid germination in the presence of acti- understanding changes in microbial community composition in the
vated carbon has a long history (Yam et al., 1990), including those affected soils. Some functions of roots have not yet been examined
cultivated for medicinal purposes (Hossain et al., 2009), but has also in this context. For example it is unknown if there are changes in
found application for other plant cultivation as diverse as oak rhizodeposition in response to biochar.

4. Management and risks of biochar for soil biota

4.1. Changing the native soil biota

Biochar is likely to be applied to agroecosystems, in a restoration


context, or in other situations in which the soil biota are either
directly or indirectly already being managed. Perhaps as a conse-
quence, concern about altering the indigenous soil biota has not
been a primary consideration, although perhaps it should be. Bio-
char, when produced, is devoid of biota. However, during storage or
transport inoculation with microbes could occur, which would then
be added e potentially inadvertently e to the target ecosystem.
This could be a particularly important consideration in hydro-
thermal carbonization, because in this case the carbonization
product is wet as a consequence of the production process, and
therefore particularly prone to colonization by microbes, for
Fig. 11. Root elongation of creosote (Larrea tridentate) test roots before and after example molds, during storage (Rillig, unpubl. observation).
contact (Day 0) with roots from another creosote plant, either in the presence or
absence of activated carbon (Mahall and Callaway, 1992; with permission); root
Biochar has not always been beneficial to soil biota abundance.
elongation remained high after contact only in the presence of activated carbon, even For example, even though positive effects have frequently been
when roots were in contact with each other. observed for mycorrhizae and total microbial biomass (see sections
1828 J. Lehmann et al. / Soil Biology & Biochemistry 43 (2011) 1812e1836

above), there are examples of negative impacts as well (Warnock above, more long-term studies are available that report changes in
et al., 2007). It is not valid to conclude from positive effects on plant communities. On 100-year-old charcoal hearths with 25% or
one organism group that a particular biochar will also have similar more charcoal-containing soil, growth of oak was suppressed
positive effects on others. For example, hydrochar can have positive (Mikan and Abrams, 1996) and plant composition on charcoal
effects on arbuscular mycorrhizae, but negative effects on plant deposits changed to a greater abundance of yellow poplar (Mikan
growth (Rillig et al., 2010). and Abrams, 1995). These changes were explained by changes in
Many of the primary concerns of negative effects of biochar on the nutrient availability in the soil. In some instances, changes can
soil biota are associated with a mineralizable or labile fraction often also be explained by enhanced activity of microorganisms, such as
quantified and described as the volatile matter, as well as with salts rhizobia (Vantsis and Bond, 1950; Rondon et al., 2007), which may
such as Na or Cl. These may be short-term effects that need to be lead to an increased abundance of legumes (Anderson and Spencer,
taken seriously in consideration and be evaluated for their suit- 1948; Major et al., 2005, 2010). Effects on root signaling have also
ability as a soil amendment. In previous research, Brown et al. been observed after additions of activated carbon that removed
(1951), Turner (1955) and Gibson and Nutman (1960) used exten- phytotoxic root exudates (Mahall and Callaway, 1992; Callaway and
sive washing procedures of biochar-type charcoals to remove both Aschehoug, 2000). As a result, abundance of invasive weeds was
organic and inorganic substances before application to soil. reduced (Ridenour and Callaway, 2001). Recent research is starting
Without pretreatment, Turner (1955) reported withering of the to identify individual compounds that are released by the plant and
petioles and discoloration of the leaves of clover plants. In addition, subsequently sorbed by biochar. One of the phytotoxic compounds
Devonald (1982) speculated whether the observed decrease in that have been observed to be neutralized was catechin (Bais et al.,
nodule size and abundance could be attributed to some properties 2003). Future research may target such compounds directly to
of the biochar that was applied to peas. identify biochar properties that optimize inactivation of phytotoxic
In the long term, these effects may be of lesser concern as labile root exudates. Possibly, application of specifically designed bio-
organic matter is mineralized and salts are leached from the soil. chars may be used as a selective tool in ecological restoration
Biochar-type substances such as chars produced by vegetation fires (Kulmatiski and Beard, 2006, Kulmatiski, 2010).
are found in almost all soils as already pointed out by Schreiner and
Brown (1912), who identified chars in all studied soils from various 4.2. Biochar as inoculant carrier
parts of the U.S. under agriculture as well as forests. This first
regional assessment of biochar-type materials concentrated on Soil additives and inoculant carriers have been used for, for
larger particulate chars separated by density (Fig. 12). Using spec- example, Azotobacter, Bacillus, Clostridium, Frankia, Pseudomonas, or
troscopic techniques rather than physical separation, there is now Rhizobium (van Elsas and Heijnen, 1991), but little is understood in
ample evidence of a ubiquitous distribution of char in soils, black C terms of their mode of action, even as far as the relatively well-
being found on all major continents (Schmidt et al., 1999; studied rhizobia are concerned (Deaker et al., 2004). Biochar-type
Skjemstad et al., 2002; Krull et al., 2008; Lehmann et al., 2008a). materials have been suggested as inoculant carriers substituting for
Nonetheless, effects on soil biota, rhizosphere ecology and plant the increasingly expensive, rare, greenhouse gas-releasing and non-
communities are also likely in the long term. While information renewable peat for some time (Gukova and Bukevich, 1941; Wu,
about soil microorganisms and fauna is only emerging, as reviewed 1958, 1960; Wu and Kuo, 1969; Tilak and Subba Rao, 1978; Ogawa,
1989; Beck, 1991). Given the previously documented positive
effects of biochar on microorganism abundance and reproduction
rates, use of biochar as an additive to commercial mycorrhizal
inoculum, or even as a carrier material seems promising. But
comprehensive discussions of the mechanisms by which biochar
properties influence inoculant efficiency and survival have not yet
occurred. This is partly because the ability to manipulate biochar
properties to potentially optimize its use as an inoculant carrier has
not been fully recognized in the past. Three aspects play a role for
evaluating the suitability of the carrier material: (i) the survival of
the inoculants during storage; (ii) the survival in the soil; and (iii) the
inoculation efficiency. Typically, past research has focused on
survival during storage, since carrier materials such as peat are
rapidly decomposed in soil and would not improve survival once
added to soil. Biochar, on the other hand, will remain in the soil and
may positively influence abundance of the inoculant organisms as
shown for microbial biomass in general and possibly inoculation
efficiency.
Adding biochar-type residues from vegetation fires to soil
significantly increased nodulation of subterranean clover in
a yellow podzolic soil in Australia (Hely et al., 1957), providing some
of the first evidence for positive effects of biochar on rhizobia.
When biochar was mixed with peat, press mud, compost or farm
yard manure, these inoculant carriers were observed to increase
rhizobial cell counts, viability from 30 days to three months and
nodulation in soybean and pigeon pea (Tilak and Subba Rao, 1978).
Kremer and Peterson (1983) compared different inoculant carriers
for peanut Rhizobia including the commonly used peat material,
Fig. 12. Black C or char particles isolated from soils in the United States (Schreiner and and found biochar materials to perform as well if not better than
Brown, 1912; with permission). peat (Fig. 13). Similarly, Sparrow and Ham (1983) reported that
J. Lehmann et al. / Soil Biology & Biochemistry 43 (2011) 1812e1836 1829

12 Rhizobia inoculant carriers are intended to protect against


Rhizobium survival (Log10 counts g inoculant)

Peat desiccation, adverse pH or toxic substances in soil, be environmen-


Biochar
10 Corn oil tally safe and non-toxic to the target organisms themselves, release
Sucrose phosphate buffer the organisms and be abundant in supply (Stephens and Rask, 2000;
-1

Deaker et al., 2004), all of which may theoretically be achieved with


8
appropriately designed biochars. However, Cassidy et al. (1996)
suggested that encapsulation may be preferable to biochar-type
6 materials because such beads may support inoculants for a longer
period of time due to better protection from environmental stress,
4 greater metabolic activity, and reduced contamination with other
microorganisms. A combination of approaches may prove beneficial.
2
4.3. Pathogens and biochar

0 Circumstantial evidence for the beneficial effect of biochar-type


0 20 40 60 80 100 120 140 materials on suppressing plant diseases, such as potato rot (Allen,
Incubation time (days) 1846a: 382) or rust and mildew (Allen, 1846b: 45) were reported
by farmers more than a century ago, and isolated studies have
Fig. 13. Survival of peanut Rhizobium strain CA001 with different inoculant carrier observed reduced damping off (caused by various pathogens) after
materials at 35  C; pH of biochar 8.7 (n ¼ 2; only sucrose phosphate buffer is signif-
icantly different at P < 0.05; determine by the most-probable-number technique on
additions of charcoal (Retan, 1915). However, little direct experi-
yeast extract-mannitol agar; redrawn after Kremer and Peterson, 1983). mentation has been conducted so far.
Biochar may act in a similar way in suppressing plant diseases as
is described for other organic amendments such as composts.
Several principal mechanisms have been proposed and partly
biochar made from hardwood (pH 8.1) could be used as an inocu- proven for composts (Hoitink and Fahy, 1986; Noble and Coventry,
lant carrier for Rhizobium phaseoli and was superior to peanut hulls, 2005), including (1) a direct release of inhibitors of plant patho-
corn cobs or polyacrylamide gel. The results varied significantly gens; (2) the promotion of microorganisms that act antagonistic to
between peat, vermiculite and biochar, depending on the different pathogens, such as parasites, through production of antibiotics, or
strains of Rhizobia used, indicating that matching carrier properties by successful competition for nutrients; (3) improved plant nutri-
and microorganisms even on the sub-species level is worth tion and vigor, leading to enhanced disease resistance; and (4)
examining. activation of plant defense mechanisms (induced systemic resis-
Khavazi et al. (2007) used biochar to adjust the pH of the tance) by enhancing certain microorganisms. Any and all of these
inoculant carrier material, which was as effective as other carrier four mechanisms may also be applicable to biochar. In addition, (5)
materials and ensured survival of Bradyrhizobium japonicum for the known strong sorption of organic compounds onto biochar may
more than six months at an acceptable level. Beck (1991) showed modify signaling between plant and pathogens, or (6) affect the
that biochars can be mixed with soil and provide a carrier material mobility and activity of the pathogen itself. The following section
that was equally effective as a high-quality peat. Similar results examines the evidence for effects of biochar on pathogens and
were obtained by mixing biochar with composts (Wu, 1960), and possible mechanisms.
10% w/w biochar additions to peat was either similar or increased Fusarium infection of asparagus was found to decrease after
inoculant survival compared to peat alone (Newbould, 1951). addition of coconut biochar and was similar to the benefits derived
Clearly, some properties of biochars offer advantages over using the from manure made from coffee residue (Matsubara et al., 2002). A
very effective peat to improve inoculant survival under some decrease in Fusarium infection of asparagus was also reported after
conditions. Mixing soot into peat outperformed any other mixture addition of biochar made by fast pyrolysis of wood powder (Elmer
with peat (Hedlin and Newton, 1948), suggesting that adsorptive and Pignatello, 2011). Also, infection of tomato with another soil-
properties may play an important role for survival of rhizobia. borne disease, bacterial wilt (Ralstonia solanacearum), was signifi-
With respect to mycorrhizae, Ogawa (1989) mentioned prelim- cantly reduced by adding wood biochar in some experiments and
inary data showing that spores of Gigaspora margarita could be consistently by adding biochar made from municipal biowaste
preserved for more than 180 days at room temperature, if mixed (Nerome et al., 2005). The disease suppression improved with
into balls of biochars made from bark or saw dust. On the other greater application rates of up to 40% (v/v), with benefits persisting
hand, Rutto and Mizutani (2006) found reduced mycorrhizal beyond 90 days after planting.
symbiosis in peach roots when activated carbon was added, which Disease severity of powdery mildew (Leveillula taurica) signifi-
may provide grounds for the hypothesis that high surface area alone cantly decreased after biochar was added to both a sandy soil and
may not provide the properties appropriate for use as an inoculant an organic potting mix (Fig. 14; Elad et al., 2010), suggesting that
carrier material. biochar acts differently from other organic matter with respect to
Some results, however, also showed that peat was clearly the studied disease. The studied biochar produced from citrus
superior to biochar, if the pH of the peat was properly adjusted, as wood using traditional charcoal-making techniques was found to
shown for E. coli as a test organism (Lochhead and Thexton, 1947). induce systemic resistance also to another foliar fungal pathogen,
In all studies reported here without exception, only one single type Botrytis cinerea (gray mold) on pepper and tomato and to the broad
of biochar or activated carbon from unknown origin and unknown mite pest (Polyphagotarsonemus latus) on pepper (Elad et al., 2010).
production conditions was used. Given the dramatically varying Since all three disease agents are spatially separate from the soil-
properties of biochars described earlier (Table 1), it can be expected applied biochar (but on above-ground plant parts), since plant
that different biochar properties will have significantly different nutrition did not differ as a result of biochar additions, and since all
effects on inoculant organisms. It is reasonable to assume that pots were watered equally, induced systemic resistance was sug-
biochars may be designed specifically for certain inoculants and gested to be the most likely explanation. This resistance was
possibly soil conditions. presumed to result from either low-level stress exerted by
1830 J. Lehmann et al. / Soil Biology & Biochemistry 43 (2011) 1812e1836

Fig. 14. Disease severity of powdery mildew (Leveillula taurica) on tomato as affected by biochar additions (a) in both a sand and a coconut fiber-tuff soilless potting medium (mix);
(b) in sand; (c) in the potting medium over 105 days (Elad et al., 2010; with permission).

phytotoxic compounds contained in the biochar (e.g., ethylene and shown. It may be desirable to investigate sorption properties of
propylene glycol) or through larger populations of microorganisms different biochars for model substances that are associated with
isolated from the biochar-treated soils that are known to induce antimicrobial and phytotoxic effects, bearing in mind that these
resistance, such as Trichoderma spp. (Graber et al., 2010). The sorption properties change over time in soil (Cheng et al., 2008;
disease suppression was apparent with the lowest tested applica- Cheng and Lehmann, 2009).
tion rate of 1% (by weight) and was most strongly expressed early in An additional consideration for the management of diseases is
the development of disease symptoms, with a calculated delay the effect of biochar on pesticide efficiency. Pesticides as well as
period of 20 days slowing the disease epidemic (Fig. 14). herbicides may be sorbed to biochars (Zheng et al., 2010) and
Elmer and Pignatello (2011) proposed the following explanation therefore be less effective (Andersen, 1968; Jordan and Smith, 1971;
for their observed decline in Fusarium infection of asparagus: bio- Yang et al., 2006). On the other hand, uptake into crops and
char may have adsorbed allelopathic compounds in replant soil leaching of these substances may be reduced (Yu et al., 2009, 2010)
such as coumaric, caffeic and ferulic acids which led to a measur- which could improve environmental health and food safety. This
able increase in mycorrhizal infection. Greater AM abundance may should be considered when managing plant diseases and weeds
then have led to suppression of the disease. with biochar.
For soil-borne root diseases, it is also conceivable that biochars
reduce compounds in the soil solution that would otherwise facili- 5. Future research
tate the ability of pathogens to detect and infect roots. Root exudates
are known to act as chemoattractants for a range of pathogens such The available literature provides ample justification for further
as Pythium (Jones et al., 1991), and to elicit germination of Pythium investigation into the effects of biochar on microbial, faunal and
spores through linoleic and oleic acids in root exudates (Windstam root abundance, community composition of various biota and their
and Nelson, 2008). Under greenhouse (Callaway and Aschehoug, functions. A greater abundance of microorganisms after biochar
2000) and field conditions (Kulmatiski and Beard, 2006; additions to soil is relatively well established (Table 4). The effects
Kulmatiski, 2010), activated carbon was shown to sorb allelopathic of biochar on soil faunal abundance, however, is barely investigated
compounds produced by plants as discussed above. This was apart from a few studies on earthworms, and resulting community
hypothesized and empirically shown already much earlier using soot composition of the entire soil biota has only been studied to
from fossil fuel combustion (Breazeale, 1906; Schreiner and Reed, a limited extent for microorganisms. Little information is available
1907). It is reasonable to hypothesize similar interferences with for purposeful use of biochars to manage roots, pathogens or
the interactions between roots and soil-borne diseases when using
fresh biochar. Such an interference may possibly include microbial
volatile organic compounds which have been shown to sorb to Table 4
Relative levels of existing knowledge on biochar effects on soil biota and our opin-
biochar-type materials and are being explored for monitoring in-
ions on suggestions for research priorities.
door air quality (Matysik et al., 2009). In addition, the mentioned
ethylene production after biochar is added to soil is known to have Research area Level of existing knowledge Research priority
a significant impact on a range of soil and plant metabolic activities Microbial abundance DDD D
(Spokas et al., 2010). Whether similar effects apply to root and foliar Faunal abundance D DDD
Root abundance DD D
pathogens and pests will need to be proven directly. Microbial community DD DD
However, disease symptoms, plant mortality of Arabidopsis and Faunal community D DD
root colonization by Pseudomonas syringae has also been observed Microbial function D DDD
to increase in the presence of activated charcoal (Bais et al., 2005). Faunal function DDD
Root function
The reason for the greater infection was the sorption of antimi- DD DDD
Biochar inoculants D DDD
crobial compounds exuded by the roots onto the surfaces of the Biochar enzyme interaction D DDD
activated charcoals. These antimicrobial exudates are effective in Biochar pathogen control D DDD
suppressing a range of pathogens (Walker et al., 2003). But the Environmental risk D DDD
relevance of results from the activated carbons with high surface þ, þþ, þþþ, low, medium and high level of existing knowledge or priority for future
areas and low functional surface groups for biochars is yet to be research.
J. Lehmann et al. / Soil Biology & Biochemistry 43 (2011) 1812e1836 1831

microorganisms apart from biochars as inoculant carriers. As for studied in combination. On the short term, characterization stan-
the example of biochar inoculants, none of the available studies dards need to be developed that adequately capture the most
uses more than one type of biochar or gives sufficient information important differences in biochar properties starting with those
about their properties or production conditions. This is symptom- mentioned above. Availability of standard biochar materials to the
atic for much of the available literature on biochar as a whole. Given research community would accelerate the knowledge gain and
the greatly varying properties of different biochars, few advances in ensure comparability between research methods. On the medium
biochar design or insights into processes responsible for observed term, an international research network may prove critical to
changes in soil ecology and biogeochemistry can be expected from address concerns over biochar effects on soil health more expedi-
such studies. Therefore, the feedstock types and production ently under a variety of soil and environmental conditions. A long-
procedures must be systematically varied and investigated for their term research vision building on such a network effort, should
effects on soil biota. include the development of a comprehensive dynamic model that
Often, preliminary information can be gleaned from studies builds on a thorough understanding of biochar properties and their
using activated carbons, which provide some guidance about bio- interactions with soil biota.
chars from woody material with high surface area that are produced
at higher temperatures (including those from gasification), and Acknowledgments
about what is presumably a fraction of surface properties of many
biochars even if they are produced at lower temperatures. None- The manuscript was supported by NSF-BREAD Basic Research in
theless, direct evidence is still to be gathered for biochars. In the Enabling Agricultural development (IOS-0965336) and the New
future, biochar studies on soil biota must include characterization of York State Energy Reseearch and Development Authority (NYSERDA).
a minimum set of properties of the specific biochars. Some general CAM and WCH acknowledge funding from NSF EAR-0911685 and
guidance for biochar characterization was provided by Joseph et al. from DOE SUN grant No. DE-FG36-08GO88073. Any opinions,
(2009), even though it is not clear what biochar properties are findings and conclusions or recommendations expressed in this
important for the control, for example, of microbial abundance at material are those of the authors and do not necessarily reflect the
any given location. Studies specifically on soil biota may at views of the National Science Foundation. The authors express their
a minimum need to document microbially available C, surface area, sincere thanks for help with data presentation by Hongyan Jin and
pore size distribution, pH, ash content, and elemental analyses as for the many useful suggestions by Andy Zimmerman and three
well as production conditions (temperature and time at highest anonymous referees.
temperature) and feedstock type. In addition, contrasting biochars
have to be compared rather than one biochar studied on its own.
Knowledge gaps needing urgent attention include biochar References
effects on faunal abundance (especially micro- and meso-fauna), on
Abeles, F.B., Morgan, P.W., Saltveit Jr., M.E., 1992. Ethylene in Plant Biology.
the ecology of biota including environmental risk, on electro- Academic Press, New York.
chemical properties as well as on the utility as inoculant carriers, on Aciego Pietry, J.C., Brookes, P.C., 2008. Relationships between soil pH and microbial
interactions with enzymes and for managing plant pathogens properties in a UK arable soil. Soil Biology and Biochemistry 40, 1856e1861.
Akiyama, K., Matsuzaki., K.-I., Hayashi, H., 2005. Plant sesquiterpenes induce hyphal
(Table 4). Soil fauna may serve as a useful indicator for environ- branching in arbuscular mycorrhizal fungi. Nature 435, 824e827.
mental risks associated with certain biochar types and could help in Allen, A.B., 1846a. The American Agriculturalist, vol. V. Saxton and Miles, New York.
constraining biochar properties for assessment of its bio-safety. In Allen, R.L., 1846b. Brief Compend of American Agriculture. Saxton and Miles, New
York.
all such studies, appropriate characterization and contrasting of Amelung, W., Zech, W., 1996. Organic species in ped surface and core fractions along
different biochars is critical. a climosequence in the prairie, North America. Geoderma 74, 193e206.
Biochars may influence chemical and physical properties of the Amonette, J., Joseph, S., 2009. Characteristics of biochar e micro-chemical proper-
ties. In: Lehmann, J., Joseph, S. (Eds.), Biochar for Environmental Management:
entire soil, such as water content, aeration or pH. However, this has
Science and Technology. Earthscan, London, pp. 13e32.
been insufficiently clarified whether changes in soil properties after Andersen, A.H., 1968. The inactivation of simazine and linuron in soil by charcoal.
biochar additions are merely an average of soil and biochar prop- Weed Research 8, 58e60.
Anderson, A.J., Spencer, D., 1948. Lime in relation to clover nodulation at sites on the
erties or whether biochar confers distinct changes to surrounding
Southern tablelands of New South Wales. The Journal of the Australian Institute
soil. In any event, all changes induced by biochars will be most of Agricultural Science 14, 39e41.
pronounced close to the surfaces of biochars. Some effects that may Andrews, G.F., Tien, C., 1981. Bacterial film growth in adsorbent surfaces. American
only occur around biochar particles, such as sorption of nutrients Institute of Chemical Engineers Journal 27, 396e403.
Angelini, J., Castro, S., Fabra, A., 2003. Alterations in root colonization and nodC gene
and organic matter, create micro-locations in soil, which we call the induction in the peanut-rhizobia interaction under acidic conditions. Plant
“biochar-sphere”. Such co-location of energy and other resources Physiology and Biochemistry 41, 289e294.
may promote abundance and efficiency of the soil biota. Our tradi- Asada, T., Ohkubo, T., Kawata, K., Oikawa, K., 2006. Ammonia adsorption on bamboo
charcoal with acid treatment. Journal of Health Science 52, 585e589.
tional view typically focuses on bulk soil properties; this perspective Ayodele, A., Oguntunde, P., Joseph, A., de Souza Dias Jr., M., 2009. Numerical analysis of
is chiefly a result of analytical limitations. Such a view is increasingly the impact of charcoal production on soil hydrological behavior, runoff response
deemed inappropriate when studying soil biota (Young and and erosion susceptibility. Revista Brasilieira de Ciencia do Solo 33, 137e145.
Bääth, E., Frostegärd, A., Pennanen, T., Fritze, H., 1995. Microbial community
Crawford, 2004), and inadequate for fostering a better under- structure and pH response in relation to soil organic matter quality in wood-ash
standing of biochar effects in soil, since biochars are mainly partic- fertilized, clear-cut or burned coniferous forest soils. Soil Biology and
ulate (Skjemstad et al., 1996; Nguyen et al., 2008; Lehmann et al., Biochemistry 27, 229e240.
Bailey, V.L., Fansler, S.J., Smith, J.L., Bolton Jr., H., 2010. Reconciling apparent vari-
2008b) and may rather emulate aggregate properties. Microorgan- ability in effects of biochar amendment on soil enzyme activities by assay
isms may directly interact with biochar surfaces and pores as optimization. Soil Biology and Biochemistry 43, 296e301.
demonstrated throughout this paper. Important questions emerge Bais, H.P., Vepachedu, R., Gilroy, S., Callaway, R.M., Vivanco, J.A., 2003. Allelopathy
and exotic plant invasion: from molecules and genes to species interactions.
from a biochar-sphere perspective: How far does the influence of
Science 301, 1377e1380.
biochar reach into the bulk soil? What are the critical soil compo- Bais, H.P., Prithiviraj, B., Jha, A.K., Ausubel, F.M., Vivanco, J.M., 2005. Mediation of
nents and characteristics influenced by its surface properties? pathogen resistance by exudation of antimicrobials from roots. Nature 434,
There will likely not be one single answer to these or other 217e221.
Ball, P.N., MacKenzie, M.D., DeLuca, T.H., Holben, W.E., 2010. Wildfire and charcoal
questions about the net effects of biochars, but answers will vary enhance nitrification and ammonium-oxidizing bacteria abundance in dry
between biochars, soil and plant conditions which should be montane forest soils. Journal of Environmental Quality 39, 1243e1253.
1832 J. Lehmann et al. / Soil Biology & Biochemistry 43 (2011) 1812e1836

Beaton, J.D., Peterson, H.B., Bauer, N., 1960. Some aspects of phosphate adsorption to Constantin, M.J., Henke, R.R., Mansur, M.A., 1977. Effect of activated charcoal on
charcoal. Soil Science Society of America Proceedings 24, 340e346. callus growth and shoot organogenesis in tobacco. In Vitro 13, 293e296.
Beck, D.P., 1991. Suitability of charcoal-amended mineral soil as carrier for Rhizo- Corbin, J.D., Avis, P.G., Wilbur, R.B., 2003. The role of phosphorus availability in the
bium inoculants. Soil Biology and Biochemistry 23, 41e44. response of soil nitrogen cycling, understory vegetation and arbuscular
Bengough, A.G., Mullins, C.E., 1990. Mechanical impedance to root growth: a review mycorrhizal inoculum potential to elevated nitrogen inputs. Water, Air, & Soil
of experimental techniques and root growth responses. Journal of Soil Science Pollution 147, 141e161.
41, 341e358. Cornelissen, G., Gustafsson, Ö, Bucheli, T.D., Jonker, M.T.O., Koelmans, A.A., Van
Berglund, L., DeLuca, T., Zackrisson, O., 2004. Activated carbon amendments to soil Noort, P.C.M., 2005. Extensive sorption of organic compounds to black carbon,
alters nitrification rates in Scots pine forests. Soil Biology and Biochemistry 36, coal, and kerogen in sediments and soils: mechanisms and consequences for
2067e2073. distribution, bioaccumulation, and biodegradation. Environmental Science and
Bird, M.I., Ascough, P.L., Young, I.M., Wood, C.V., Scott, A.C., 2008. X-ray microtomo- Technology 39, 6881e6895.
graphic imaging of charcoal. Journal of Archaeological Sciences 35, 2698e2706. Covacevich, F., Marino, M.A., Echeverri ca, H.E., 2006. The phosphorus source deter-
Birk, J.J., Steiner, C., Teixeira, W.C., Zech, W., Glaser, B., 2009. Microbial response to mines the arbuscular mycorrhizal potential and the native mycorrhizal coloni-
charcoal amendments and fertilization of a highly weathered tropical soil. In: zation of tall fescue and wheatgrass. European Journal of Soil Biology 42, 127e138.
Woods, W.I., Teixeira, W.G., Lehmann, J., Steiner, C., WinklerPrins, A.M.G.A., Cragg, R.G., Bardgett, R.D., 2001. How changes in soil fauna diversity and compo-
Rebellato, L. (Eds.), Amazonian Dark Earths: Wim Sombroek’s Vision. Springer, sition within a trophic group influences decomposition processes. Soil Biology
Berlin, pp. 309e324. and Biochemistry 33, 2073e2081.
Blackwell, P., Krull, E., Butler, G., Herbert, A., Solaiman, Z., 2010. Effect of banded Czimczik, C.I., Preston, C.M., Schmidt, M.W.I., Werner, R.A., Schulze, E.-D., 2002.
biochar on dryland wheat production and fertiliser use in south-western Effects of charring on mass, organic carbon, and stable carbon isotope
Australia: an agronomic and economic perspective. Australian Journal of Soil composition of wood. Organic Geochemistry 33, 1207e1223.
Research 48, 531e545. Dachnowski, A., 1908. The toxic property of bog water and bog soil. Botanical
Bond, D.R., 2010. Electrodes as electron acceptors, and the bacteria who love them. Gazette 46, 130e143.
In: Barton, L.L., Mandl, M., Loy, A. (Eds.), Geomicrobiology: Molecular and Das, K.C., Garcia-Perez, M., Bibens, B., Melear, N., 2008. Slow pyrolysis of poultry
Environmental Perspective. Springer, Berlin, pp. 385e399. litter and pine woody biomass: impact of chars and bio-oils on microbial
Bourke, J., Manley-Harris, M., Fushimi, C., Dowaki, K., Nonoura, T., Antal, M.J., 2007. growth. Journal of Environmental Science and Health A 43, 714e724.
Do all carbonized charcoals have the same chemical structure? 2. A model of Deaker, R., Roughley, R.J., Kennedy, I.R., 2004. Legume seed inoculation tech-
the chemical structure of carbonized charcoal. Industrial and Engineering nologyda review. Soil Biology and Biochemistry 36, 1275e1288.
Chemistry Research 46, 5954e5967. Deenik, J.L., McClellan, T., Uehara, G., Antal, N.J., Campbell, S., 2010. Charcoal volatile
Breazeale, J.F., 1906. Effect of certain solids upon the growth of seedlings in water matter content influences plant growth and soil nitrogen transformations. Soil
cultures. Botanical Gazette 41, 54e63. Science Society of America Journal 74, 1259e1270.
Brewer, C.E., Schmidt-Rohr, K., Satrio, J.A., Brown, R.C., 2009. Characterization of DeLuca, T., Nilsson, M.C., Zackrisson, O., 2002. Nitrogen mineralization and phenol
biochar from fast pyrolysis and gasification systems. Environmental Progress accumulation along a fire chronosequence in northern Sweden. Oecologia 133,
and Sustainable Energy 28, 386e396. 206e214.
Brown, R., Greenwood, A.D., Johnson, A.W., Long, A.G., 1951. The stimulant involved DeLuca, T.H., MacKenzie, M.D., Gundale, M.J., Holben, W.E., 2006. Wildfire-produced
in the germination of Orobanche minor Sm. 1. Assay technique and bulk prep- charcoal directly influences nitrogen cycling in ponderosa pine forests. Soil
aration of the stimulant. Biochemical Journal 48, 559e564. Science Society of America Journal 70, 448e453.
Brussaard, L., de Ruiter, P.C., Brown, G.G., 2007. Soil biodiversity for agricultural DeLuca, T.H., MacKenzie, M.D., Gundale, M.J., 2009. Biochar effects on soil nutrient
sustainability. Agriculture, Ecosystems and Environment 121, 233e244. transformations. In: Lehmann, J., Joseph, S. (Eds.), Biochar for Environmental
Brussaard, L., 1997. Biodiversity and ecosystem functioning in soil. AMBIO 26, Management: Science and Technology. Earthscan, London, pp. 251e270.
563e570. Derenne, S., Largeau, C., 2001. A review of some important families of refractory
Bunting, B.T., Lundberg, J., 1987. The humus profile e concept, class and reality. macromolecules: composition, origin, and fate in soils and sediments. Soil
Geoderma 40, 17e36. Science 166, 833e847.
Busscher, W.J., Novak, J.M., Evans, D.E., Watts, D.W., Niandou, M.A.S., Ahmedna, M., Devonald, V.G., 1982. The effect of wood charcoal on the growth and nodulation of
2010. Influence of pecan biochar on physical properties of a Norfolk loamy sand. garden peas in pot culture. Plant and Soil 66, 125e127.
Soil Science 175, 10e14. Downie, A., Crosky, A., Munroe, P., 2009. Physical properties of biochar. In:
Callaway, R.M., Aschehoug, E.T., 2000. Invasive plants versus their new and old Lehmann, J., Joseph, S. (Eds.), Biochar for Environmental Management: Science
neighbors: a mechanism for exotic invasion. Science 290, 521e523. and Technology. Earthscan, London, pp. 13e32.
Carcaillet, C., 2001. Soil particles reworking evidences by AMS 14C dating of char- Durenkamp, M., Luo, Y., Brookes, P.C., 2010. Impact of black carbon addition to soil
coal. Comptes Rendus de l’Académie des Sciences de Paris, Séries 2, Sciences de on the determination of soil microbial biomass by fumigation extraction. Soil
la Terre et des Planètes 332, 21e28. Biology and Biochemistry 42, 2026e2029.
Carney, K.M., Matson, P.A., 2005. Plant communities, soil microorganisms, and soil Eckmeier, E., Gerlach, R., Skjemstad, J.O., Ehrmann, O., Schmidt, M.W.I., 2007. Minor
carbon cycling: does altering the world belowground matter to ecosystem changes in soil organic carbon and charcoal concentrations detected in
functioning? Ecosystems 8, 928e940. a temperate deciduous forest a year after an experimental slash-and-burn.
Cassidy, M.B., Lee, H., Trevors, J.T., 1996. Environmental applications of immobilized Biogeosciences 4, 377e383.
microbial cells: a review. Journal of Industrial Microbiology 16, 79e101. Ehrhardt, H.M., Rehm, H.J., 1985. Phenol degradation by microorganisms adsorbed
Chan, K.Y., Xu, Z., 2009. Biochar: nutrient properties and their enhancement. In: on activated carbon. Applied Microbiology and Biotechnology 21, 32e36.
Lehmann, J., Joseph, S. (Eds.), Biochar for Environmental Management: Science Elad, Y., Rav David, D., Meller Harel, Y., Borenshtein, M., Ben Kalifa, H., Silber, A.,
and Technology. Earthscan, London, pp. 67e84. Graber, E.R., 2010. Induction of systemic resistance in plants by biochar, a soil-
Chan, K.Y., Van Zwieten, L., Meszaros, I., Downie, A., Joseph, S., 2007. Agronomic applied carbon sequestering agent. Phytopathology 100, 913e921.
values of green waste biochar as a soil amendment. Australian Journal of Soil Elmer, W.H., Pignatello, J.J., 2011. Effect of biochar amendments on mycorrhizal
Research 45, 629e634. associations and Fusarium crown and root rot of asparagus in replant soils. Plant
Chen, H., Yao, J., Wang, F., Choi, M.M.F., Bramanti, E., Zaray, G., 2009. Study on the Disease published online.
toxic effects of diphenol compounds on soil microbial activity by a combination Ensminger, P.W., Culbertson, C.G., Powell, H.M., 1953. Antigenic Hemophilus
of methods. Journal of Hazardous Materials 167, 846e851. pertussis vaccines grown on charcoal agar. The Journal of Infectious Diseases 93,
Cheng, C.H., Lehmann, J., 2009. Ageing of black carbon along a temperature 266e268.
gradient. Chemosphere 75, 1021e1027. Ezawa, T., Yamamoto, K., Yoshida, S., 2002. Enhancement of the effectiveness of
Cheng, C.H., Lehmann, J., Thies, J.E., Burton, S.D., Engelhard, M.H., 2006. Oxidation of indigenous arbuscular mycorrhizal fungi by inorganic soil amendments. Soil
black carbon by biotic and abiotic processes. Organic Geochemistry 37, 1477e1488. Science and Plant Nutrition 48, 897e900.
Cheng, C.H., Lehmann, J., Engelhard, M., 2008. Natural oxidation of black carbon in Fischer, Z., Bienkowski, P., 1999. Some remarks about the effect of smoke from
soils: changes in molecular form and surface charge along a climosequence. charcoal kilns on soil degradation. Environmental Monitoring and Assessment
Geochimica et Cosmochimica Acta 72, 1598e1610. 58, 349e358.
Cho, Y.K., Bailey, J.E., 1978. Immobilization of enzymes on activated carbon: prop- Frankenberger, W.T., Arshad, M., 1995. Phytohormones in SoilsdMicrobial
erties of immobilized glucoamylase, glucose oxidase, and gluconolactonase. Production and Function. Marcel Dekker, New York.
Biotechnology and Bioengineering 10, 1651e1665. Freitas, J.C.C., Emmerich, F.G., Bonagamba, T.J., 2000. High-resolution solid-state
Cho, Y.M., Ghosh, U., Kennedy, A.J., Grossman, A., Ray, G., Tomaszewski, J.E., NMR study of the occurrence and thermal transformations of silicon-
Smithenry, D.W., Bridges, T.S., Luthy, R.G., 2009. Field application of activated containing species in biomass materials. Chemistry of Materials 12, 711e718.
carbon amendment for in-situ stabilization of polychlorinated biphenyls in Freitas, J.C.C., Passamani, E.C., Orlando, M.T.D., Emmerich, F.G., Garcia, F.,
marine sediment. Environmental Science Technology 43, 3815e3823. Sampaio, L.C., Bonagamba, T.J., 2002. Effects of ferromagnetic inclusions on 13C
Clough, T.J., Bertram, J.E., Ray, J.L., Condron, L.M., O’Callaghan, M., Sherlock, R.R., MAS NMR Spectra of heat-treated peat samples. Energy and Fuels 16,
Wells, N.S., 2010. Unweathered wood biochar impact on nitrous oxide emis- 1068e1075.
sions from a bovine-urine-amended pasture soil. Soil Science Society of Fridborg, G., Eriksson, T., 1975. Effects of activated charcoal on growth and
America Journal 74, 852e860. morphogenesis in cell cultures. Physiologia Plantarum 34, 306e308.
Compant, S., Clément, S., Sessitsch, A., 2010. Plant growth-promoting bacteria in the Fridborg, G., Pedersen, M., Landström, L.E., Eriksson, T., 1978. The effect of activated
rhizo- and endosphere of plants: their role, colonization, mechanisms involved charcoal on tissue cultures: adsorption of metabolites inhibiting morphogen-
and prospects for utilization. Soil Biology and Biochemistry 42, 669e678. esis. Physiologia Plantarum 43, 104e106.
J. Lehmann et al. / Soil Biology & Biochemistry 43 (2011) 1812e1836 1833

Fukami, T., Dickie, I.A., Wilkie, J.P., Paulus, B.C., Park, D., Roberts, A., Buchanan, P.K., Jin, H., 2010. Characterization of microbial life colonizing biochar and biochar-
Allen, R.B., 2010. Assembly history dictates ecosystem functioning: evidence amended soils. PhD Dissertation, Cornell University, Ithaca, NY.
from wood decomposer communities. Ecology Letters 13, 675e684. Jones, S.W., Donaldson, S.P., Deacon, J.W., 1991. Behavior of zoospores and zoospore
Gani, S.A., Mukherjee, D.C., Chattoraj, D.K., 1999. Adsorption of biopolymer at solid cysts in relation to root infection by Pythium aphanidermatum. New Phytologist
liquid interfaces. 1. Affinities of DNA to hydrophobic and hydrophilic solid 117, 289e301.
surfaces. Langmuir 15, 7130e7138. Jordan, P.D., Smith, L.W., 1971. Adsorption and deactivation of atrazine and diuron
Gaur, A., Adholeya, A., 2000. Effects of the particle size of soil-less substrates upon by charcoals. Weed Science 19, 541e544.
AM fungus inoculum production. Mycorrhiza 10, 43e48. Joseph, S., Peacocke, C., Lehmann, J., Munroe, P., 2009. Developing a biochar char-
George, N., Davies, J.T., 1988. Parameters affecting adsorption of microorganisms on acterization and test methods. In: Lehmann, J., Joseph, S. (Eds.), Biochar for
activated charcoal cloth. Journal of Chemical Technology and Biotechnology 43, Environmental Management: Science and Technology. Earthscan Publ., London,
173e186. pp. 107e126.
Gibson, A.H., Nutman, P.S., 1960. Studies on the physiology of nodule formation. VII. Joseph, S.D., Camps-Arbestain, M., Lin, Y., Munroe, P., Chia, C.H., Hook, J., van
A reappraisal of the effect of preplanting. Annals of Botany 24, 420e433. Zwieten, L., Kimber, S., Cowie, A., Singh, B.P., Lehmann, J., Foidl, N., Smernik, R.J.,
Glaser, B., Lehmann, J., Zech, W., 2002. Ameliorating physical and chemical prop- Amonette, J.E., 2010. An investigation into the reactions of biochar in soil.
erties of highly weathered soils in the tropics with charcoal e a review. Biology Australian Journal of Soil Research 48, 501e515.
and Fertility of Soils 35, 219e230. Kaplan, D.L., Kaplan, A.M., 1985. Biodegradation of N-nitrosodimethylamine in
Glass, V., Kennett, S.J., 1939. The effect of various forms of particulate carbon on the aqueous and soil systems. Applied and Environmental Microbiology 50,
growth of Gonococcus and Meningococcus. Journal of Pathology and Bacteri- 1077e1086.
ology 49, 125e133. Karaosmanoglu, F., Isigigur-Ergundenler, A., Sever, A., 2000. Biochar from the straw-
Gomez-Eyles, J.L., Sizmur, T., Collins, C.D., Hodson, M.E., 2011. Effects of biochar and stalk of rapeseed plant. Energy and Fuels 14, 336e339.
the earthworm Eisenia fetida on the bioavailability of polycyclic aromatic Kasozi, G.N., Zimmerman, A.R., Nkedi-Kizza, P., Gao, B., 2010. Catechol and humic
hydrocarbons and potentially toxic elements. Environmental Pollution 159, acid sorption onto a range of laboratory-produced black carbons (biochars).
616e622. Environmental Science and Technology 44, 6189e6195.
Gouveia, S.E.M., Pessenda, L.C.R., 2000. Datation par le 14C de charbons inclus dans Keiluweit, M., Nico, P.S., Johnson, M.G., Kleber, M., 2010. Dynamic molecular
le sol pour l’étude du rôle de la remontée biologique de matière et du collu- structure of plant-derived black carbon (biochar). Environmental Science and
vionnement dans la formation de latosols de l’état de São Paulo, Brésil. Comptes Technology 44, 1247e1253.
Rendus de l’Académie des Sciences de Paris, Séries 2, Sciences de la Terre et de Khavazi, K., Rejali, F., Seguin, P., Miransari, M., 2007. Effects of carrier, sterilisation
la Planètes 330, 133e138. method, and incubation on survival of Bradyrhizobium japonicum in soybean
Graber, E.R., Harel, Y.M., Kolton, M., Cytryn, E., Silber, A., David, D.R., Tsechansky, L., (Glycine max L.) inoculants. Enzyme and Microbial Technology 41, 780e784.
Borenshtein, M., Elad, Y., 2010. Biochar impact on development and produc- Khodadad, C.L.M., Zimmerman, A.R., Green, S.J., Uthandi, S., Foster, J.S., 2011. Taxa-
tivity of pepper and tomato grown in fertigated soilless media. Plant and Soil specific changes in soil microbial community composition induced by pyro-
337, 481e496. genic carbon amendments. Soil Biology and Biochemistry 43, 385e392.
Grossman, J.M., O’Neill, B.E., Tsai, S.M., Liang, B., Neves, E., Lehmann, J., Thies, J.E., Killham, K., Firestone, M.K., 1984. Salt stress control of intracellular solutes in
2010. Amazonian anthrosols support similar microbial communities that differ Streptomycetes indigenous to saline soils. Applied and Environmental Microbi-
distinctly from those extant in adjacent, unmodified soils of the same miner- ology 47, 301e306.
alogy. Microbial Ecology 60, 192e205. Killham, K., 1985. A physiological determination of the impact of environmental
Gryndler, M., Larsen, J., Hrselova c, H., Rezaccova c, V., Gryndlerova c, H., Kuba
ct, J., stress on the activity of microbial biomass. Environmental Pollution, Series A
2006. Organic and mineral fertilization, respectively, increase and decrease the 38, 283e294.
development of external mycelium of arbuscular mycorrhizal fungi in a long- Kim, J.-S., Sparovek, S., Longo, R.M., De Melo, W.J., Crowley, D., 2007. Bacterial
term field experiment. Mycorrhiza 16, 159e166. diversity of terra preta and pristine forest soil from the Western Amazon. Soil
Gukova, M.M., Bukevich, W.S., 1941. Influence of aeration and soil temperature on Biology and Biochemistry 39, 648e690.
the development of inoculated and non-inoculated soya plants. Comptes Kimetu, J.M., Lehmann, J., 2010. Stability and stabilization of biochar and green
Rendus de l’Academie des Sciences de l’URSS 31, 937e940. manure in soil with different organic carbon contents. Australian Journal of Soil
Gundale, M.J., DeLuca, T.H., 2006. Temperature and substrate influence the chemical Research 48, 577e585.
properties of charcoal in the ponderosa pine/Douglas-fir ecosystem. Forest Kishimoto, S., Sugiura, G., 1985. Charcoal as a soil conditioner, In: Symposium on
Ecology and Management 231, 86e93. Forest Products Research, International Achievements for the Future. vol. 5: pp.
Haefele, S.M., Knoblauch, C., Gummert, M., Konboon, Y., Koyama, S., 2009. Black 12/23/1e12/23/15.
carbon (biochar) in rice-based systems: characteristics and opportunities. In: Klein, B., Bopp, M., 1971. Effect of activated charcoal in agar on the culture of lower
Woods, W.I., Teixeira, W.G., Lehmann, J., Steiner, C., WinklerPrins, A.M.G.A., plants. Nature 230, 474.
Rebellato, L. (Eds.), Amazonian Dark Earths: Wim Sombroek’s Vision. Springer, Knoblauch, C., Maarifat, A.A., Pfeiffer, E.M., Haefele, S.M., 2011. Degradability of
Berlin, pp. 445e463. black carbon and its impact on trace gas fluxes and carbon turnover in paddy
Hamer, U., Marschner, B., Brodowski, S., Amelung, W., 2004. Interactive priming of soils. Soil Biology and Biochemistry published online.
black carbon and glucose mineralization. Organic Geochemistry 35, 823e830. Koch, B., Ostermann, M., Höke, H., Hempel, D.C., 1991. Sand and activated carbon as
Hedlin, R.A., Newton, J.D., 1948. Some factors influencing the growth and survival biofilm carriers for microbial degradation of phenols and nitrogen-containing
of rhizobia in humus and soil cultures. Canadian Journal of Research 26, aromatic compounds. Water Research 25, 1e8.
174e187. Koelmans, A.A., Jonker, M.T.O., Cornelissen, G., Bucheli, T.D., Van Noort, P.C.M.,
Hely, F.W., Bergersen, F.J., Brockwell, J., 1957. Microbial antagonisms in the rhizo- Gustafsson, Ö, 2006. Black carbon: the reverse of its dark side. Chemosphere 63,
sphere as a factor in the failure of inoculation of subterranian clover. Australian 365e377.
Journal of Agricultural Research 8, 24e44. Kolb, S.E., Fermanich, K.J., Dornbush, M.E., 2009. Effect of charcoal quantity on
Hidetoshi, A., Benjamin, K.S., Haefele, S., Khamdong, S., Koki, H., Yoshiyuki, K., microbial biomass and activity in temperate soils. Soil Science Society of
Yoshio, I., Tatsuhiko, S., Takeshi, H., 2009. Biochar amendment techniques for America Journal 73, 1173e1181.
upland rice production in Northern Laos. 1. Soil physical properties, leaf SPAD Kremer, R.J., Peterson, H.L., 1983. Effects of carrier and temperature on survival of
and grain yield. Field Crops Research 111, 81e84. Rhizobium spp. in legume inocula: development of an improved type of inoc-
Hockaday, W.C., Grannas, A.M., Kim, S., Hatcher, P.G., 2007. The transformation and ulant. Applied and Environmental Microbiology 45, 1790e1794.
mobility of charcoal in a fire-impacted watershed. Geochimica et Cosmochimica Krull, E., Lehmann, J., Skjemstad, J., Baldock, J., Spouncer, L., 2008. The global extent
Acta 71, 3432e3445. of black C in soils: is it everywhere? In: Schröder, Hans G. (Ed.), Grasslands:
Hofrichter, M., Ziegenhagen, D., Sorge, S., Ullrich, R., Bublitz, F., Fritsche, W., 1999. Ecology, Management and Restoration. Nova Science, Hauppauge, NY, pp.
Degradation of lignite (low-rank coal) by lignolytic basidiomycetes and their 13e17.
peroxidase system. Applied Microbiology and Biotechnology 52, 78e84. Kulmatiski, A., Beard, K.H., 2006. Activated carbon as a restoration tool: potential for
Hofrichter, M., 2002. Review: lignin conversion by manganese peroxidase (MnP). control of invasive plants in abandoned agricultural fields. Restoration Ecology
Enzyme and Microbial Technology 30, 454e466. 14, 251e257.
Hoitink, H.A.J., Fahy, P.C., 1986. Basis for the control of soil-borne plant pathogens Kulmatiski, A., 2010. Changing soils to manage plant communities: activated carbon
with composts. Annual Review of Phytopathology 24, 93e114. as a restoration tool in ex-arable fields. Restoration Ecology 19, 102e110.
Hossain, M.M., Sharma, M., Pathak, P., 2009. Cost effective protocol for in vitro mass Kumar, S., Jain, M.C., Chhonkar, P.K.,1987. A note on the stimulation of biogas production
propagation of Cymbidium aloifolium (L.) Sw. e a medicinally important orchid. from cattle dung by addition of charcoal. Biological Wastes 20, 1209e1215.
Engineering in Life Science 9, 444e453. Kuzyakov, Y., Subbotina, I., Chen, H., Bogomolova, I., Xu, X., 2009. Black carbon
Hunsicker, M., Almeida, T., 1976. Powdered activated carbon improves anaerobic decomposition and incorporation into microbial biomass estimated by 14C
digestion. Water Sewage Works 123, 62e63. labeling. Soil Biology and Biochemistry 41, 210e219.
Inderjit, Callaway, R.M., 2003. Experimental designs for the study of allelopathy. Laird, D.A., 2008. The charcoal vision: a winewinewin scenario for simultaneously
Plant and Soil 256, 1e11. producing bioenergy, permanently sequestering carbon, while improving soil
Jackson, R.M., 1958. The ecology of fungi in the soil with special reference to fun- and water quality. Agronomy Journal 100, 178e181.
gistasis. PhD thesis, University of London, Rothamsted Experimental Station, LaMontagne, M.G., Michel, F.C., Holden, P.A., Reddy, C.A., 2002. Evaluation of
Harpenden, UK. extraction and purification methods for obtaining PCR-amplifiable DNA from
Jain, V., Nainawatee, H.S., 2002. Plant flavonoids: signals to legume nodulation and compost for microbial community analysis. Journal of Microbiological Methods
soil microorganisms. Journal of Plant Biochemistry and Biotechnology 11, 1e10. 49, 255e264.
1834 J. Lehmann et al. / Soil Biology & Biochemistry 43 (2011) 1812e1836

Lau, J.A., Puliafico, K.P., Kopshever, J.A., Steltzer, H., Jarvis, E.P., Schwarzländer, M., conditions: a new mechanism involving transmission of physical signals.
Strauss, S.Y., Hufbauer, R.A., 2008. Inference of allelopathy is complicated by Journal of Bacteriology 177, 688e693.
effects of activated carbon on plant growth. New Phytologist 178, 412e423. Matysik, S., Herbarth, O., Mueller, A., 2009. Determination of microbial volatile
Lavelle, P., 1988. Earthworm activities and the soil system. Biology and Fertility of organic compounds (MVOCs) by passive sampling onto charcoal sorbents.
Soils 6, 237e259. Chemosphere 76, 114e119.
Lee, J.G., Cutler, I.B., 1975. Formation of silicon carbide from rice hulls. Ceramics McCreary, R.L., 1999. Electrochemical properties of carbon surfaces. In: Wieckowski, A.
Bulletin 54, 195e198. (Ed.), Interfacial Electrochemistry. Marcel Dekker, Basel, pp. 631e646.
Lehmann, J., Joseph, S., 2009. Biochar for environmental management: an intro- McLeod, P.B., van den Heuvel-Greve, M.J., Luoma, S.N., Luthy, R.G., 2007. Biological
duction. In: Lehmann, J., Joseph, S. (Eds.), Biochar for Environmental Manage- uptake of polychlorinated biphenyls by Macoma balthica from sediment
ment: Science and Technology. Earthscan, London, pp. 1e12. amended with activated carbon. Environmental Toxicology and Chemistry 26,
Lehmann, J., Sohi, S., 2008. Comment on “Fire-derived charcoal causes loss of forest 980e987.
humus”. Science 321, 1295. Mikan, C.J., Abrams, M.D., 1995. Altered forest composition and soil properties of
Lehmann, J., da Silva Jr., J.P., Steiner, C., Nehls, T., Zech, W., Glaser, B., 2003. Nutrient historic charcoal hearths in southeastern Pennsylvania. Canadian Journal of
availability and leaching in an archaeological Anthrosol and a Ferralsol of the Forest Research 25, 687e696.
Central Amazon basin: fertilizer, manure and charcoal amendments. Plant and Mikan, C.J., Abrams, M.D., 1996. Mechanisms inhibiting the forest development of
Soil 249, 343e357. historic charcoal hearths in southeastern Pennsylvania. Canadian Journal of
Lehmann, J., Liang, B., Solomon, D., Lerotic, M., Luizão, F., Kinyangi, F., Schäfer, T., Forestry Research 25, 687e696.
Wirick, S., Jacobsen, C., 2005. Near-edge X-ray absorption fine structure (NEXAFS) Millward, R.N., Bridges, T.S., Ghosh, U., Zimmerman, J.R., Luthy, R.G., 2005. Addition
spectroscopy for mapping nano-scale distribution of organic carbon forms in soil: of activated carbon to sediments to reduce PCB bioaccumulation by a poly-
application to black carbon particles. Global Biogeochemical Cycles 19, GB1013. chaete (Neanthes arenaceodentata) and an amphipod (Leptocheirus plumulosus).
Lehmann, J., Gaunt, J., Rondon, M., 2006. Bio-char sequestration in terrestrial Environmental Science and Technology 39, 2880e2887.
ecosystems e a review. Mitigation and Adaptation Strategies for Global Change Mishulow, L., Sharpe, L.S., Cohen, L.L., 1953. Beef-heart charcoal agar for the
11, 403e427. preparation of Pertussis vaccines. American Journal of Public Health 43,
Lehmann, J., Skjemstad, J.O., Sohi, S., Carter, J., Barson, M., Falloon, P., Coleman, K., 1466e1472.
Woodbury, P., Krull, E., 2008a. Australian climate-carbon cycle feedback Miura, A., Shiratani, E., Yoshinaga, I., Hitomi, T., Hamada, K., Takaki, K., 2007.
reduced by soil black carbon. Nature Geoscience 1, 832e835. Characteristics of the adsorption of dissolved organic matter by charcoals
Lehmann, J., Solomon, D., Kinyangi, J., Dathe, L., Wirick, S., Jacobsen, C., 2008b. carbonized at different temperatures. Japan Agricultural Research Quarterly 41,
Spatial complexity of soil organic matter forms at nanometre scales. Nature 211e228.
Geoscience 1, 238e242. Moore, J.C., Walter, D.E., Hunt, H.W., 1988. Arthropod regulation of micro- and
Lehmann, J., Czimczik, C., Laird, D., Sohi, S., 2009. Stability of biochar in soil. In: mesobiota in below-ground detrital food webs. Annual Reviews of Entomology
Lehmann, J., Joseph, S. (Eds.), Biochar for Environmental Management: Science 33, 419e439.
and Technology. Earthscan, London, pp. 183e205. Moore, J.C., Berlow, E.L., Coleman, D.C., de Ruiter, P.C., Dong, Q., Hastings, A.,
Lehmann, J., 2007a. Bio-energy in the black. Frontiers in Ecology and the Envi- Johnson, N.C., McCann, K.S., Melville, K., Morin, P.J., Nadelhoffer, K.,
ronment 5, 381e387. Rosemond, A.D., Post, D.M., Sabo, J.L., Scow, K.M., Vanni, M.J., Wall, D.H., 2004.
Lehmann, J., 2007b. A handful of carbon. Nature 447, 143e144. Detritus, trophic dynamics and biodiversity. Ecological Letters 7, 584e600.
Leinweber, P., Kruse, J., Walley, F.L., Gillespie, A., Eckardt, K.-U., Blyth, R., Regier, T., Moreno-Castilla, C., 2004. Adsorption of organic molecules from aqueous solutions
2007. Nitrogen K-edge XANES e an overview of reference compounds used to on carbon materials. Carbon 42, 83e94.
identify ’unknown’ organic nitrogen in environmental samples. Journal of Murage, E.W., Voroney, P., Beyaert, R.P., 2007. Turnover of carbon in the free light
Synchrotron Radiation 14, 500e511. fraction with and without charcoal as determined using the 13C natural abun-
Levy, G., 1982. Gastrointestinal clearance of drugs with activated carbon. New dance method. Geoderma 138, 133e143.
England Journal of Medicine 307, 676e678. Nelson, J.M., Griffin, E.G., 1916. Adsorption of invertase. Journal of the American
Liang, B., Lehmann, J., Solomon, D., Kinyangi, J., Grossman, J., O’Neill, B., Chemical Society 38, 1109e1115.
Skjemstad, J.O., Thies, J., Luizão, F.J., Petersen, J., Neves, E.G., 2006. Black carbon Nerome, M., Toyota, K., Islam, T.M., Nishijima, T., Matsuoka, T., Sato, K.,
increases cation exchange capacity in soils. Soil Science Society of America Yamaguchi, Y., 2005. Suppression of bacterial wilt of tomato by incorporation of
Journal 70, 1719e1730. municipal biowaste charcoal into soil. Soil Microorganisms 59, 9e14.
Liang, B., Lehmann, J., Sohi, S.P., Thies, J.E., O’Neill, B., Trujillo, L., Gaunt, J., Newbould, F.H.S., 1951. Studies on humus type legume inoculants. I. Growth and
Solomon, D., Grossman, J., Neves, E.G., Luizão, F.J., 2010. Black carbon affects the survival in storage. Scientific Agriculture 31, 463e469.
cycling of non-black carbon in soil. Organic Geochemistry 41, 206e213. Nguyen, B., Lehmann, J., 2009. Black carbon decomposition under varying water
Liu, Y., Wang, M., Feng, Z., Xu, Z., Dong, S., 2005. The direct electron transfer of regimes. Organic Geochemistry 40, 846e853.
glucose oxidase and glucose biosensor based on carbon nanotubes/chitosan Nguyen, T.V., Thu, T.T., Claeys, M., Angenon, G., 2007. Agrobacterium-mediated
matrix. Biosensors and Bioelectronics 21, 984e988. transformation of sorghum (Sorghum bicolor (L.) Moench) using an improved
Lochhead, A.G., Chase, F.E., 1943. Qualitative studies of soil microorganisms: V. in vitro regeneration system. Plant Cell, Tissue and Organ Culture 91, 155e164.
Nutritional requirements of the predominant bacterial flora. Soil Science 55, Nguyen, B., Lehmann, J., Kinyangi, J., Smernik, R., Riha, S.J., Engelhard, M.H., 2008.
185e196. Long-term black carbon dynamics in cultivated soil. Biogeochemistry 89,
Lochhead, A.G., Thexton, R.H., 1947. Growth and survival of bacteria in peat. I. 295e308.
Powdered peat and related products. Canadian Journal of Research 25, 1e13. Nguyen, B., Lehmann, J., Hockaday, W.C., Joseph, S., Masiello, C.A., 2010. Tempera-
MacKenzie, M.D., DeLuca, T.H., 2006. Charcoal and shrubs modify soil processes in ture sensitivity of black carbon decomposition and oxidation. Environmental
ponderosa pine forests of western Montana. Plant and Soil 287, 257e267. Science and Technology 44, 3324e3331.
Mahall, B.E., Callaway, R.M., 1992. Root communication mechanisms and intra- Noble, R., Coventry, E., 2005. Suppression of soil-borne plant diseases with
community distributions of two Mojave desert shrubs. Ecology 73 2145e1251. composts: a review. Biocontrol Science and Technology 15, 3e20.
Major, J., Steiner, C., DiTommaso, A., Falcão, N.P.S., Lehmann, J., 2005. Weed Noguera, D., Rondon, M., Laossi, K.-R., Hoyos, V., Lavelle, P., de Carvalho, M.H.C.,
composition and cover after three years of soil fertility management in the Barot, S., 2010. Contrasted effect of biochar and earthworms on rice growth
central Brazilian Amazon: compost, fertilizer, manure and charcoal applica- and resource allocation in different soils. Soil Biology and Biochemistry 42,
tions. Weed Biology and Management 5, 69e76. 1017e1027.
Major, J., Lehmann, J., Rondon, M., Goodale, C., 2010. Fate of soil-applied black Novick, S.J., Rozzell, J.D., 2005. Immobilization of enzymes by covalent attachment.
carbon: downward migration, leaching and soil respiration. Global Change In: Barredo, J.L. (Ed.), Microbial Enzymes and Biotransformations. Methods in
Biology 16, 1366e1379. Biotechnology, vol. 17. Springer, Berlin, pp. 247e271.
Makoto, K., Tamai, Y., Kim, Y.S., Koike, T., 2010. Buried charcoal layer and ectomy- Nutman, P.S., 1952. Host factors influencing infection and nodule development in
corrhizae cooperatively promote the growth of Larix gmelinii seedlings. Plant leguminous plants. Proceedings of the Royal Society of London, Series B, Bio-
and Soil 327, 143e152. logical Sciences 139, 176e185.
Malik, K.A., 1990. Use of activated charcoal for the preservation of anaerobic pho- O’Neill, B., Grossman, J., Tsai, M.T., Gomes, J.E., Lehmann, J., Peterson, J., Neves, E.,
totrophic and other sensitive bacteria by freeze-drying. Journal of Microbio- Thies, J.E., 2009. Bacterial community composition in Brazilian Anthrosols and
logical Methods 12, 117e124. adjacent soils characterized using culturing and molecular identification.
Marhan, S., Scheu, S., 2005. Effects of sand and litter availability on organic matter Microbial Ecology 58, 23e35.
decomposition in soil and in casts of Lumbricus terrestris L. Geoderma 128, O’Neill, B., 2007. Microbial communities in Amazonian Dark Earth soils analyzed
155e166. by culture-based and molecular approaches. MS thesis, Cornell University,
Matlack, G.R., 2001. Factors determining the distribution of soil nematodes in Ithaca NY.
a commercial forest landscape. Forest Ecology and Management 146, 129e143. Ogawa, M., Okimori, Y., 2010. Pioneering works in biochar research, Japan.
Matsubara, Y., Hasegawa, N., Fukui, H., 2002. Incidence of Fusarium root rot in Australian Journal of Soil Research 48, 489e500.
asparagus seedlings infected with arbuscular mycorrhizal fungus as affected by Ogawa, M., 1989. Inoculation methods of VAM fungi: charcoal ball method and rice
several soil amendments. Journal of the Japanese Society for Horticultural hull method. In: Hattori, T., Ishida, Y., Maruyama, Y., Morita, R.Y., Uchida, A. (Eds.),
Science 71, 370e374. Recent Advances in Microbial Ecology. Proceedings of the 5th Int. Symp on
Matsuhashi, M., Pankrushina, A.N., Endoh, K., Watanabe, H., Mano, Y., Hyodo, M., Microbial Ecology (ISME 5). Japan Scientific Societies Press, Tokyo, pp. 247e252.
Fujita, T., Kunugita, K., Kaneko, T., Otani, S., 1995. Studies on carbon material Ogawa, M., 1994. Symbiosis of people and nature in the tropics. III. Tropical agri-
requirements for bacterial proliferation and spore germination under stress culture using charcoal. Farming Japan 28, 21e35.
J. Lehmann et al. / Soil Biology & Biochemistry 43 (2011) 1812e1836 1835

Otsuka, S., Sudiana, I., Komori, A., Isobe, K., Deguchi, S., Nishiyama, M., Shimizu, H., Schmidt, M.W.I., Noack, A.G., 2000. Black carbon in soils and sediments: analysis,
Senoo, K., 2008. Community structure of soil bacteria ina tropical rainforest distribution, implications, and current challenges. Global Biogeochemical Cycles
several years after fire. Microbes and Environment 23, 49e56. 14, 777e793.
Özçimen, D., Karaosmanolu, F., 2004. Production and characterization of bio-oil and Schmidt, M.W.I., Skjemstad, J.O., Gehrt, E., Kögel-Knabner, I., 1999. Charred organic
biochar from rapeseed cake. Renewable Energy 29, 779e787. carbon in German chernozemic soils. European Journal of Soil Science 50,
Pan, M.J., Van Staden, J., 1998. The use of charcoal in in vitro culture e a review. 351e365.
Plant Growth Regulation 26, 155e163. Schreiner, O., Brown, B.E., 1912. Occurrence and Nature of Carbonized Material in
Peng, L., You, S.-J., Wang, J.-Y., 2010. Carbon nanotubes as electrode modifier Soil. United States Department of Agriculture, Bureau of Soils. Bulletin No. 90.
promoting direct electron transfer from Shewanella oneidensis. Biosensors and Schreiner, O., Reed, H.S., 1907. The production of deleterious excretions by roots.
Bioelectronics 25, 1248e1251. Bulletin of the Torrey Botanical Club 34, 279e303.
Phillips, D.H., Foss, J.E., Buckner, E.R., Evans, R.M., FitzPatrick, E.A., 2000. Response of Schwartz, M.W., Hoeksema, J.D., Gehring, C.A., Johnson, N.C., Klironomos, J.N.,
surface horizons in an oak forest to prescribed burning. Soil Science Society of Abbott, L.K., Pringle, A., 2006. The promise and the potential consequences of the
America Journal 64, 754e760. global transport of mycorrhizal fungal inoculum. Ecological Letters 9, 501e515.
Pietikäinen, J., Kiikkilä, O., Fritze, H., 2000. Charcoal as a habitat for microbes and its Sexstone, A.J., Revsbech, N.P., Parkin, T.P., Tiedje, J.M., 1985. Direct measurement of
effects on the microbial community of the underlying humus. Oikos 89, oxygen profiles and denitrification rates in soil aggregates. Soil Science Society
231e242. of America Journal 49, 645e651.
Pignatello, J.J., Kwon, S., Lu, Y., 2006. Effect of natural organic substances on the Singh, B.P., Hatton, B.J., Singh, B., Cowie, A.L., Kathuria, A., 2010. Influence of bio-
surface and adsorptive properties of environmental black carbon (char): chars on nitrous oxide emission and nitrogen leaching from two contrasting
attenuation of surface activity by humic and fulvic acids. Environmental Science soils. Journal of Environmental Quality 39, 1224e1235.
and Technology 40, 7757e7763. Skinner, J.J., Beattie, J.H., 1916. A study on the action of carbon black and similar
Pintos, B., Manzanera, J.A., Bueno, M.A., 2010. Oak somatic and gametic embryos adsorbing materials in soils. Soil Science 2, 93e102.
maturation is affected by charcoal and specific aminoacids mixture. Annals of Skjemstad, J.O., Clarke, P., Taylor, J.A., Oades, J.M., McClure, S.G., 1996. The chemistry
Forest Science 67 Article number 205. and nature of protected carbon in soil. Australian Journal of Soil Research 34,
Pollock, M.R., 1947. The growth of H. pertussis on media without blood. British 251e271.
Journal of Experimental Pathology 28, 295e307. Skjemstad, J.O., Reicosky, D.C., Wilts, A.R., McGowen, J.A., 2002. Charcoal carbon in
Ponge, J.-F., Ballot, S., Rossi, J.-P., Lavelle, P., Betsch, J.-M., Gaucher, P., 2006. Ingestion U.S. agricultural soils. Soil Science Society of America Journal 66, 1249e1255.
of charcoal by the Amazonian earthworm Pontoscolex corethrurus: a potential Smernik, R., 2009. Biochar and sorption of organic compounds. In: Lehmann, J.,
for tropical soil fertility. Soil Biology and Biochemistry 38, 2008e2009. Joseph, S. (Eds.), Biochar for Environmental Managament: Science and Tech-
Portet, C., Yushin, G., Gogotsi, Y., 2007. Electrochemical performance of carbon nology. Earthscan, London, pp. 289e300.
onions, nanodiamonds, carbon black and multiwalled nanotubes in electrical Sohi, S., Krull, E., Lopez-Capel, E., Bol, R., 2010. A review of biochar and its use and
double layer capacitors. Carbon 45, 2511e2518. function in soil. Advances in Agronomy 105, 47e82.
Quirós, M., Garci, A.B., Montes-Moran, M.A., 2011. Influence of the support surface Solaiman, Z.M., Blackwell, P., Abbott, L.K., Storer, P., 2010. Direct and residual effect
properties on the protein loading and activity of lipase/mesoporous carbon of biochar application on mycorrhizal colonization, growth and nutrition of
biocatalysts. Carbon 49, 406e415. wheat. Australian Journal of Soil Research 48, 546e554.
Rapaport, E., Zamecnik, P.C., Baril, E.F., 1981. HeLa cell DNA polymerase alpha is Sparrow Jr., S.D., Ham, G.E., 1983. Survival of Rhizobium phaseoli in six carrier
tightly associated with tryptophanyl-tRNA synthetase and diadenosine 50 , 500 0 - materials. Agronomy Journal 75, 181e184.
P1, P4-tetraphosphate binding activities. Proceedings of the National Academy Spokas, K.A., Reicosky, D.C., 2009. Impacts of sixteen different biochars on soil
of Sciences of the United States of America 78, 838e842. greenhouse gas production. Annals of Environmental Science 3, 179e193.
Retan, G.A., 1915. Charcoal as a means of solving some nursery problems. Forestry Spokas, K.A., Koskinen, W.C., Baker, J.M., Reicosky, D.C., 2009. Impacts of woodchip
Quarterly 13, 25e30. biochar additions on greenhouse gas production and sorption/degradation of
Ridenour, W.M., Callaway, R.M., 2001. The relative importance of allelopathy in two herbicides in a Minnesota soil. Chemosphere 77, 574e581.
interference: the effects of an invasive weed on a native bunchgrass. Oecologia Spokas, K.A., Baker, J.M., Reicosky, D.C., 2010. Ethylene: potential key for biochar
126, 444e450. amendment impacts. Plant and Soil 333, 443e452.
Rillig, M.C., Mummey, D.L., 2006. Mycorrhizas and soil structure. New Phytologist Steiner, C., Teixeira, W.G., Lehmann, J., Zech, W., 2004. Microbial response to char-
171, 41e53. coal amendments of highly weathered soils and Amazonian Dark Earths in
Rillig, M.C., Wagner, M., Salem, M., Antunes, P.M., George, C., Ramke, H.G., Central Amazonia e preliminary results. In: Glaser, B., Woods, W.I. (Eds.),
Titirici, M.M., Antonietti, M., 2010. Material derived from hydrothermal Amazonian Dark Earths: Explorations in Time and Space. Springer, Berlin,
carbonization: effects on plant growth and arbuscular mycorrhiza. Applied Soil Germany, pp. 195e212.
Ecology 45, 238e242. Steiner, C., Das, K.C., Garcia, M., Förster, B., Zech, W., 2008a. Charcoal and smoke
Rivera-Utrilla, J., Bautilsta-Toledo, I., Ferro-Carcia, M.A., Moreno-Catilla, C., 2001. extract stimulate the soil microbial community in a highly weathered xanthic
Activated carbon surface modifications by adsorption of bacteria and their ferralsol. Pedobiologia-International Journal of Soil Biology 51, 359e366.
effect on aqueous lead adsorption. Journal of Chemical Technology and Steiner, C., Glaser, B., Teixeira, W.G., Lehmann, J., Blum, W.E.H., Zech, W., 2008b.
Biotechnology 76, 1209e1215. Nitrogen retention and plant uptake on a highly weathered central Amazonian
Ron, E.Z., Rosenberg, E., 2001. Natural roles of biosurfactants. Environmental Ferralsol amended with compost and charcoal. Journal of Plant Nutrition and
Microbiology 3, 229e236. Soil Science 171, 893e899.
Rondon, M., Ramirez, J.A., Lehmann, J., 2005. Greenhouse gas emissions decrease Steiner, C., Garcia, M., Zech, W., 2009. Effects of charcoal as slow release nutrient
with charcoal additions to tropical soils. In: Proceedings of the Third USDA carrier on NePeK dynamics and soil microbial population: pot experiments
Symposium on Greenhouse Gases and Carbon Sequestration. Soil Carbon Center, with ferralsol substrate. In: Woods, W.I., Teixeira, W.G., Lehmann, J., Steiner, C.,
Kansas State University, United States Department of Agriculture, Baltimore, p. 208. WinklerPrins, A.M.G.A., Rebellato, L. (Eds.), Amazonian Dark Earths: Wim
http://soilcarboncenter.k-state.edu/conference/USDA%2520Abstracts%2520html/ Sombroek’s Vision. Springer, Berlin, pp. 325e338.
Abstract%2520Rondon.htm. Steiner, C., Das, K.C., Melear, N., Lakly, D., 2010. Reduced nitrogen loss during poultry
Rousk, J., Brookes, P.C., Bååth, E., 2009. Contrasting soil pH effects on fungal and litter composting using biochar. Journal of Environmental Quality 39, 1236e1242.
bacterial growth suggest functional redundancy in carbon mineralization. Stephens, J.H.G., Rask, H.M., 2000. Inoculant production and formulation. Field
Applied and Environmental Microbiology 75, 1589e1596. Crops Research 65, 249e258.
Rousk, J., Bååth, E., Brookes, P.C., Lauber, C.L., Lozupone, C., Caporaso, J.G., Knight, R., Taghizadeh-Toosi, A., Clough, T.J., Condron, L.M., Sherlock, R.R., Anderson, C.R.,
Fierer, N., 2010. Soil bacterial and fungal communities across a pH gradient in Craigie, R.A., 2011. Biochar incorporation into pasture soil suppresses in situ
an arable soil. The ISME 4, 134e151. nitrous oxide emissions from ruminant urine patches. Journal of Environmental
Rutto, K.L., Mizutani, F., 2006. Effect of mycorrhizal inoculation and activated carbon Quality 40, 468e476.
on growth and nutrition in peach (Prunus persica Batsch) seedlings treated with Taketani, R.G., Tsai, S.M., 2010. The influence of different land uses on the structure
peach root bark extracts. Journal of the Japanese Society of Horticultural of archaeal communities in Amazonian Anthrosols based on 16S rRNA and
Sciences 75, 463e468. amoA genes. Microbial Ecology 59, 734e743.
Saito, M., Marumoto, T., 2002. Inoculation with arbuscular mycorrhizal fungi: the Taylor, C.B., 1951. The nutritional requirements of the predominant bacterial flora of
status quo in Japan and the future prospects. Plant and Soil 244, 273e279. soil. Journal of Applied Microbiology 14, 101e111.
Samonin, V.V., Elikova, E.E., 2004. A study of the adsorption of bacterial cells on Thies, J.E., Rillig, M., 2009. Characteristics of biochar: biological properties. In:
porous materials. Microbiology 73, 696e701. Lehmann, J., Joseph, S. (Eds.), Biochar for Environmental Management: Science
Sampietro, D.A., Vattuone, M.A., 2006. Nature of the interference mechanism of and Technology. Earthscan, London, pp. 85e105.
sugarcane. Plant and Soil 280, 157e169. Tilak, K.V.B.R., Subba Rao, N.S., 1978. Carriers for legume (rhizobium) inoculants.
Scheer, C., Grace, P.R., Rowlings, D.W., Kimber, S., Van Zwieten, L., 2011. Effect of Fertilizer News 23, 25e28.
biochar amendment on the soil-atmosphere exchange of greenhouse gases Tisdall, J.M., Oades, J.M., 1982. Organic matter and water-stable aggregates in soils.
from an intensive subtropical pasture in northern New South Wales, Australia. Journal of Soil Science 33, 141e163.
Plant and Soil published online. Topoliantz, S., Ponge, J.F., 2003. Burrowing activity of the geophagous earthworm
Schimel, J., Balser, T.C., Wallenstein, M., 2007. Microbial stress-response physiology Pontoscolex corethurus (Oligochaeta: Glossoscolecidae) in the presence of
and its implications for ecosystem function. Ecology 88, 1386e1394. charcoal. Applied Soil Ecology 23, 267e271.
Schink, B., 1997. Energetics of syntrophic cooperation in methanogenic degradation. Topoliantz, S., Ponge, J.F., 2005. Charcoal consumption and casting activity by
Microbiology and Molecular Biology Reviews 61, 262e280. Pontoscolex corethurus (Glossoscolecidae). Applied Soil Ecology 28, 217e224.
1836 J. Lehmann et al. / Soil Biology & Biochemistry 43 (2011) 1812e1836

Treseder, K.K., Allen, M.F., 2002. Direct nitrogen and phosphorus limitation of arbus- Wu, M.M.H., Kuo, K.C., 1969. Influence of autoclaved compost carrier on the survival
cular mycorrhizal fungi: a model and field test. New Phytologist 155, 507e515. of rhizobia for legume inoculant. Soils and Fertilizers in Taiwan 16, 46e51.
Turner, E.R., 1955. The effect of certain adsorbents on the nodulation of clover Wu, M.M.H., 1958. Studies on Rhizobium inoculation of legumes in Taiwan. Journal
plants. Annals of Botany 19, 149e160. of Agriculture and Forestry 7, 1e48.
Uvarov, A., 2000. Effects of smoke emissions from a charcoal kiln on the functioning Wu, M.M.H., 1960. Research on powdered legume inoculant of root nodule bacteria
of forest soil systems: a microcosm study. Environmental Monitoring and in Taiwan. Journal of the Agricultural Association of China 32, 48e58.
Assessment 60, 337e357. Yam, T.Y., Ernst, R., Arditti, J., Nair, H., Weatherhead, M.A., 1990. Charcoal in
Van Elsas, J.D., Heijnen, C.E., 1991. Methods for the introduction of bacteria into soil: orchid seed germination and tissue culture media: a review. Lindleyana 5,
a review. Biology and Fertility of Soils 10, 127e133. 256e265.
Van Zwieten, L., Singh, B.P., Joseph, S., Kimber, S., Cowie, A., Chan, K.Y., 2009. Biochar Yamamoto, N., Lopez, G., 1985. Bacterial abundance in relation to surface area and
and emissions of non-CO2 greenhouse gases from soil. In: Lehmann, J., Joseph, S. organic content of marine sediments. Journal of Experimental Marine Biology
(Eds.), Biochar for Environmental Management: Science and Technology. and Ecology 90, 209e220.
Earthscan, London, pp. 227e249. Yamato, M., Okimori, Y., Wibowo, I.F., Anshori, S., Ogawa, M., 2006. Effects of the
Van Zwieten, L., Kimber, S., Morris, S., Chan, K.Y., Downie, A., Rust, J., Joseph, S., application of charred bark of Acacia mangium on the yield of maize, cowpea
Cowie, A., 2010a. Effects of biochar from slow pyrolysis of papermill waste on and peanut, and soil chemical properties in South Sumatra, Indonesia. Soil
agronomic performance and soil fertility. Plant and Soil 327, 235e246. Science and Plant Nutrition 52, 489e495.
Van Zwieten, L., Kimber, S., Morris, S., Downie, A., Berger, E., Rust, J., Scheer, C., Yanai, Y., Toyota, K., Okazaki, M., 2007. Effects of charcoal addition on N2O emissions
2010b. Influence of biochars on flux of N2O and CO2 from Ferrosol. Australian from soil resulting from rewetting air-dried soil in short-term laboratory
Journal of Soil Research 48, 555e568. experiments. Soil Science and Plant Nutrition 53, 181e188.
Vantsis, J.T., Bond, G., 1950. The effect of charcoal on the growth of leguminous Yang, Y.N., Sheng, G.Y., Huang, M.S., 2006. Bioavailability of diuron in soil
plants in sand culture. Annals of Applied Biology 37, 159e168. containing wheatestraw-derived char. Science of the Total Environment
Vookova, B., Kormutak, A., 2001. Effect of sucrose concentration, charcoal, and 354, 170e178.
indole-3-butyric acid on germination of Abies numidica somatic embryos. Yang, Y., Hunter, W., Tai, S., Crowley, D., Gan, J., 2009. Effect of activated carbon on
Biologia Plantarium 44, 181e184. microbial bioavailability of phenanthrene in soil. Environmental Toxicology and
Walker, T.S., Bais, H.P., Grotewold, E., Vivanco, J.M., 2003. Root exudation and Chemistry 28, 2283e2288.
rhizosphere biology. Plant Physiology 132, 44e51. Yin, B., Crowley, D., Sparovek, G., De Melo, W.J., Borneman, J., 2000. Bacterial
Wallstedt, A., Coughlan, A., Munson, A.D., Nilsson, M.-C., Margolis, H.A., 2002. functional redundancy along a soil reclamation gradient. Applied and Envi-
Mechanisms of interaction between Kalmia angustifulia cover and Picea mariana ronmental Microbiology 66, 4361e4365.
seedlings. Canadian Journal of Forest Research 32, 2022e2031. Young, I.M., Crawford, J.W., 2004. Interactions and self-organization in the
Wardle, D.A., Zackrisson, O., Nilsson, M.C., 1998. The charcoal effect in boreal soilemicrobe complex. Science 304, 1634e1637.
forests: mechanisms and ecological consequences. Oecologia 115, 419e426. Yu, X., Ying, G., Kookana, R.S., 2009. Plant uptake of pesticides with biochar addi-
Wardle, D.A., Nilsson, M.C., Zackrisson, O., 2008. Fire-derived charcoal causes loss of tions to soil. Chemosphere 76, 665e671.
forest humus. Science 320 629e629. Yu, X., Pan, L., Ying, G., Kookana, R.S., 2010. Enhanced and irreversible sorption of
Warnock, D.D., Lehmann, J., Kuyper, T.W., Rillig, M.C., 2007. Mycorrhizal responses pesticide pyrimethanil by soil amended with biochars. Journal of Environ-
to biochar in soil e concepts and mechanisms. Plant and Soil 300, 9e20. mental Sciences 22, 615e620.
Warnock, D.D., Mummey, D.L., McBride, B., Major, J., Lehmann, J., Rillig, M.C., 2010. Zackrisson, O., Nilsson, M.C., Wardle, D.A., 1996. Key ecological function of charcoal
Influences of non-herbaceous biochar on arbuscular mycorrhizal fungal abun- from wildfire in the boreal forest. Oikos 77, 10e19.
dances in roots and soils: results from growth-chamber and field experiments. Zhang, A., Cui, L., Pa, G., Li, L., Hussain, Q., Zhang, X., Zheng, J., Crowley, D., 2010.
Applied Soil Ecology 46, 450e456. Effect of biochar amendment on yield and methane and nitrous oxide emissions
Weatherhead, M.A., Burdon, J., Henshaw, G.G., 1978. Some effects of activated from a rice paddy from Tai Lake plain, China. Agriculture, Ecosystems and
charcoal as an additive to plant tissue culture media. Zeitschrift für Pflanzen- Environment 139, 469e475.
physiologie 89, 141e147. Zheng, W., Guo, M., Chow, T., Bennett, D.N., Rajagopalan, N., 2010. Sorption prop-
Willmann, G., Fakoussa, R.M., 1997. Extracellular oxidative enzymes of coal- erties of greenwaste biochar for two triazine pesticides. Journal of Hazardous
attacking fungi. Fuel Processing Technology 52, 27e41. Materials 181, 121e126.
Wilson, J.B., 1988. A review of the evidence on the control of shoot:root ratio in Zimmerman, A., Gao, B., Ahn, M.Y., 2011. Positive and negative mineralization
relation to models. Annals of Botany 61, 433e449. priming effects among a variety of biochar-amended soils. Soil Biology and
Windstam, S., Nelson, E.B., 2008. Temporal release of fatty acids and sugars in the Biochemistry 43, 1169e1179.
spermosphere: impacts on Enterobacter cloacae induced biological control. Zimmerman, A., 2010. Abiotic and microbial oxidation of laboratory-produced black
Applied and Environmental Microbiology 74, 4292e4299. carbon (biochar). Environmental Science and Technology 44, 1295e1301.

View publication stats

You might also like