You are on page 1of 18

Environmental Pollution 227 (2017) 98e115

Contents lists available at ScienceDirect

Environmental Pollution
journal homepage: www.elsevier.com/locate/envpol

Effects and mechanisms of biochar-microbe interactions in soil


improvement and pollution remediation: A review*
Xiaomin Zhu a, b, Baoliang Chen a, b, *, Lizhong Zhu a, b, Baoshan Xing c
a
Department of Environmental Science, Zhejiang University, Hangzhou 310058, China
b
Zhejiang Provincial Key Laboratory of Organic Pollution Process and Control, Hangzhou 310058, China
c
Stockbridge School of Agriculture, University of Massachusetts, Amherst, MA 01003, United States

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

Article history: Biochars have attracted tremendous attention due to their effects on soil improvement; they enhance
Received 14 November 2016 carbon storage, soil fertility and quality, and contaminant (organic and heavy metal) immobilization and
Received in revised form transformation. These effects could be achieved by modifying soil microbial habitats and (or) directly
1 April 2017
influencing microbial metabolisms, which together induce changes in microbial activity and microbial
Accepted 13 April 2017
community structures. This review links microbial responses, including microbial activity, community
structures and soil enzyme activities, with changes in soil properties caused by biochars. In particular, we
summarized possible mechanisms that are involved in the effects that biochar-microbe interactions have
Keywords:
Biochar amendment
on soil carbon sequestration and pollution remediation. Special attention has been paid to biochar effects
Soil improvement on the formation and protection of soil aggregates, biochar adsorption of contaminants, biochar-
Microbial community mediated transformation of soil contaminants by microorganisms, and biochar-facilitated electron
Carbon sequestration transfer between microbial cells and contaminants and soil organic matter. Certain reactive organic
Contaminant mitigation compounds and heavy metals in biochar may induce toxicity to soil microorganisms. Adsorption and
Interaction mechanisms hydrolysis of signaling molecules by biochar interrupts microbial interspecific communications, poten-
tially altering soil microbial community structures. Further research is urged to verify the proposed
mechanisms involved in biochar-microbiota interactions for soil remediation and improvement.
© 2017 Elsevier Ltd. All rights reserved.

1. Introduction cycling and function (Biederman and Harpole, 2013; Lauber et al.,
2009; Rousk et al., 2009, 2010). There are some components in
For the purposes of soil remediation, biochar generally refers to biochar, including minerals, volatile organic compounds (VOCs),
a carbon-rich solid that is produced by the pyrolysis of biomass in and free radicals (Spokas et al., 2011), that can potentially influence
oxygen-limited conditions (Beesley et al., 2011; Chen and Chen, microbial activity, reshape the soil microbial community, and
2009). Biochars are applied to soil due to their potential benefits change the soil enzyme activity that catalyzes various key biogeo-
for carbon sequestration, soil fertility, and contaminant immobili- chemical processes including soil organic matter turnover and
zation (Cao et al., 2009; Chen et al., 2008a; Jeffery et al., 2015). The elemental cycles (e.g., N, P, and S) (Paz-Ferreiro et al., 2014). Due to
physiochemical properties of biochar are responsible for changes in the various positive effects on the soil properties and microbes,
soil character including changes in pH, nutrient maintenance, and biochars are considered effective agents for soil remediation.
water retention, which can induce heterogeneous responses in However, the variability among different types of biochar makes its
microbial species. This response can result in changes in microbial effects on soil remediation quite unpredictable, and the specific
community structure and can consequently alter soil element mechanisms of biochar-microbe interactions are still unclear.
The use of contaminant-degrading microbe inoculation
(mycoremediation) together with biochar can enhance the bio-
logical degradation of pollutants (e.g., PAHs) (Chen and Ding, 2012;
*
This paper has been recommended for acceptance by Dr. Hageman Kimberly Jill. Chen et al., 2012a; Garcia-Delgado et al., 2015), providing a prom-
* Corresponding author. Department of Environmental Science, Zhejiang Uni-
ising method for soil contaminant remediation. Such a process is
versity, Hangzhou 310058, China.
E-mail addresses: zhuxm@zju.edu.cn (X. Zhu), blchen@zju.edu.cn (B. Chen), zlz@ considered a combination of the immobilization of the pollutants
zju.edu.cn (L. Zhu), bx@umass.edu (B. Xing). by the biochar and the further degradation of these pollutants by

http://dx.doi.org/10.1016/j.envpol.2017.04.032
0269-7491/© 2017 Elsevier Ltd. All rights reserved.
X. Zhu et al. / Environmental Pollution 227 (2017) 98e115 99

microbes. The adsorption of organic contaminants, toxic heavy ergosterol extraction, quantitative real-time polymerase chain re-
metals, and hazardous anions (e.g., ClO 4 ) by biochar immobilizes action (q-PCR), fluorescence in situ hybridization (FISH), phos-
them and prevents their leaching into the groundwater (Cao et al., pholipid fatty acid quantitation (PLFA), molecular fingerprinting of
2009; Chen and Yuan, 2011; Chen et al., 2012d; Fang et al., 2014b; 16S rRNA gene fragments including denaturing gradient gel elec-
Jones et al., 2011; Yang et al., 2016b). Further transformation and trophoresis (DGGE) and terminal restriction fragment length
detoxification of environmental pollutants by microbes as cata- polymorphism (TRFLP), and high-throughput sequencing (also
lyzed by biochar have drawn increasing attention in recent studies known as next-generation sequencing, or NGS) of soil microbial
(Dong et al., 2014; Oh et al., 2013). Persistent free radicals (PFRs) genes (Chen et al., 2013; Hale et al., 2014; Kolton et al., 2011; Mackie
that are formed on biochar during thermal decomposition of the et al., 2015; Rousk et al., 2009). Changes in the relative abundances
feedstocks activate reactive oxygen species (ROS) (Fang et al., of Acidobacteria, Actinobacteria, Gemmatimonadetes, and Verru-
2015b; Kluepfel et al., 2014; Yu et al., 2015), and the electron comicrobia are frequently detected using high-throughput
transfer between biochar and microbial cells plays an important sequencing, under treatment with biochar (Mackie et al., 2015;
role in organic contaminant degradation and heavy metal trans- Nielsen et al., 2014). With higher resolution to the species level,
formation (Dong et al., 2014; Fang et al., 2015a; Yang et al., 2016a; the metagenomics sequencing of microbial genes is able to realize
Yu et al., 2015). function annotation reflected by the soil microbial community
Biochar can participate in soil processes such as organic matter structure changes (Ja €ckel et al., 2004). Such a process is essential to
decomposition as it takes part in the direct extracellular electron explain the effects of biochar on soil remediation (Chen et al., 2013;
transfer (DEET) between soil organic matter (or soil minerals) and Hale et al., 2014; Kolton et al., 2011; Mackie et al., 2015; Rousk et al.,
microbial cells, as well as in the direct interspecific electron transfer 2009). Since the mechanisms underlying biochar's effects on mi-
between microbial cells (DIET) (Chen et al., 2014; Fang et al., 2014a). crobes and related soil functions and processes are still not quite
The identification and quantification of the reactive components of clear, this review focuses on the synthesis of several possible
biochar particles that are responsible for the electron transfer be- mechanisms based on the published research.
tween the biochar and soil microbes are essential to investigate The influences of biochar on microbial activity are diverse and
biochar-involved elemental cycling. The electron transfer between seven possible mechanisms are demonstrated in the central circle
biochar particles and soil minerals, organic matter, pollutant mol- of Fig. 1 (from which points 1 to 3 can be classified into direct in-
ecules, and microbial cells, as well as in response of the microbial fluences, and points 4 to 7 indirect influences): (1) biochar provides
community to the reactive components of the biochar, is an shelter for soil microbes with pore structures and surfaces
emerging research field that seeks to further clarify the effects of (Quilliam et al., 2013a); (2) biochar supplies nutrients to soil mi-
biochar on soil biogeochemical processes. crobes for their growth with those nutrients and ions adsorbed on
Several studies in the past decade extensively discussed the biochar particles (Joseph et al., 2013); (3) biochar triggers potential
structure, physiochemical properties, and structure-function re- toxicity with VOCs and environmentally persistent free radicals
lationships of biochar with respect to the apparent effects of bio- (Fang et al., 2014a); (4) biochar modifies microbial habitats by
char on soil (Ahmad et al., 2014; Ameloot et al., 2013; Chen and improving soil properties that are essential for microbial growth
Yuan, 2011; Chen et al., 2008a, 2012b; Lehmann et al., 2011; (including aeration conditions, water content, and pH) (Quilliam
Warnock et al., 2007); however, an understanding of biochar- et al., 2013a); (5) biochar induces changes in enzyme activities
microbe interactions is a research gap that would link biochar that affect soil elemental cycles related to microbes (Lehmann et al.,
properties with many soil processes, e.g., carbon storage and 2011; Yang et al., 2016b); (6) biochar interrupts microbial intra- and
contaminant degradation. With regard to the newly found electron inter-specific communication between microbial cells via a com-
transfer and free radical activation functions of biochar, the bination of sorption and the hydrolysis of signaling molecules (Gao
mechanisms of biochar effects can be more clearly identified (Fang et al., 2016; Masiello et al., 2013); it should be noted that biochar
et al., 2015b; Kappler et al., 2014). Therefore, this review aims to may contain some molecules that can work as signals for microbial
seek answers for the following questions: (1) What are the most communication; and (7) biochar enhances the sorption and
essential physiochemical properties of biochar that influence the degradation of soil contaminants and reduces their bioavailability
microbial activity and community, and how do these properties and toxicity to microbes (Beesley et al., 2010; Qin et al., 2013;
influence the microbes? (2) Mechanistically, how does the inter- Stefaniuk and Oleszczuk, 2016). The proposed mechanisms
action of the soil microbes with different types of biochar affect soil involved in biochar-microbe interactions need further experi-
carbon sequestration and contaminant dissipation? Answers to mental verification, and a research emphasis should be placed on
these questions are essential to identifying the cause of the het- the linkage between biochar-microbe interaction mechanisms and
erogeneous effects of biochar in existing research and to develop their environmental effects.
the required understanding to accurately predict biochar effects,
both of which are key challenges in this research field. 2.1. Biochar provides shelter for microbes

2. Driving mechanisms of biochar-microbe interaction in soil One hypothesis of the benefits of biochar for microorganisms is
that biochars can be shelters for microbes due to their pore struc-
Biochar affects the soil microbial activity and biomass, changes tures. Biochars provide more habitable pore volume per unit vol-
the soil bacteria to fungi ratio and soil enzyme activity, and re- ume than soil does (Quilliam et al., 2013a). Microbial living cells can
shapes the microbial community structure (Ahmad et al., 2016; attach on biochar surfaces; in such cases, biochars with large spe-
George et al., 2012; Mackie et al., 2015; Nielsen et al., 2014; cific surface areas can provide habitats for microbes as well (Abit
Rutigliano et al., 2014). Note that biochar application may signifi- et al., 2012). However, the colonization of bacterial cells and
cantly alter the microbial community structure even when it does fungal hyphae has spatial heterogeneity between the external and
not change the overall microbial activity and biomass. To clearly internal pores of biochar (Quilliam et al., 2013a). Different microbial
interpret the microbial responses to biochar application in soils, colonization patterns on the surfaces and in the pores of biochar
gene copy numbers can serve as a more sensitive parameter than can be explained by three phenomena: 1) there is less nutrient
microbial biomass (Chen et al., 2013). Various techniques are used accessibility in biochar pores than in natural soil pores, 2) the
to test microbial activity and community structure, including biochar pores can be blocked with soil organic matter (e.g., humic
100 X. Zhu et al. / Environmental Pollution 227 (2017) 98e115

Fig. 1. Proposed mechanisms of biochar-microbe interactions and the environmental effects of biochar. The central circular area illustrates the interaction between biochar and
microbes, while the enclosing four boxes represent the effects of their interaction on carbon sequestration, soil processes (elemental cycling), contaminant degradation, and plant
growth. Interactions between the biochar and the microbes and its effects include: (1) biochar can act as a microbial shelter with its pore structure; (2) through sorption of nutrient
cations via functional groups, biochar can improve soil CEC and maintain nutrients for microbial growth; (3) free radicals and VOCs on biochar can be toxic to some soil microbes,
inhibit soilborne pathogens, and favor plant growth; (4) biochar can improve soil properties (e.g., pH, water content, and aeration conditions), and change the growth pattern of soil
microbes; (5) biochar can adsorb enzyme molecules, influence soil enzyme activities and elemental cycles; (6) biochar can adsorb and enhance the hydrolysis of signaling mol-
ecules, and consequently interrupt microbial communication and alter microbial community structure; (7) biochar can enhance the sorption (via biochar surface functional groups)
and degradation of soil contaminants (facilitated through electron transfer between biochar, microbes, and contaminants), which can reduce the toxicity of contaminants to soil
microbes. The interactions between biochar and soil microbes can alter the microbial community and their metabolic pathways (which can be revealed by metagenomics analysis of
microbial DNA sequencing), resulting in changed soil processes. There are interactions among different environmental effects as well.

Fig. 2. The interaction between microbial cells and biochar particles as observed with scanning electron microscopes (SEM). (A) Pine biochar inoculated with a co-culture of
Geobacter metallireducens (rods) and Methanosarcina barkeri (spheres) for a short time period (20 days, scale bar ¼ 1 mm) (Chen et al., 2014) and (B) Field-aged biochars buried in
agricultural soil for a long time period (3 years, scale bar ¼ 5 mm) (Quilliam et al., 2013a). (A) and (B) both show the attachment of microbial cells on biochar surfaces and pores. In
(C), the mineral phases have entered the pores of wheat-straw biochar (Joseph et al., 2013). The SEM images show that the mineral phases have undergone considerable reactions
that produce complex morphologies (Joseph et al., 2013).

acids), and 3) toxic substances, such as PAHs, may be present in Methanosarcina barkeri are able to attach themselves on biochar
biochar (especially in fresh biochar) (Kasozi et al., 2010; Quilliam surfaces during a very short time period (20 days, as shown in
et al., 2013a, 2013b). Microbial colonization on biochar surfaces Fig. 2A). The colonization of soil microbes (both fungal hyphae and
and pores is also dependent on the biochar aging process, which single bacterial cells) on surfaces and in the pores of biochar can be
can be considered a temporal heterogeneity (Quilliam et al., 2013a). improved by adjusting the aging periods of biochar (3 years, as
Bacterial cells from co-cultures of Geobacter metallireducens and shown in Fig. 2B) (Quilliam et al., 2013a).
X. Zhu et al. / Environmental Pollution 227 (2017) 98e115 101

2.2. Biochar supplies nutrients to soil microbes Zimmerman, 2013). In this case, biochar can act as a slow-release
fertilizer, bringing long-term benefits to soil fertility and microbi-
Biochar contains a range of nutrients (e.g., K, Mg, Na, N, and P) al growth. Notably, another reason that biochar supplies nutrients
(Chathurika et al., 2016; Rodriguez-Vila et al., 2016) and enriches for soil microbes might be that it regulates soil microbial functions
soil nutrients via sorption that is due to its large surface area, high that are essential for nutrient cycling. For example, biochar can
pore volume, and negative surface charge (Chen et al., 2012c). enhance the abundance of rhizobacteria that are able to transform
Cation exchange capacity (CEC) is a critical indicator of soil's ability organic S and P into bio-available forms, which further promotes
to retain cationic nutrients and supply nutrients to support mi- the growth of Lolium perenne (Fox et al., 2014) and are reasonably
crobial activity. The improved soil CEC that results from biochar suspected to promote the growth of other microbes that can only
application reflects a higher nutrient retention capability and a utilize inorganic S and P.
lower nutrient loss through leaching, which is beneficial for soil Biochar stores and supplies nutrients to soil microbes via the
microbial activity (Lehmann, 2007b), especially for microbes living sorption of nutrient cations and inorganic anions with its surface
in soils with low organic matter content (de Andrade et al., 2015; functional groups, especially oxygen-containing groups such as the
Laird et al., 2010; Mukherjee et al., 2011; Silber et al., 2010). The carboxylate group (Fig. 4) (Chen et al., 2015; El-Naggar et al., 2015;
nutrient conditions of biochar (indicated by ash content) are largely Jien and Wang, 2013; Liang et al., 2006; Mukherjee et al., 2011;
determined by its feedstock and pyrolysis temperatures. Biochars Yuan et al., 2016). Some studies showed that biochar maximized
produced from herbaceous material (crop residues) and manure CEC at low and moderate pyrolysis temperatures, while to the
generally have higher ash content than wood biochar and are contrary, some other studies showed that biochar CEC increased
consequently able to supply more nutrients (Fig. 3A) (Akhter et al., with the pyrolysis temperature (Lehmann, 2007a; Mukherjee et al.,
2015). A clear rising curve between ash content with the pyrolysis 2011; Yuan et al., 2011). The feedstock types and pyrolysis program
temperatures has been found for crop residue biochar and manure parameters, including temperatures, heating rate, and holding
biochar (Xu and Chen, 2013). The nutrients from biochar can be time, primarily determine the biochar functional groups and
released with different rates into soils (Mukherjee and consequently the capability of biochar to improve the soil CEC

Fig. 3. Biochar properties as function of pyrolysis temperature and feedstock types. (A) to (D) show the changes in ash content, specific surface area (SSA), pH, and volatile matter
content (VM), respectively, with pyrolysis temperature. Different colors and shapes indicate different feedstock types, and linear (or nonlinear) regressions between biochar
properties with pyrolysis temperatures are analyzed; the regression and its 95% confidence and 95% prediction bands are illustrated (feedstock types are not distinguished, analyzed
with the software SigmaPlot). Positive relationships between pH and ash content with pyrolysis temperature are found, while negative relationships between VM content with
pyrolysis temperature is found. An increased ash content with temperature is more obvious for manure biochar and crop residue biochar than for wood biochar (A), while an
increased SSA with pyrolysis temperature is only found for wood biochar (B). “Other biochars” include distilled grain biochar and sugar cane bagasse biochar. Data are from Table S1.
102 X. Zhu et al. / Environmental Pollution 227 (2017) 98e115

Fig. 4. Schematic diagram showing the roles of biochar functional groups (AFG ¼ acidic functional groups, SOM ¼ soil organic matter): (1) The AFG are responsible for the liming
effect of biochar which modifies the soil microbial habitat; (2) the electrostatic attraction between the carboxyl groups of biochar with the nutrient cations effectively retains
nutrients to ensure a nutrient supply to soil microbes and (3) to immobilize heavy metals, thus reducing heavy metal toxicity to microbial cells; (4) electrostatic attraction, as well as
polar and non-polar organic attraction, of humic acid molecules can result in the adsorption of soil organic matter that is beneficial for carbon sequestration (further discussed in
later chapter); (5) hydrogen bonding between eOH groups on biochar with oxygenated anions can adsorb inorganic anions to supply nutrients or reduce anion contaminant
toxicity; (6) electron transfer to form free radicals on the biochar surface can facilitate organic contaminant degradation and heavy metal transformation (to be detailed discussed in
later chapter) and can reduce contaminant toxicity to microbes.

(Table S1) (Chen et al., 2008a; Jeong et al., 2016; Mukherjee and increases the recalcitrant SOC pool, enhancing soil carbon seques-
Zimmerman, 2013; Yuan et al., 2011). The biochar CEC was pH tration (Yanardag et al., 2015). Bacteria and fungi have their own
dependent, increasing from low to neutral environmental pH preferences for different carbon sources and have different toler-
values (Lehmann, 2007a), which indicated possible interactive ef- ances to changes in environmental factors, such as pH and water
fects between the changes in the pH and the CEC in soils with condition (Rousk et al., 2009; Zhang et al., 2015). Compared with
biochar. Additionally, interaction between biochars and soil min- bacteria, fungi are able to colonize soil macroaggregations
erals can be responsible for the long-term maintenance of the (>200 mm) that contain a higher SOC, ratio of C/N, and total N
minerals during the aging of biochar (Novak et al., 2009). For (Zhang et al., 2015). This ability may be due to the advantages of the
example, a complex reaction between biochar (wheat straw) and living strategy of fungi; specifically, the translocation of nutrients
soil clay as well as extra added chemical fertilizers (including and water within a continuous hyphal network enables fungi to
CO(NH2)2, KCl and monoammonium phosphate) can result in some colonize poor carbon sources with a high C/N, such as biochar
unusual morphologies (not identified) on biochar surfaces, as (Ascough et al., 2010). Therefore, biochar application that promotes
indicated in Fig. 2C (Joseph et al., 2013). From the above results, the the formation of macroaggregates should favor fungi growth rather
various aspects of biochar preparation including feedstock types, than bacteria under the same environmental condition. Both
pyrolysis temperatures, the aging period, and the relationships atomic H/C and O/C ratios commonly decrease with increased py-
between various biochar properties (e.g., pH and CEC) should be rolysis temperature, indicating a more intensive aromatic and
fully considered during the field application of biochar to improve nonpolar structure of the biochar (Fig. S1) (Hale et al., 2015; Xiao
the nutrient supply to soil microbes. et al., 2016). Biochar C/N ratios are mainly dependent on feed-
The aromaticity of biochar is responsible for its recalcitrant stock types; for example, the C/N of wood biochar is normally
nature against microbial decomposition (reflected by the atomic H/ higher than that of manure biochar (Cantrell et al., 2012; Chen et al.,
C and O/C ratios), and some fractions of biochar can also function as 2008a). In addition, the increased pyrolysis temperature can cause
a carbon source for soil microbes (the quality of biochar as a carbon elevated C/N in some feedstocks due to the condensed aromatic
source is indicated by its C/N ratio, Table S1, Fig. S1) (Demisie et al., carbon structure (Table S1, Fig. S1B) (Cantrell et al., 2012; Spokas
2014). Biochars are typically low in available carbon for microbial et al., 2011). Therefore, the pyrolysis temperature and feedstock
utilization because they contain higher C/N ratios than their feed- type are useful when considering the stability of biochar C against
stocks and are thus difficult to degrade by microorganisms because microbial decomposition in practical application.
of the lack in N supply. As a consequence, biochar amendment
X. Zhu et al. / Environmental Pollution 227 (2017) 98e115 103

2.3. Potential toxicity of biochar to microbial cells particulate matter; these radicals can reduce oxygen to form su-
peroxide, which then forms H2O2 that can initiate Fenton reactions
Certain compounds in biochar are known as microbial in- in the presence of transition metal ions such as Fe2þ and Cu2þ
hibitors, including benzene (the dominant product of pyrolysis (which are ubiquitous in biological systems). These Fenton re-
during glowing combustion of charcoal), methoxyphenols and actions generate OH that causes DNA strand breaks, which lead to
phenols (the products of pyrolysis of hemicelluloses and lignin), DNA damage (Dellinger et al., 2001). On the other hand, free radi-
carboxylic acids, ketones, furans (that are generally presented as cals from biochar can play an important role in the degradation of
the sorbed VOCs on biochar), and PAHs (Ghidotti et al., 2017; Lyu organic pollutants with the ROS induced by the biochar, especially
et al., 2016; Spokas et al., 2011). Such compounds can be identi- the strongest species: the hydroxyl radical (OH) and the sulfate
fied from water or organic solvent extractions of biochar. Organic radical anion (SO4 ) (Fang et al., 2015a, 2015b), which will be dis-
solvent extracts from biochar contain organic compounds from cussed later. Both the potential toxicity of biochar free radicals on
many classes, including n-alkanoic acids, hydroxyl and acetoxy soil microorganisms and the positive effects of biochar free radicals
acids, benzoic acids, diols, triols, and phenols, while water extracts on pollutant degradation are two sides of the coin to be considered
from biochar contain dicarboxylic acids, aromatic organic acids and in the evaluation of the environmental effects of biochar.
polyol along with hydroxy acids, n-alkanoic acids, and benzoic
acids (Table 1) (Graber et al., 2010). Biochar VOCs vary with feed- 2.4. Biochar modifies microbial habitats
stock type, elemental composition, pyrolysis temperature, and
heating procedures and conditions (Spokas et al., 2011). Higher Biochar can modify microbial habitats by improving the soil's
concentrations and toxicities of PAHs and polychlorinated dioxins physical properties. Biochar porosity can decrease soil bulk density,
and furans (PCDD/DF) are found in biochar generated at moderate improve soil aeration condition (Abel et al., 2013), and control the
temperatures (300 and 400  C) than at high temperatures transport of soil microbes in biochar amended soil (Abit et al.,
(>400  C) (Lyu et al., 2016). The diversity of composition for the 2012). Biochar can increase the available water content that in-
VOCs species sorbed on biochar can be the main contributing factor fluences nutrient accessibility to microbial cells (Abel et al., 2013).
to the various responses of soil microbial activity to biochar. In addition, the biochar can increase water content at the perma-
Although volatile organic compounds (VOCs) in fresh biochar can nent wilting point, which indicates the capability of biochar, via its
support the survival of some microbes (such as Bacillus mucilagi- high porosity, to store water unavailable to plants; keeping water in
nosus) as a carbon source (Sun et al., 2015), they can induce po- this manner is beneficial, especially in sandy and degraded soils
tential toxicity to microbes (not species-specific) when present in (Abel et al., 2013). Moreover, the improved water retention capacity
high concentrations (especially for some low molecular weight means that there is a greater capability of the soil to hold water
oxygenated VOCs, including acids, alcohols, and carbonyls) (Ennis against dry-wet cycles in the natural environment, which can favor
et al., 2012; Spokas et al., 2011). As microbial inhibitors, VOCs can the maintenance of a stable microbial activity (Liang et al., 2014).
also induce direct toxicity to microbial soil pathogens, thus The pyrolysis parameters (mainly temperature, heating rate and
benefitting plant growth (Graber et al., 2010). However, the inhi- time) and feedstock compositions (e.g., lignin and lipid concen-
bition may not be very specific to pathogens, and the alteration of trations) of the preparation of biochar control its porosity, carbon
the microbial community needs further investigation to evaluate stability, and the surface adsorption of nutrients (Abit et al., 2012;
the environmental benefits and risks of biochar application. Cantrell et al., 2012; Chen et al., 2008a). Although an increasing
Persistent free radicals (PFRs) that are generated and stabilized trend of specific surface area with pyrolysis temperatures is found
during the pyrolysis of biochar, including semiquinones, phenoxyls, for wood biochar, this relationship is not found for biochars
cyclopentadienyls, and phenols, can also induce toxicity to mi- generated from other feedstock types such as manure and crop
crobes (Truong et al., 2010). The formation of free radicals on bio- residue, indicating the extraordinarily important role of feedstock
char includes replacing surface functional groups on metal oxides type on the formation of large SSAs (Table S1, Fig. 3B) (Chen et al.,
with aromatics while metal oxides are being reduced (Fig. 5I) 2008a; Wang et al., 2015). The role of biochar in improving soil
(Balakrishna et al., 2009; Doong et al., 2014; Fang et al., 2014a) and properties and modifying microbial habitats can be dependent on
breaking chemical bonds in macromolecules (Fig. 5II) (Dellinger the feedstock types and pyrolysis procedures used in making
et al., 2007; Liao et al., 2014). The PFRs are abundant (approxi- biochar.
mately 1017 spins g1 pine needle biochar) and depends on the Biochar can be an effective liming agent to neutralize soil pH
pyrolysis temperature and time (Fang et al., 2015a). The adverse (Yuan et al., 2011). An increase of the soil pH by 0.2e0.3 units after
impacts initiated by biochar PFRs were previously misidentified biochar application can be the main factor affecting the soil mi-
because of the lack of proper methods to separate those effects crobial community, in contrast to other chemical variables in soil
caused by free radicals from those caused by co-existing molecular such as the total C content, total N content, electrical conductivity
pollutants; they have yet to be taken into account in current (EC), and NO þ
3 and NH4 concentrations (Nielsen et al., 2014). An
research (Liao et al., 2014; Truong et al., 2010). increase in the soil pH and a decrease in the toxicity of exchange-
Free radicals can induce oxidative stress in living microbial cells, able Al in acidic soils with biochar amendment (Qian et al., 2013)
reduce cellular glutathione (GSH), glutathione peroxidase (GPx), increase the bacteria abundance in the pH range from 4 to 7; the
and superoxide dismutase (SOD) levels, and decrease cell mem- abundances of bacteria are positively correlated with soil pH (Leal
brane integrity with the generation of reactive oxygen species et al., 2015; Rousk et al., 2010). Fungi and bacteria have different
(ROS) such as hydroxyl radicals (OH), the superoxide radical anion sensitivities to soil pH (Rousk et al., 2009), and bacteria are
(O2 ), and hydrogen peroxide (H2O2) (Balakrishna et al., 2009; generally more tolerant to a narrower range of pH and are more
Dellinger et al., 2007; Liao et al., 2014). Free radicals interfere sensitive to soil pH changes than fungi are (Rousk et al., 2010). As a
with cytochrome P450s (e.g., CYP1A2, a member of the ubiquitous result, there are likely to be different responses from bacteria and
superfamily of enzymes that catalyze the mixed-function oxygen- fungi to biochar-induced changes in soil pH, and the microbial
ation of both endogenous and foreign compounds) and competi- community of fungi and bacteria may react in different ways to
tively inhibit the metabolism of exogenous organic substrates biochar-induced changes in soil pH, which can result in an alter-
(Reed et al., 2015). Semiquinone radicals (QH) are one type of free ation of the overall microbial community structure. A pyrose-
radicals that are found in combustion-generated ultrafine quencing analysis of the soil bacterial community showed
104 X. Zhu et al. / Environmental Pollution 227 (2017) 98e115

Table 1
Organic compounds frequently detected from various biochar types as VOCs, organic solvent extracts, or water extracts of biochar.
X. Zhu et al. / Environmental Pollution 227 (2017) 98e115 105

significant correlation between the soil pH and the soil bacterial consequence (Zhang et al., 2017). Pyrolysis temperature domi-
community composition. The relative abundances of Acidobacteria, nantly determined biochar pH values, which usually rise with
Actinobacteria, and Bacteroidetes were largely driven by the range of increased pyrolysis temperature for various feedstocks (Fig. 3C).
soil pH values (3.5e9.0) at the continental scale from North to This trend is mainly attributed to the decreased number of acidic
South America (Lauber et al., 2009). Increasing the soil pH with functional groups (AFG) and decreased volatile matter (VM) con-
biochar application altered the abundance, diversity, and compo- tent (Fig. 3D) (Mukherjee et al., 2011), yet there is an increased
sition of nitrifying bacteria in soil and changed soil nitrification as a alkalinity (due to carbonates and alkaline ash content) of biochar
106 X. Zhu et al. / Environmental Pollution 227 (2017) 98e115

Fig. 5. Schematic of the mechanisms of PFR formation and free radical generation on biochar, including: I) the interaction between organic compounds containing oxygenous
functional groups and metal oxidation, and II) the breaking of chemical bonds in macromolecules during heating and cooling. The figure is modified from the reference (Fang et al.,
2014a; Liao et al., 2014).

with increased pyrolysis temperature (Smebye et al., 2016; Yuan The sorption (binding) of enzymes on biochar and soil organic
et al., 2011). Therefore, high temperature biochars are normally matter can change the kinetic properties of enzyme activity
more effective at improving the soil pH (Cantrell et al., 2012; Cao (Lammirato et al., 2011; Nannipieri et al., 2012) and therefore is the
and Harris, 2010; Mukherjee et al., 2014; Yuan et al., 2011). most important mechanism regulating the soil enzyme activity
(Zimmerman and Ahn, 2010). The adsorption efficiency of the
enzyme and substrates depends on the biochar structure: adsorp-
2.5. Biochar alters soil enzyme activity tion of enzyme molecules on biochar surfaces is considered to be
driven by non-coulombic forces between the uncharged regions of
Most of the elemental turnover in soil, which determines the the protein and the uncharged regions of the biochar surface, and
nutrient bioavailability and includes turnover of C, N, P, and S, is the adsorption of small molecular polar substrates (e.g., disaccha-
catalyzed by enzymes. Soil enzymatic activities respond faster than ride) on charred fractions (especially activated carbon) is stabilized
other soil variables to soil management and disturbance introduced via hydrogen bonding to polar surface groups (e.g., COOH, SO4H,
into soil and are therefore indicators of biological changes and soil PO4H) on the sorbents (Lammirato et al., 2011). Changes in surface
quality (Bandick and Dick, 1999). Enzyme activities respond to functional groups in aged biochar alter the adsorption capacity of
biochar application in various ways, mainly depending on the types enzyme and substrates, thus affecting enzyme activity (Gibson
of enzymes and biochar, the biochar application rate, and soil et al., 2016). For example, the oxidation of aromatic carbon and
properties (Table S2). The decreased microbial abundance and ac- the introduction of aliphatic C-H groups on biochar during the
tivity of soil enzymes in relation to organic matter decomposition abiotic aging process improved laccase and peroxidase activities
can enhance C sequestration (Luo and Gu, 2016). Possible mecha- and enhanced fungal respiration (Gibson et al., 2016). Biochar can
nisms involved in the influence of biochar on enzyme activity reduce the activation energy (Ea, related to the temperature
include but may not be limited to: (1) biochar adsorbs extracellular sensitivity of an enzyme) of an enzyme-catalyzed reaction and
enzyme molecules and/or substrates on the surface or blocks the down-regulate the enzymatic sensitivity to temperature changes
reaction site of enzymes (Bailey et al., 2011) in a way regulates their (in terms of Q10), resulting in enhanced activities of b-glucosidase
apparent affinity with substrates (Paz-Ferreiro et al., 2015); (2) and arysulfatase (Paz-Ferreiro et al., 2015). A reduced Ea under
biochar influences enzyme activity with changes in soil physi- biochar application indicates higher substrate affinity (in term of
ochemical properties (especially pH) (Zimmerman and Ahn, 2010); apparent Km) and less temperature sensitivity of the enzyme,
and (3) biochar releases some small molecules that are speculated possibly due to sorption of substrates and enzymes on biochar (Paz-
to act as allosteric regulators or inhibitors of specific enzymes (such Ferreiro et al., 2012). On the other hand, soil enzymes respond
as a possible up-regulation of beNeacetylglucosaminidase activity quickly to soil management (e.g., organic matter amendment)
with ethylene) (Bailey et al., 2011).
X. Zhu et al. / Environmental Pollution 227 (2017) 98e115 107

(Bandick and Dick, 1999), and thus changes in soil properties with that can act as microbial inhibitors (Spokas et al., 2011) may be the
biochar application can reasonably influence soil enzyme activities. reason for the uncertainty in biochar-triggered pathogen resis-
For the third mechanism, inhibitors released from biochar can tance. To better reveal the mechanisms, the identification and
interfere with enzyme-catalyzed reactions as well: for example, quantification of reactive compounds from biochar, which are able
during pyrolysis, plant biochars can release a number of benzofu- to interfere with microbial communication, is an important
rans, polycyclic aromatic hydrocarbons, and heterocyclic com- research field.
pounds, which are inhibitory compounds to soil enzymes
(Lammirato et al., 2011). 2.7. Biochar reduces the toxicity of contaminants to soil microbes

2.6. Biochar affects intra- and interspecific communication of Biochar as a soil ameliorant can decrease the toxicity of soil
organisms contaminants to soil microbes (Koltowski et al., 2017). Willow
biochar (pyrolyzed at 700  C) can reduce the microbe mortality
Biochar modifies microbial cell-to-cell communication via the while increasing the reproduction of Folsomia candida in soils
sorption of signaling molecules (e.g., an N-acyl-homoserine contaminated with heavy metals and organic pollutants (PAHs) and
lactone, AHL) (Masiello et al., 2013) and, perhaps more importantly, can decrease the leachate toxicity to the bacterium Vibrio fischeri
by enhancing the hydrolysis of AHL (Gao et al., 2016). The inter- (Koltowski et al., 2017). The immobilization of soil contaminants
cellular signaling molecule N-3-oxo-dodecanoyl-L-homoserine (including heavy metals such as Al, Cd, Co, Cr, Mn, and Ni, and
lactone is an AHL that is used by many gram-negative soil bacteria organic pollutants such as PAHs) on biochar and the consequent
(e.g., nitrogen-fixing plant symbionts and pathogens that cause soft reduction of their bioavailability can be the main reason for the
rot in plants) to regulate gene expression and intraspecific alleviated toxicity of soil contaminants to microbes and the
communication (Masiello et al., 2013). When sorption is the main elevated microbial biomass (Qian et al., 2013, 2015; Seneviratne
mechanism, this effect is dependent on pyrolysis temperature, et al., 2017; Zielinska and Oleszczuk, 2016). Rice straw biochar
which influences the biochar adsorption capacity (Masiello et al., application (5% application rate) can result in an increase of the
2013). High-temperature biochar (700  C), with higher specific organic-bound fraction of heavy metals (Cd, Cu, Pb, and Zn) by up to
surface area (SSA), adsorbed the signaling molecule (AHL) and 68% (Lu et al., 2017). The reduced heavy metal stress on N-fixing
interrupted the inter-cell transfer of the signal to a larger extent bacteria (Bradyrhizobium japonicum) could further benefit the N
than moderate temperature biochar (300  C), with a much lower supply for plant growth (Seneviratne et al., 2017). In addition to the
SSA, did (Masiello et al., 2013). By adjusting the soil pH, biochar is effect of reducing the toxicity of contaminants to soil microbes, thus
able to inhibit or promote the exchange of signaling compounds enhancing microbial activity, interactions between the biochar and
that regulate specific soil microbial activity (Warnock et al., 2007). soil microbes have more profound effects on the environmental
Additionally, the increase in soil pH induced by biochar can drive fate of soil contaminants, including their immobilization and
the hydrolysis of the signaling AHLs, thereby reducing the number degradation, which are to be discussed in a later chapter of this
of bioavailable AHLs and consequently inactivating the bacterial review.
cell-to-cell communication (Gao et al., 2016). While some signaling
molecules for bacterial communication (such as AHL) are pH- 3. Soil carbon storage under biochar-microbe interaction
sensitive, some fungal signals (such as farnesol, a fungal auto-
inducer) are less sensitive to pH, possibly leading to shifts in the Biochar can enhance soil carbon sequestration by various
fungi to bacteria ratio (and a change in the soil microbial commu- mechanisms. In addition to the stable carbon structure and its own
nity structure) with biochar application (Gao et al., 2016). recalcitrant nature (Guo and Chen, 2014; Xiao et al., 2014), biochar
Biochar can modify microbe-to-plant communication in the can affect carbon cycle dynamics that involve microbial participa-
rhizosphere (Harel et al., 2012), affect the competition between tion, leading to either positive or negative priming effects on native
beneficial soil microbes and soilborne pathogens (Akhter et al., organic matter, which represent enhanced or inhibited degradation
2015), and trigger systemic plant defenses to soilborne pathogens of native soil organic carbon (SOC), respectively, when extra carbon
(Akhter et al., 2015; Elad et al., 2011). Application of biochar suc- sources are added (Bamminger et al., 2014; Cely et al., 2014;
cessfully induced resistance against two foliar fungal pathogens Kuzyakov et al., 2000; Zimmerman et al., 2011). The interactions
(Botrytis cinerea and Leveillula taurica) for both pepper and tomato between biochar, soil microbes, and native SOC account for the
plants (Elad et al., 2010). This action is regulated by either the stability of biochar C and native SOC, in addition to having an
enhanced interaction between plant roots and the plant-growth- overall effect on carbon sequestration (Fig. 6). The balance between
promoting rhizobacteria (PGPR) or fungi (PGPF) (Harel et al., the decomposition and sequestration of biochar C results in an
2012) or by the direct toxic effects of certain biochar components overall contribution of biochar to the recalcitrant C pool. In addi-
on microbial pathogens (Graber et al., 2010). Both wood biochar tion, the effects of biochar on the microbial decomposition and
and green waste biochar, when amended to the potting medium, physical protection of native SOC determine the fate of native SOC.
stimulated a range of general defense pathways of strawberry The stability of biochar C, effects of biochar on the stability of native
plants and as a result suppressed anthracnose disease caused by the SOC, and the possible mechanisms accounting for C sequestration
fungal pathogens Botrytis cinerea, Colletotrichum acutatum, and are the focuses of this chapter.
Podosphaera aphanis (Harel et al., 2012). It was found that higher
application rates (3%) of biochar induced higher defense ability 3.1. Carbon stability of biochar
against the pathogen-caused plant disease than low application
rates (1%) of biochar did (Harel et al., 2012). Moreover, there are also As biochars are recalcitrant for microbial degradation, they are
various influences of different biochar types (wood and green waste considered stable carbon pools and have carbon sequestration po-
biochar) on the competition between arbuscular mycorrhizal fungi tential. However, biochars produced from different conditions may
(AMF) and the soilborne pathogens (e.g., Fusarium oxysporum f. sp. have different stability, which can be reflected by the amount of
lycopersici) in the mycorrhization on tomato roots (Akhter et al., aromatic structure in the biochar, as indicated by the H/C and O/C
2015). The variability in the amount and types of biochar VOCs atomic ratios, and the amount of volatile matter which can be
108 X. Zhu et al. / Environmental Pollution 227 (2017) 98e115

3.2. Effects on native organic carbon stability

There is abundant evidence that biochar application can affect


the stability of native soil organic carbon (SOC) in terms of a
“priming effect” (Ameloot et al., 2014; Jien et al., 2015; Zimmerman
et al., 2011). This effect can be caused by microbial community
changes that are induced by biochar, which in turn increase or
decrease the decomposition of native SOC (thus determining the
stability of the native SOC). The effect may also be caused by the
physical protection of native SOC on biochar (Fig. 6) (Ameloot et al.,
2013). An alteration of the soil microbial community to a higher
fungi-to-bacteria ratio can suppress the degradation of native SOC
and cause a large negative priming effect up to 68% (Bamminger
et al., 2014). The priming effects of biochars on native SOC are
dependent on the feedstock characteristics and pyrolysis condi-
tions, which determine the difficulty of utilizing biochar organic
carbon fractions as a carbon source by soil microbes.
The dominant properties of biochar that can influence the
priming effects on native SOC include the carbon content, carbon
Fig. 6. Schematics showing the interactions between biochar, soil microbes and native aromaticity, volatile matter content, the amount of fixed and easily
SOC, in particular their effects on soil carbon stability and the decomposition of biochar oxidized carbon components, and various surface properties (Cely
C and native SOC. The volatile matter (VM) on biochar can be used by microbes as a C et al., 2014). Due to the differences in carbon stability, woodchip
source and contributes to microbial respiration. Some soil microbes can utilize aro-
biochar had a negative priming effect, while biochars generated
matic C structures as well. Native SOC provides soil microbes with a major C source,
and the resulting microbial decomposition of native SOC contributes to microbial from wheat husk and sewage sludge caused a positive priming
respiration. On the other hand, biochar and soil microbes can protect native SOC from effect in short-term (45 days) incubations (Cely et al., 2014). In
decomposition by forming soil aggregates with certain functional groups on the bio- contrast to the role of the aromatic carbon structures of biochar in
char and the microbial secretion of glomalin, respectively. The resulting soil aggregates the priming effect, the volatile matter content of biochar (i.e., labile
of organic matter, together with the aromatic C structure of biochar as a recalcitrant C
pool in soil, can contribute to the C sequestration.
carbon fraction) correlated positively with the initial biological
oxygen demand (BOD), implying that positive priming effects can
be due to the microbial utilization of the labile carbon fraction
within the applied substrate, which can enhance the microbial
utilized by soil microbes as a carbon source (Fig. 6). The reaction of growth that controls the positive priming effect on native SOC
soil microbes to biochar and hence the stability of biochar are (Ronsse et al., 2013). On the other hand, biochar application can
driven by the feedstock types and pyrolysis temperatures of the enhance the effectiveness of microbial energy utilization by
biochars (Ameloot et al., 2015; Steinbeiss et al., 2009). Maize stover simultaneously decreasing the metabolic quotient (qCO2) and
biochar had higher C, N, and P content as well as higher carbon increasing the soil microbial biomass (or in some cases, by reducing
stability than rice and wheat straw biochars did; as indicated by its soil respiration with unchanged microbial biomass) (Bamminger
aromatic C¼C structure, it also had less carbon mineralization et al., 2014; Zheng et al., 2016), which accounts for one mecha-
during one year (Purakayastha et al., 2015). Oilseeds (castor bean nism of the negative priming effect of biochar on native SOC. The
seeds) contain high amounts of aliphatic components, and the decreased qCO2 of soil microbes can be achieved by modifying the
biochar generated from them has some quickly decomposing thermodynamic parameters of soil enzyme activities with biochar
compounds that result in greater mineralization than found in applications (Paz-Ferreiro et al., 2015).
wood biochar (Rittl et al., 2015). On the other hand, pyrolysis The oxidation of biochar surfaces during the aging process in-
temperatures can result in variable carbon structures of biochar, troduces oxygenated functional groups (e.g., carboxylic and ester
generally leading to enhanced carbon stability with increased py- groups) on biochar surfaces (Novak et al., 2009; Qian and Chen,
rolysis temperature (Xiao et al., 2014). 2014); therefore, it can enhance the stability of native SOC by
Biochar stability affects the microbial decomposition of biochar, physically protecting soil aggregates (Fig. 6) (Heitkoetter and
which is defined as biological aging (Czimczik and Masiello, 2007; Marschner, 2015). The introduction of the oxygenated functional
Zimmerman, 2010). It is suggested that labile and semi-labile groups on biochar surfaces can lead to enhanced biochar and soil
fractions make up the non-stable biochar fractions, and the char- CECs (Hale et al., 2011; Lin et al., 2012), and higher sorption of soil
acteristics of stable and non-stable fractions of biochars determine DOM on biochar surfaces can provide acidic functional groups and
their resistance to chemical and biological degradation (Masek further increase the surface charge of the biochar-SOM aggregates
et al., 2013). Microbes involved in the transformation of aromatic (Fig. 6) (Heitkoetter and Marschner, 2015). Moreover, the
compounds may be the first colonizers and decomposers of fresh oxidation-introduced carboxylic functional groups on aged biochar
pyrogenic carbon, including biochar (Zimmermann et al., 2012); as surfaces serve as additional binding sites, which can intimately
such, these microbes influence the carbon stability. For example, attach mineral phases composed of Al, Si, and Fe and form large
biochar application has caused increases of the relative abundance aggregates (Lin et al., 2012; Qian and Chen, 2014). As a conse-
of chitin and cellulose degraders (Chitinophaga and Cellvibrio, quence, biochar is considerably involved in the preservation of the
respectively) and aromatic compound degraders (Hydrogenophaga soil humus, including humic acids (HAs), fulvic acids (FAs) and
and Dechloromonas) on the genus level, likely being induced by the humins (HMs) in soil, which can be ascribed to the presence of
enrichment of aromatic compounds such as phenol, methylphenol, abundant oxygenated functional groups that favor soil aggregation
and dihydroxybenzenes in biochar (Kolton et al., 2011). Enrichment formation via sorption (Hua et al., 2015).
of the chitin-, cellulose-, and aromatic compound-decomposing To evaluate the effects of biochar application on the soil carbon
microbial genera indicated their important role in the biochar sequestration and the deposition of native SOC, it is crucial to
biological aging process. distinguish between the microbial decomposition of organic
X. Zhu et al. / Environmental Pollution 227 (2017) 98e115 109

carbon from biochar and that from native SOC. For this purpose, a of soil arbuscular mycorrhizal fungi that are able to secrete glo-
two-component model can be used to indicate the carbon stability malin, the protein that promotes soil microaggregation due to its
of biochar carbon and native SOC in a system receiving biochar impact on hydration (King, 2011); and (5) the influences on the soil
application (Galvez et al., 2012; Murray et al., 2015; Zimmerman microbial community and the modification of soil enzyme activities
et al., 2011), as shown in equation (1): that control soil organic carbon decomposition (Paz-Ferreiro et al.,
2015). Among these factors, the protection of soil aggregates and
-k1t -k2t
Cmineralized (%) ¼ f  (1e ) þ (100f)  (1e ) Eq. (1) modification of enzyme activities are related to microbial activity
and are to be mainly discussed here.
where Cmineralized is the carbon fraction that is cumulatively Biochar can act as a binding agent of soil organic matter during
mineralized, f is the carbon fraction (%) of the active or fast turnover the formation of aggregates and can thus promote the macro-
carbon pool (native SOC), k1 and k2 represent the mineralization aggregation (George et al., 2012) that enhances the stability of soil
rate constants of the active and resistant carbon pools (biochar aggregates (Sun and Lu, 2014). The number of stable macroaggre-
carbon), respectively, and t is the incubation time in days. The mean gates (>250 mm) can be increased by approximately 25% in the case
retention time (MRT) of the respective carbon pools can be calcu- of a sandy soil after 80 days of incubation, which is helpful for soil
lated as the inverse of the mineralization rate constants k1 and k2, pore water retention, soil carbon sequestration, and microbial
i.e., 1/k1 and 1/k2 (Murray et al., 2015). After a given application of protection (Awad et al., 2013). The aggregation of large-size soil
biochar, the shifts of the parameters within this model can indicate particles is partly due to the adsorption of soil organic compounds,
changes in carbon stability and the effectiveness of carbon such as humic acids, on biochar surfaces (Kasozi et al., 2010).
sequestration. This model indicates that less than 3% of the applied Changed patterns of soil aggregation will alter microbial commu-
biochar was lost through CO2 evolution, with an MRT of nity structures in soil, since gram-positive and gram-negative
600 years at an annual mean temperature of 26  C (or, 3264 years at bacteria tend to possess different strategies in occupying soil par-
10  C) (Major et al., 2010). The model also indicates that biochar can ticle fractions. There is a greater abundance of gram-positive bac-
reduce the mineralization rate constant of native SOC and increase teria in silt and clay fractions, which are older and have more
its MRT, especially in acidic soil (Galvez et al., 2012). microbially processed soil organic matter, while there are more
gram-negative bacteria in fractions with larger (>200 mm) aggre-
gates, as these bacteria are fond of fresh plants as C sources (Zhang
3.3. Mechanisms of biochar-enhanced soil carbon sequestration et al., 2015). Similarly, fungal abundance was found to be reduced
with the decreasing size of soil aggregation due to the C substrates
The factors involved in the maintenance of soil carbon stability being less available in soil fractions with aggregates smaller than
with biochar application (Fig. 7) include: (1) the recalcitrant nature 200 mm (Zhang et al., 2015).
of biochars, due to their aromatic carbon structure and the crystal The effects of biochar amendment on soil microaggregation are
silicon structure in silica-carbon complexes formed during pyrol- not only related to the direct adsorption of soil organic matter on
ysis (Guo and Chen, 2014; Xiao et al., 2014); (2) the interaction biochar surfaces but also to the alteration of the microbial com-
between surface oxygenated functional groups such as carboxyl munity (especially fungi) that produces proteins such as glomalin,
with soil minerals are favorable to maintaining the carbon stability hydrophobins, and chaplins that positively influence micro-
of biochar during the aging process (Chen et al., 2015; Li et al., aggregate formation and stability (King, 2011; Rillig et al., 2005).
2014); (3) the adsorption of soil organic matter that is beneficial Glomalin, produced by a phylogenetically narrow range of arbus-
for soil aggregate formation (George et al., 2012); (4) the availability cular mycorrhizal fungi within the order Glomales, has the function
to increase the hydrophobicity and stability of microaggregates
(Rillig et al., 2002). Hydrophobins, occurring widely in numerous
basidiomycetes and ascomycetes, take vital parts in mycelia for-
mation and microaggregate protection (Kershaw and Talbot, 1998).
Chaplins, released mainly by streptomycetes in phylum Actino-
bacteria, are proteins of bacterial origin that generate hydrophobic
soil surfaces (Gebbink et al., 2005; Sawyer et al., 2011). The hy-
drophobic nature of the above proteins prevents water penetration
into soil organic matter and results in microaggregate formation
(Rillig, 2005). Thus, biochar-enhanced soil carbon sequestration can
be reasonably attributed to biochar functions on soil microbiota
that maintain soil microaggregate stability through the protein
secretion.

4. Fate of soil contaminants under biochar-microbe


interactions

4.1. Immobilization of contaminants by biochars

Fig. 7. Proposed carbon sequestration mechanisms affected by biochar, including: (A)


Biochar is an effective sorbent that immobilizes organic con-
the stable carbon structure of biochar (aromatic carbon structure and crystal silicon
structure in silica-carbon complexes) (Guo and Chen, 2014; Xiao et al., 2014); (B) the taminants and heavy metals through various mechanisms (Ahmad
reaction of biochar with soil minerals, which forms a complex structure that protects et al., 2014; Chen and Chen, 2009; Chen et al., 2008a, 2008b, 2012b,
biochar from microbial degradation (Chen et al., 2015; Li et al., 2014); (C) biochar 2012c). Electrostatic attraction, polar and non-polar organic-
adsorption of soil organic matter (SOM), forming aggregates that protect the degra- attraction to the carbonized phase of biochar, and partitioning to
dation of SOM (George et al., 2012); (D) protection of soil aggregates by fungal hyphae
and the secreted glomalin (King, 2011); (E) modification of soil enzyme activities that
the non-carbonized phase of biochar are forces involved in the
control soil organic carbon decomposition (Paz-Ferreiro et al., 2015). These processes interactions of biochar with organic contaminants (Chen and Chen,
can reduce the CO2 emission from the decomposition of biochar and SOM. 2009; Chen et al., 2008a; Huang and Chen, 2010). Ion exchange,
110 X. Zhu et al. / Environmental Pollution 227 (2017) 98e115

anionic metal attraction, precipitation, and cationic metal attrac- electron-accepting capability (EAC) of the electron acceptors, e.g.,
tion are efficient mechanisms for the interaction of biochar with quinonyl moieties (Kluepfel et al., 2014; Yang et al., 2016a). Biochar
inorganic contaminants (Ahmad et al., 2014, 2016; Cao et al., 2009; could store hundreds of micromoles of electrons per gram of mass,
Qian and Chen, 2013; Xu and Chen, 2015). In some cases, biochar depending on the feedstock and hydrolysis temperatures during
can be even more effective than activated carbon to immobilize the process of accepting electrons, and could further donate them
heavy metals, e.g., a study found six times more effective Pb to minerals, causing their reduction (Kappler et al., 2014). It is
sorption (up to 680 mmol Pb kg1) by dairy manure-derived bio- suggested that the redox activity of biochar was related to the
char than activated carbon, mainly attributed to Pb precipitation water-soluble organic components in biochar suspensions and
(84e87%) in form of bePb9(PO4)6 and Pb3(CO3)2 as well as surface dissolved organic matter (DOM), which is composed of phenols
sorption (13e16%) via C¼C (p-electron) and eOePb bonds through (Graber et al., 2014). In contrast to the dissolved compounds that
interactions with biochar (Cao et al., 2009). The sorption of certain can be extracted with water from biochar particles, the solid phase
herbicides (such as simazine) in micropores or on the surfaces of of biochar particles was recognized to be responsible for the elec-
biochar prevents their accessibility to microbial cells, extracellular tron transfer that caused the microbial ferrihydrite reduction
enzymes, and plants, in addition to preventing them from leaching (Kappler et al., 2014). Direct contact between microbial cells,
into the groundwater (Jones et al., 2011). A simple modification of pollutant molecules, and biochar particles through surface
biochar with iron oxide can endow such magnetic biochar with a adsorption can favor the electron transfer between them and
hybrid sorption capability of organic pollutants and phosphate, further accelerate the pollutant degradation, which is a process that
which enables biochar a multifunctional material for agricultural needs further investigation. Changes in the redox activity of biochar
and environmental applications (Chen et al., 2011). are able to enhance the direct interspecific electron transfer be-
Recent research reveal the novel roles of biochar on alleviating tween microbial cells (DIET) and direct extracellular electron
the aluminum (Al) toxicity (Qian and Chen, 2013) by modifying the transfer between microbial cells with minerals, organic matter,
speciation of Al(OH)2þ or Al(OH)3þ, rather than by a direct elec- contaminants, and heavy metals (DEET), resulting in the degrada-
trostatic attraction of Al3þ with negatively charged sites on biochar tion of organic pollutants (e.g., pentachlorophenol) and the
(Qian et al., 2013). The oxidation of biochar surfaces during the decomposition of organic matter (e.g., humic acids) (Fig. 8).
aging process can provide extra binding sites for heavy metals Cd Together with RAMs, the electrical conductivity (EC) of biochar
and Al, though the competition sorption between Cd and Al exist contributes to its role as an electron shuttle between microbial cells
(Qian and Chen, 2014; Qian et al., 2015). Novel roles of biogenic and pollutants, with a proposed mechanism that involves the for-
silicon in biochars are found to alleviate Al phytotoxicity and to mation of a p-p network within the graphite-like aromatic struc-
prevent plant uptake of Al, indicating that Si particles can reduce ture of biochar, which has a high EC. The shuttle function of biochar
the amount of soil exchangeable Al and that Si released from bio- is responsible for the electron transfer to the organic compounds
chars can form Si-Al compounds in the epidermis of wheat roots (e.g., PCP) that are adsorbed on its surface (Chen et al., 2014; Yu
(Qian et al., 2016). The above research show the necessity to further et al., 2015), as shown in Fig. 8. Biochar had various ECs, depend-
investigate the elemental composition of biochar and the role of ing on feedstock type and pyrolysis temperatures, roughly ranging
biochar on the biogeochemical cycles of soil elements, which can be from 2.11 to 4.41 mS cm1 g1 in pine biochar and much higher
involved with soil microbes. (from 216 to 2217 mS cm1) in manure biochar (Cantrell et al., 2012;
Chen et al., 2014).
4.2. Influences on contaminant transformation and dissipation In addition to the biochar-facilitated microbial degradation of
hazardous pollutants, biochar can reduce the adverse effects from
Evidence has accumulated that biochars are able to act as organic contaminants and heavy metals on soil microbes and plants
electron shuttles between microorganisms and pollutants to by adsorbing such hazards and further directing electron transfer
enhance microbial degradation (Fig. 8) (Yu et al., 2015). Biochar can between them (Choppala et al., 2012). For metal(loid) ions, such as
bridge the electron transfer from the microbial cells to the mineral Cr and As, whose toxicity depends on their valence, the regulation
(Fig. 8) (Kappler et al., 2014), acting as the direct electron acceptor of electron transfer by biochar may control their fate and risks
of acetate for microbial extracellular respiration and growth (Yu (Fig. 8). Biochar DOM serves both as electron donor and acceptor to
et al., 2015) and as the electron donator to stimulate the microbi- reduce toxic Cr(VI) to the less toxic and relatively immobile Cr(III)
al reduction of the Fe(III) oxyhydroxide mineral ferrihydrite (Fig. 8) and to oxidize As(III) to As(V). This function is most likely due to the
(Kappler et al., 2014). When the reduction occurs as a result of the high reactivity of Cr with many functional groups in biochar DOM
electron transfer from Geobacter sulfurreducens, biochar may that are able to donate electrons to reduce Cr(VI) and oxidize As(III)
further serve as an electron donator for the metabolism of other (Choppala et al., 2012; Dong et al., 2014). Losses of Cr(VI) from
microorganisms, in such way to catalyze biochemical processes suspensions of biochar-applied soil and decreased toxicity, induced
(Kappler et al., 2014; Yu et al., 2015). The presence of biochar has by Cr, to sunflowers has been found as well (Choppala et al., 2012).
significantly enhanced the degradation of pentachlorophenol (PCP) By immobilizing Cr and reducing its concentration in soil pore
by a species of bacteria (Geobacter Sulfurreducens), which is hardly water, biochar application can mitigate Cr contamination and
able to degrade PCP alone; this enhancement was ascribed to enhance soil microbial activity and sunflower plant growth
biochar-facilitated electron transfer between microbial cells and (Choppala et al., 2012).
PCP molecule (Yu et al., 2015). The activation of persistent free radicals formed in biochar
The electron transfer capabilities of biochar are influenced by (Fig. 5) is commonly considered to be the mechanism of biochar-
the pyrolysis temperature; the effects are mainly attributed to the facilitated contaminant degradation (Fang et al., 2014a, 2015a,
phenolic moieties formed under low temperatures and quinonyl 2015b). The oxidation of biochar during the aging process re-
groups formed under moderate and high pyrolysis temperatures leases redox-active, acidic and phenolic organics, and the interac-
(Fig. 8) (Kluepfel et al., 2014). The pool of redox-active moieties tion with transition metals produces PFRs, making biochar
(RAMs) in biochar is dominated by different components and is continuously participate in redox reactions and hazardous organic
quantified by its electron exchange capability (EEC), which is compound degradation (Fang et al., 2014a; Graber et al., 2014). Free
mainly composed of the electron-donating capability (EDC) of the radicals in biochar exhibited excellent reactivity to generate OH
electron donators, e.g., phenolic moieties in biochars, and the with H2O2 decomposition, consuming approximately 12 spins of
X. Zhu et al. / Environmental Pollution 227 (2017) 98e115 111

Fig. 8. Schematic diagram showing mechanisms of biochar acting as an electron shuttle enhancing direct interspecific electron transfer between microbial cells (DIET) and direct
extracellular electron transfer between microbial cells with minerals, organic matter, contaminants, and heavy metals (DEET) (Chen et al., 2014; Dong et al., 2014; Kluepfel et al.,
2014; Xu et al., 2016; Yu et al., 2015). Electrons can be transferred from one microbial cell to the component with an electron-accepting capacity (EAC, e.g., quinone) in biochar,
shuttled between EAC and the component with an electron-donating capacity (EDC, e.g., phenol) and be accepted either by another microbial cell, the organic contaminant
(pentachlorophenol), the hematite minerals, or humic acids, enhancing their transformation or degradation. The graphite-like structure on biochar can realize the conductivity.
Electron transfer occurring in biochar dissolved organic matter (DOM) facilitates the transformation of metal(loid)s to nontoxic forms: Cr(VI) to Cr(III) and As(III) to As(V).

free radicals to trap one [OH] molecule (Fang et al., 2015b); sulfate changes and predators) to ensure microbial growth that contrib-
radicals (SO 4 ), formed by persulfate activation, could efficiently utes to contaminant degradation; (2) persistent free radicals could
degrade contaminants including 2-chlorobiphenyl (2-CB), poly- assist biochar as a catalyst and as an electron shuttle to enhance the
chlorinated biphenyls (PCBs), and diethyl phthalate (DEP) (Fang electron transfer between the microbial cells and pollutant mole-
et al., 2014a, 2015a, 2015b). cules, thus facilitating heavy metal transformation and organic
As a practical application, the combination of biochar and mi- pollutant degradation (Kappler et al., 2014; Yu et al., 2015).
crobial remediation strategies is proposed as an effective method
for soil organic contaminant remediation (e.g., for soil PAH pollu-
5. Perspective and conclusions
tion management) (Chen and Ding, 2012; Chen et al., 2012a;
Garcia-Delgado et al., 2015). Using biochar as carriers of PAH-
Heterogeneous effects on soil remediation are found from the
degrading bacteria (Pseudomonas putida, Pseudomonas aeruginosa,
complex combination of biochar and soils with various properties.
Acinetobacter radioresistens, and an unidentified indigenous bac-
Through the massive research with respect to biochar effects on soil
terium), the immobilized-microorganism technique (IMT) finds its
physiochemical characters and microbial activities, the most
effective application on alleviating soil PAHs contamination (Chen
influential factors point to feedstock types and pyrolysis tempera-
et al., 2012a; Galitskaya et al., 2016). For the strategy of combined
ture, which determine biochar properties including surface prop-
biochar-microbial remediation, biochars are used as a pretreatment
erties, structures, elemental composition, redox capacity,
to immobilize and concentrate organic contaminant PAHs, followed
conductivity, pH, CEC, and VOCs. Such properties play key roles in
by an inoculation of PAH-degrading microbes (e.g., white rot fungus
microbial activity and soil process. According to the specific pur-
Phanerochaete chrysosporium and Pleurotus Ostreatus) to perform
poses, different biochar types should be considered. To improve soil
the final degradation of the PAHs in soil (Chen and Ding, 2012;
pH, especially for acidic soils, the utilization of wood biochar pro-
Garcia-Delgado et al., 2015). The sequential application of biochar
duced under high temperature can be proper as it loses most of the
and white rot fungi has induced higher degradation of phenan-
P P acidic functional groups during pyrolysis. To enhance soil fertility,
threne, anthracene, fluoranthene, pyrene, 3-rings, 4-rings, and
P crop residue biochar and manure biochar with high pyrolysis
13 PAH than the single addition of this species of fungi (Chen and
temperatures have the priority, because their high ash content can
Ding, 2012; Garcia-Delgado et al., 2015). Two possible mechanisms
improve soil CEC, facilitating nutrient maintenance for microbial
that could be involved in the influence of biochar on the microbial
and plant growth. To enhance carbon sequestration, high temper-
dissipation of contaminants: (1) the attachment of microbial cells
ature biochar, especially from woods, should be considered because
to biochar particles and biochar-enhanced soil microaggregates
they generally have stable carbon structure and high C/N ratio that
that provide suitable habitats and prevent harsh environmental
make them recalcitrant to microbial decomposition. To control
changes (such as pH changes, water retention, temperature
soilborne pathogens, low temperature biochars may have the
112 X. Zhu et al. / Environmental Pollution 227 (2017) 98e115

advantages, because the relatively high small molecular VOCs on biochar stability and soil organisms: review and research needs. Eur. J. Soil Sci.
64, 379e390.
such biochar can be microbial inhibitors. Moreover, moderate
Ameloot, N., Sleutel, S., Case, S.D.C., Alberti, G., McNamara, N.P., Zavalloni, C.,
temperature biochar with relatively higher adsorption capacity and Vervisch, B., delle Vedove, G., De Neve, S., 2014. C mineralization and microbial
electron transfer capability can be considered for the purpose of soil activity in four biochar field experiments several years after incorporation. Soil
contamination remediation. Biol. Biochem. 78, 195e203.
Ameloot, N., Sleutel, S., Das, K.C., Kanagaratnam, J., de Neve, S., 2015. Biochar
Understanding the key roles of biochar properties on microbial amendment to soils with contrasting organic matter level: effects on N
activity in different soil types is essential to answering the most mineralization and biological soil properties. Glob. Change Biol. Bioenergy 7,
challenging questions, such as: (1) Under which conditions can the 135e144.
Ascough, P.L., Sturrock, C.J., Bird, M.I., 2010. Investigation of growth responses in
application of biochar reach the desired benefits, and how can the saprophytic fungi to charred biomass. Isot. Environ. Health Stud. 46, 64e77.
trade-offs between various environmental effects of biochar be Awad, Y.M., Blagodatskaya, E., Ok, Y.S., Kuzyakov, Y., 2013. Effects of polyacrylamide,
balanced? (2) Why, mechanistically, is a given biochar beneficial in biopolymer and biochar on the decomposition of C-14-labelled maize residues
and on their stabilization in soil aggregates. Eur. J. Soil Sci. 64, 488e499.
some soils but not in other soils? (3) How can the biological effects Bailey, V.L., Fansler, S.J., Smith, J.L., Bolton Jr., H., 2011. Reconciling apparent vari-
of biochar be better predicted? Focused on certain key functions of ability in effects of biochar amendment on soil enzyme activities by assay
biochar, this review synthesizes the biochar-microbe interaction optimization. Soil Biol. Biochem. 43, 296e301.
Balakrishna, S., Lomnicki, S., McAvey, K.M., Cole, R.B., Dellinger, B., Cormier, S.A.,
mechanisms pertinent to those questions, covering: (1) the mi- 2009. Environmentally persistent free radicals amplify ultrafine particle
crobial community modification by biochar, via alteration of mediated cellular oxidative stress and cytotoxicity. Part. Fibre Toxicol. http://
nutrient availability and soil characters, (2) the electron transfer dx.doi.org/10.1186/1743-8977-6-11.
Bamminger, C., Zaiser, N., Zinsser, P., Lamers, M., Kammann, C., Marhan, S., 2014.
between microbial cells and contaminants facilitated by biochar,
Effects of biochar, earthworms, and litter addition on soil microbial activity and
(3) the potential microbial toxicity triggered by biochar, (4) the abundance in a temperate agricultural soil. Biol. Fertil. Soils 50, 1189e1200.
carbon stability in biochar with regard to the role of microbial Bandick, A.K., Dick, R.P., 1999. Field management effects on soil enzyme activities.
degradation and the effects of biochar on the soil organic carbon Soil Biol. Biochem. 31, 1471e1479.
Beesley, L., Moreno-Jime nez, E., Gomez-Eyles, J.L., 2010. Effects of biochar and
stability, and (5) the mechanisms of biochar-modified plant resis- greenwaste compost amendments on mobility, bioavailability and toxicity of
tance to soilborne pathogens. New developments in the meta- inorganic and organic contaminants in a multi-element polluted soil. Environ.
genomics analysis of soil microbe genes will no doubt be pivotal to Pollut. 158, 2282e2287.
Beesley, L., Moreno-Jimenez, E., Gomez-Eyles, J.L., Harris, E., Robinson, B., Sizmur, T.,
uncovering the hidden dimensions of biochar-microbe interactions 2011. A review of biochars' potential role in the remediation, revegetation and
that link biochar properties with microbial functions. Under- restoration of contaminated soils. Environ. Pollut. 159, 3269e3282.
standing the mechanisms of the interaction between biochar and Biederman, L.A., Harpole, W.S., 2013. Biochar and its effects on plant productivity
and nutrient cycling: a meta-analysis. Glob. Change Biol. Bioenergy 5, 202e214.
soil microbes is essential to revealing the mechanisms of the het- Cantrell, K.B., Hunt, P.G., Uchimiya, M., Novak, J.M., Ro, K.S., 2012. Impact of pyrolysis
erogeneous effects of biochar on soil remediation. temperature and manure source on physicochemical characteristics of biochar.
Bioresour. Technol. 107, 419e428.
Cao, X.D., Harris, W., 2010. Properties of dairy-manure-derived biochar pertinent to
Competing financial interests its potential use in remediation. Bioresour. Technol. 101, 5222e5228.
Cao, X.D., Ma, L.N., Gao, B., Harris, W., 2009. Dairy-manure derived biochar effec-
The authors declare that they have no competing interests. tively sorbs lead and atrazine. Environ. Sci. Technol. 43, 3285e3291.
Cely, P., Tarquis, A.M., Paz-Ferreiro, J., Mendez, A., Gasco, G., 2014. Factors driving
the carbon mineralization priming effect in a sandy loam soil amended with
Acknowledgements different types of biochar. Solid Earth 5, 585e594.
Chathurika, J.A.S., Kumaragamage, D., Zvomuya, F., Akinremi, O.O., Flaten, D.N.,
Indraratne, S.P., Dandeniya, W.S., 2016. Woodchip biochar with or without
This project was supported by the National Natural Science synthetic fertilizers affects soil properties and available phosphorus in two
Foundation of China (Grant 21425730, 21537005, 21621005, and alkaline, chernozemic soils. Can. J. Soil Sci. 96, 472e484.
21607125), The National Key Technology R&D Program Chen, B., Chen, Z., 2009. Sorption of naphthalene and 1-naphthol by biochars of
orange peels with different pyrolytic temperatures. Chemosphere 76, 127e133.
(2015BAC02B01), and the Postdoctoral Science Foundation of China Chen, B., Chen, Z., Lv, S., 2011. A novel magnetic biochar efficiently sorbs organic
(Grant 2015M581943). Special thanks give to the anonymous re- pollutants and phosphate. Bioresour. Technol. 102, 716e723.
viewers who contribute greatly to improve the quality of this Chen, B., Ding, J., 2012. Biosorption and biodegradation of phenanthrene and pyrene
in sterilized and unsterilized soil slurry systems stimulated by Phanerochaete
review.
chrysosporium. J. Hazard. Mater 229, 159e169.
Chen, B., Yuan, M., 2011. Enhanced sorption of polycyclic aromatic hydrocarbons by
Appendix ASupplementary data soil amended with biochar. J. Soils Sed. 11, 62e71.
Chen, B., Yuan, M., Qian, L., 2012a. Enhanced bioremediation of PAH-contaminated
soil by immobilized bacteria with plant residue and biochar as carriers. J. Soils
Supplementary data related to this article can be found at http:// Sed. 12, 1350e1359.
dx.doi.org/10.1016/j.envpol.2017.04.032. Chen, B., Zhou, D., Zhu, L., 2008a. Transitional adsorption and partition of nonpolar
and polar aromatic contaminants by biochars of pine needles with different
pyrolytic temperatures. Environ. Sci. Technol. 42, 5137e5143.
References Chen, B., Zhou, D., Zhu, L., Shen, X., 2008b. Sorption characteristics and mechanisms
of organic contaminant to carbonaceous biosorbents in aqueous solution. Sci.
Abel, S., Peters, A., Trinks, S., Schonsky, H., Facklam, M., Wessolek, G., 2013. Impact China Ser. B-Chem. 51, 464e472.
of biochar and hydrochar addition on water retention and water repellency of Chen, J.H., Liu, X.Y., Zheng, J.W., Zhang, B., Lu, H.F., Chi, Z.Z., Pan, G.X., Li, L.Q.,
sandy soil. Geoderma 202, 183e191. Zheng, J.F., Zhang, X.H., Wang, J.F., Yu, X.Y., 2013. Biochar soil amendment
Abit, S.M., Bolster, C.H., Cai, P., Walker, S.L., 2012. Influence of feedstock and py- increased bacterial but decreased fungal gene abundance with shifts in com-
rolysis temperature of biochar amendments on transport of Escherichia coli in munity structure in a slightly acid rice paddy from Southwest China. Appl. Soil
saturated and unsaturated soil. Environ. Sci. Technol. 46, 8097e8105. Ecol. 71, 33e44.
Ahmad, M., Ok, Y.S., Kim, B.-Y., Ahn, J.-H., Lee, Y.H., Zhang, M., Moon, D.H., Al- Chen, S.S., Rotaru, A.E., Shrestha, P.M., Malvankar, N.S., Liu, F.H., Fan, W., Nevin, K.P.,
Wabel, M.I., Lee, S.S., 2016. Impact of soybean stover- and pine needle-derived Lovley, D.R., 2014. Promoting interspecies electron transfer with biochar. Sci.
biochars on Pb and As mobility, microbial community, and carbon stability in a Rep. http://dx.doi.org/10.1038/srep05019.
contaminated agricultural soil. J. Environ. Manag. 166, 131e139. Chen, Z., Chen, B., Chiou, C.T., 2012b. Fast and slow rates of naphthalene sorption to
Ahmad, M., Rajapaksha, A.U., Lim, J.E., Zhang, M., Bolan, N., Mohan, D., biochars produced at different temperatures. Environ. Sci. Technol. 46,
Vithanage, M., Lee, S.S., Ok, Y.S., 2014. Biochar as a sorbent for contaminant 11104e11111.
management in soil and water: a review. Chemosphere 99, 19e33. Chen, Z., Chen, B., Zhou, D., Chen, W., 2012c. Bisolute sorption and thermodynamic
Akhter, A., Hage-Ahmed, K., Soja, G., Steinkellner, S., 2015. Compost and biochar behavior of organic pollutants to biomass-derived biochars at two pyrolytic
alter mycorrhization, tomato root exudation, and development of Fusarium temperatures. Environ. Sci. Technol. 46, 12476e12483.
oxysporum f. sp lycopersici. Front. Plant. Sci. http://dx.doi.org/10.3389/ Chen, Z., Fang, Y., Xu, Y., Chen, B., 2012d. Adsorption of Pb2þ by rice straw derived-
fpls.2015.00529. biochar and its influential factors. Acta Sci. Circumst. 32, 769e776.
Ameloot, N., Graber, E.R., Verheijen, F.G.A., De Neve, S., 2013. Interactions between Chen, Z., Xiao, X., Chen, B., Zhu, L., 2015. Quantification of chemical states,
X. Zhu et al. / Environmental Pollution 227 (2017) 98e115 113

dissociation constants and contents of oxygen-containing groups on the surface biochars: interactive effects of carbon and silicon components. Environ. Sci.
of biochars produced at different temperatures. Environ. Sci. Technol. 49, Technol. 48, 9103e9112.
309e317. Hale, L., Luth, M., Crowley, D., 2015. Biochar characteristics relate to its utility as an
Choppala, G.K., Bolan, N.S., Megharaj, M., Chen, Z., Naidu, R., 2012. The influence of alternative soil inoculum carrier to peat and vermiculite. Soil Biol. Biochem. 81,
biochar and black carbon on reduction and bioavailability of chromate in soils. 228e235.
J. Environ. Qual. 41, 1175e1184. Hale, L., Luth, M., Kenney, R., Crowley, D., 2014. Evaluation of pinewood biochar as a
Czimczik, C.I., Masiello, C.A., 2007. Controls on black carbon storage in soils. Glob. carrier of bacterial strain Enterobacter cloacae UW5 for soil inoculation. Appl.
Biogeochem. Cycl. http://dx.doi.org/10.1029/2006gb002798. Soil Ecol. 84, 192e199.
de Andrade, C.A., Silveira Bibar, M.P., Coscione, A.R., Moreno Pires, A.M., Soares, A.G., Hale, S.E., Hanley, K., Lehmann, J., Zimmerman, A.R., Cornelissen, G., 2011. Effects of
2015. Mineralization and effects of poultry litter biochar on soil cation exchange chemical, biological, and physical aging as well as soil addition on the sorption
capacity. Pesq. Agropecu. Bras. 50, 407e416. of pyrene to activated carbon and biochar. Environ. Sci. Technol. 45,
Dellinger, B., Loninicki, S., Khachatryan, L., Maskos, Z., Hall, R.W., Adounkpe, J., 10445e10453.
McFerrin, C., Truong, H., 2007. Formation and stabilization of persistent free Harel, Y.M., Elad, Y., Rav-David, D., Borenstein, M., Shulchani, R., Lew, B., Graber, E.R.,
radicals. Proc. Combust. Inst. 31, 521e528. 2012. Biochar mediates systemic response of strawberry to foliar fungal path-
Dellinger, B., Pryor, W.A., Cueto, R., Squadrito, G.L., Hegde, V., Deutsch, W.A., 2001. ogens. Plant Soil 357, 245e257.
Role of free radicals in the toxicity of airborne fine particulate matter. Chem. Heitkoetter, J., Marschner, B., 2015. Interactive effects of biochar ageing in soils
Res. Toxicol. 14, 1371e1377. related to feedstock, pyrolysis temperature, and historic charcoal production.
Demisie, W., Liu, Z., Zhang, M., 2014. Effect of biochar on carbon fractions and Geoderma 245, 56e64.
enzyme activity of red soil. Catena 121, 214e221. Hua, L., Wang, Y.T., Wang, T.P., Ma, H.R., 2015. Effect of biochar on organic matter
Dong, X., Ma, L.Q., Gress, J., Harris, W., Li, Y., 2014. Enhanced Cr(VI) reduction and conservation and metabolic quotient of soil. Environ. Prog. Sustain. Energy 34,
As(III) oxidation in ice phase: important role of dissolved organic matter from 1467e1472.
biochar. J. Hazard. Mater 267, 62e70. Huang, W., Chen, B., 2010. Interaction mechanisms of organic contaminants with
Doong, R.A., Lee, C.C., Lien, C.M., 2014. Enhanced dechlorination of carbon tetra- burned straw ash charcoal. J. Environ. Sci. 22, 1586e1594.
chloride by Geobacter sulfurreducens in the presence of naturally occurring €ckel, U., Schnell, S., Conrad, R., 2004. Microbial ethylene production and inhibition
Ja
quinones and ferrihydrite. Chemosphere 97, 54e63. of methanotrophic activity in a deciduous forest soil. Soil Biol. Biochem. 36,
El-Naggar, A.H., Usman, A.R.A., Al-Omran, A., Ok, Y.S., Ahmad, M., Al-Wabel, M.I., 835e840.
2015. Carbon mineralization and nutrient availability in calcareous sandy soils Jeffery, S., Bezemer, T.M., Cornelissen, G., Kuyper, T.W., Lehmann, J., Mommer, L.,
amended with woody waste biochar. Chemosphere 138, 67e73. Sohi, S.P., van de Voorde, T.F.J., Wardle, D.A., van Groenigen, J.W., 2015. The way
Elad, Y., Cytryn, E., Harel, Y.M., Lew, B., Graber, E.R., 2011. The biochar effect: plant forward in biochar research: targeting trade-offs between the potential wins.
resistance to biotic stresses. Phytopathol. Mediterr. 50, 335e349. Glob. Change Biol. Bioenergy 7, 1e13.
Elad, Y., David, D.R., Harel, Y.M., Borenshtein, M., Ben Kalifa, H., Silber, A., Jeong, C.Y., Dodla, S.K., Wang, J.J., 2016. Fundamental and molecular composition
Graber, E.R., 2010. Induction of systemic resistance in plants by biochar, a soil- characteristics of biochars produced from sugarcane and rice crop residues and
applied carbon sequestering agent. Phytopathology 100, 913e921. by-products. Chemosphere 142, 4e13.
Ennis, C.J., Evans, A.G., Islam, M., Ralebitso-Senior, T.K., Senior, E., 2012. Biochar: Jien, S.-H., Wang, C.-C., Lee, C.-H., Lee, T.-Y., 2015. Stabilization of organic matter by
carbon sequestration, land remediation, and impacts on soil microbiology. Crit. biochar application in compost-amended soils with contrasting pH values and
Rev. Environ. Sci. Technol. 42, 2311e2364. textures. Sustainability 7, 13317e13333.
Fang, G.D., Gao, J., Liu, C., Dionysiou, D.D., Wang, Y., Zhou, D.M., 2014a. Key role of Jien, S.-H., Wang, C.-S., 2013. Effects of biochar on soil properties and erosion po-
persistent free radicals in hydrogen peroxide activation by biochar: implications tential in a highly weathered soil. Catena 110, 225e233.
to organic contaminant degradation. Environ. Sci. Technol. 48, 1902e1910. Jones, D.L., Edwards-Jones, G., Murphy, D.V., 2011. Biochar mediated alterations in
Fang, G.D., Liu, C., Gao, J., Dionysiou, D.D., Zhou, D.M., 2015a. Manipulation of herbicide breakdown and leaching in soil. Soil Biol. Biochem. 43, 804e813.
persistent free radicals in biochar to activate persulfate for contaminant Joseph, S., Graber, E.R., Chia, C., Munroe, P., Donne, S., Thomas, T., Nielsen, S.,
degradation. Environ. Sci. Technol. 49, 5645e5653. Marjo, C., Rutlidge, H., Pan, G.X., Li, L., Taylor, P., Rawal, A., Hook, J., 2013. Shifting
Fang, G.D., Zhu, C.Y., Dionysiou, D.D., Gao, J., Zhou, D.M., 2015b. Mechanism of hy- paradigms: development of high-efficiency biochar fertilizers based on nano-
droxyl radical generation from biochar suspensions: implications to diethyl structures and soluble components. Carbon Manag. 4, 323e343.
phthalate degradation. Bioresour. Technol. 176, 210e217. Kappler, A., Wuestner, M.L., Ruecker, A., Harter, J., Halama, M., Behrens, S., 2014.
Fang, Q., Chen, B., Lin, Y., Guan, Y., 2014b. Aromatic and hydrophobic surfaces of Biochar as an electron shuttle between bacteria and Fe(III) minerals. Environ.
wood-derived biochar enhance perchlorate adsorption via hydrogen bonding to Sci. Technol. Lett. 1, 339e344.
oxygen-containing organic groups. Environ. Sci. Technol. 48, 279e288. Kasozi, G.N., Zimmerman, A.R., Nkedi-Kizza, P., Gao, B., 2010. Catechol and humic
Fox, A., Kwapinski, W., Griffiths, B.S., Schmalenberger, A., 2014. The role of sulfur- acid sorption onto a range of laboratory-produced black carbons (biochars).
and phosphorus-mobilizing bacteria in biochar-induced growth promotion of Environ. Sci. Technol. 44, 6189e6195.
Lolium perenne. FEMS Microbiol. Ecol. 90, 78e91. Kershaw, M.J., Talbot, N.J., 1998. Hydrophobins and repellents: proteins with
Galitskaya, P., Akhmetzyanova, L., Selivanovskaya, S., 2016. Biochar-carrying hy- fundamental roles in fungal morphogenesis. Fungal Genet. Biol. 23, 18e33.
drocarbon decomposers promote degradation during the early stage of biore- King, G.M., 2011. Enhancing soil carbon storage for carbon remediation: potential
mediation. Biogeosciences 13, 5739e5752. contributions and constraints by microbes. Trends Microbiol. 19, 75e84.
Galvez, A., Sinicco, T., Cayuela, M.L., Mingorance, M.D., Fornasier, F., Mondini, C., Kluepfel, L., Keiluweit, M., Kleber, M., Sander, M., 2014. Redox properties of plant
2012. Short term effects of bioenergy by-products on soil C and N dynamics, biomass-derived black carbon (biochar). Environ. Sci. Technol. 48, 5601e5611.
nutrient availability and biochemical properties. Agric. Ecosyst. Environ. 160, Kolton, M., Harel, Y.M., Pasternak, Z., Graber, E.R., Elad, Y., Cytryn, E., 2011. Impact of
3e14. biochar application to soil on the root-associated bacterial community structure
Gao, X., Cheng, H.-Y., Del Valle, I., Liu, S., Masiello, C.A., Silberg, J.J., 2016. Charcoal of fully developed greenhouse pepper plants. Appl. Environ. Microbiol. 77,
disrupts soil microbial communication through a combination of signal sorp- 4924e4930.
tion and hydrolysis. Acs Omega 1, 226e233. Koltowski, M., Charmas, B., Skubiszewska-Zieba, J., Oleszczuk, P., 2017. Effect of
Garcia-Delgado, C., Alfaro-Barta, I., Eymar, E., 2015. Combination of biochar biochar activation by different methods on toxicity of soil contaminated by
amendment and mycoremediation for polycyclic aromatic hydrocarbons industrial activity. Ecotoxicol. Environ. Saf. 136, 119e125.
immobilization and biodegradation in creosote-contaminated soil. J. Hazard. Kuzyakov, Y., Friedel, J.K., Stahr, K., 2000. Review of mechanisms and quantification
Mater 285, 259e266. of priming effects. Soil Biol. Biochem. 32, 1485e1498.
Gebbink, M., Claessen, D., Bouma, B., Dijkhuizen, L., Wosten, H.A.B., 2005. Amyloids Laird, D.A., Fleming, P., Davis, D.D., Horton, R., Wang, B., Karlen, D.L., 2010. Impact of
- a functional coat for microorganisms. Nat. Rev. Microbiol. 3, 333e341. biochar amendments on the quality of a typical Midwestern agricultural soil.
George, C., Wagner, M., Kuecke, M., Rillig, M.C., 2012. Divergent consequences of Geoderma 158, 443e449.
hydrochar in the plant-soil system: arbuscular mycorrhiza, nodulation, plant Lammirato, C., Miltner, A., Kaestner, M., 2011. Effects of wood char and activated
growth and soil aggregation effects. Appl. Soil Ecol. 59, 68e72. carbon on the hydrolysis of cellobiose by beta-glucosidase from Aspergillus
Ghidotti, M., Fabbri, D., Hornung, A., 2017. Profiles of volatile organic compounds in niger. Soil Biol. Biochem. 43, 1936e1942.
biochar: insights into process conditions and quality assessment. Acs Sustain. Lauber, C.L., Hamady, M., Knight, R., Fierer, N., 2009. Pyrosequencing-based
Chem. Eng. 5, 510e517. assessment of soil pH as a predictor of soil bacterial community structure at the
Gibson, C., Berry, T.D., Wang, R., Spencer, J.A., Johnston, C.T., Jiang, Y., Bird, J.A., continental scale. Appl. Environ. Microbiol. 75, 5111e5120.
Filley, T.R., 2016. Weathering of pyrogenic organic matter induces fungal Leal, O.D., Dick, D.P., Lombardi, K.C., Maciel, V.G., Gonzalez-Perez, J.A., Knicker, H.,
oxidative enzyme response in single culture inoculation experiments. Org. 2015. Soil chemical properties and organic matter composition of a subtropical
Geochem 92, 32e41. Cambisol after charcoal fine residues incorporation. J. Soils Sed. 15, 805e815.
Graber, E.R., Harel, Y.M., Kolton, M., Cytryn, E., Silber, A., David, D.R., Tsechansky, L., Lehmann, J., 2007a. Bio-energy in the black. Front. Ecol. Environ. 5, 381e387.
Borenshtein, M., Elad, Y., 2010. Biochar impact on development and produc- Lehmann, J., 2007b. A handful of carbon. Nature 447, 143e144.
tivity of pepper and tomato grown in fertigated soilless media. Plant Soil 337, Lehmann, J., Rillig, M.C., Thies, J., Masiello, C.A., Hockaday, W.C., Crowley, D., 2011.
481e496. Biochar effects on soil biota - a review. Soil Biol. Biochem. 43, 1812e1836.
Graber, E.R., Tsechansky, L., Lew, B., Cohen, E., 2014. Reducing capacity of water Li, F.Y., Cao, X.D., Zhao, L., Wang, J.F., Ding, Z.L., 2014. Effects of mineral additives on
extracts of biochars and their solubilization of soil Mn and Fe. Eur. J. Soil Sci. 65, biochar formation: carbon retention, stability, and properties. Environ. Sci.
162e172. Technol. 48, 11211e11217.
Guo, J., Chen, B., 2014. Insights on the molecular mechanism for the recalcitrance of Liang, B., Lehmann, J., Solomon, D., Kinyangi, J., Grossman, J., O'Neill, B.,
114 X. Zhu et al. / Environmental Pollution 227 (2017) 98e115

Skjemstad, J.O., Thies, J., Luizao, F.J., Petersen, J., Neves, E.G., 2006. Black Carbon microorganisms? Soil Biol. Biochem. 65, 287e293.
increases cation exchange capacity in soils. Soil Sci. Soc. Am. J. 70, 1719e1730. Quilliam, R.S., Rangecroft, S., Emmett, B.A., Deluca, T.H., Jones, D.L., 2013b. Is biochar
Liang, C., Zhu, X., Fu, S., Mendez, A., Gasco, G., Paz-Ferreiro, J., 2014. Biochar alters a source or sink for polycyclic aromatic hydrocarbon (PAH) compounds in
the resistance and resilience to drought in a tropical soil. Environ. Res. Lett. agricultural soils? Glob. Change Biol. Bioenergy 5, 96e103.
http://dx.doi.org/10.1088/1748e9326/9/6/064013. Reed, J.R., dela Cruz, A.L.N., Lomnicki, S.M., Backes, W.L., 2015. Environmentally
Liao, S.H., Pan, B., Li, H., Zhang, D., Xing, B.S., 2014. Detecting free radicals in biochars persistent free radical-containing particulate matter competitively inhibits
and determining their ability to inhibit the germination and growth of corn, metabolism by cytochrome P450 1A2. Toxicol. Appl. Pharmacol. 289, 223e230.
wheat and rice seedlings. Environ. Sci. Technol. 48, 8581e8587. Rillig, M.C., 2005. A connection between fungal hydrophobins and soil water
Lin, Y., Munroe, P., Joseph, S., Kimber, S., Van Zwieten, L., 2012. Nanoscale organo- repellency? Pedobiologia 49, 395e399.
mineral reactions of biochars in ferrosol: an investigation using microscopy. Rillig, M.C., Lutgen, E.R., Ramsey, P.W., Klironomos, J.N., Gannon, J.E., 2005. Micro-
Plant Soil 357, 369e380. biota accompanying different arbuscular mycorrhizal fungal isolates influence
Lu, K., Yang, X., Gielen, G., Bolan, N., Ok, Y.S., Niazi, N.K., Xu, S., Yuan, G., Chen, X., soil aggregation. Pedobiologia 49, 251e259.
Zhang, X., Liu, D., Song, Z., Liu, X., Wang, H., 2017. Effect of bamboo and rice Rillig, M.C., Wright, S.F., Eviner, V.T., 2002. The role of arbuscular mycorrhizal fungi
straw biochars on the mobility and redistribution of heavy metals (Cd, Cu, Pb and glomalin in soil aggregation: comparing effects of five plant species. Plant
and Zn) in contaminated soil. J. Environ. Manag. 186, 285e292. Soil 238, 325e333.
Luo, L., Gu, J.-D., 2016. Alteration of extracellular enzyme activity and microbial Rittl, T.F., Novotny, E.H., Balieiro, F.C., Hoffland, E., Alves, B.J.R., Kuyper, T.W., 2015.
abundance by biochar addition: implication for carbon sequestration in sub- Negative priming of native soil organic carbon mineralization by oilseed bio-
tropical mangrove sediment. J. Environ. Manag. 182, 29e36. chars of contrasting quality. Eur. J. Soil Sci. 66, 714e721.
Lyu, H., He, Y., Tang, J., Hecker, M., Liu, Q., Jones, P.D., Codling, G., Giesy, J.P., 2016. Rodriguez-Vila, A., Forjan, R., Guedes, R.S., Covelo, E.F., 2016. Changes on the phy-
Effect of pyrolysis temperature on potential toxicity of biochar if applied to the toavailability of nutrients in a mine soil reclaimed with compost and biochar.
environment. Environ. Pollut. 218, 1e7. Water Air Soil Pollut. http://dx.doi.org/10.1007/s11270-016-3155-x.
Mackie, K.A., Marhan, S., Ditterich, F., Schmidt, H.P., Kandeler, E., 2015. The effects of Ronsse, F., van Hecke, S., Dickinson, D., Prins, W., 2013. Production and character-
biochar and compost amendments on copper immobilization and soil micro- ization of slow pyrolysis biochar: influence of feedstock type and pyrolysis
organisms in a temperate vineyard. Agric. Ecosyst. Environ. 201, 58e69. conditions. Glob. Change Biol. Bioenergy 5, 104e115.
Major, J., Lehmann, J., Rondon, M., Goodale, C., 2010. Fate of soil-applied black Rousk, J., Baath, E., Brookes, P.C., Lauber, C.L., Lozupone, C., Caporaso, J.G., Knight, R.,
carbon: downward migration, leaching and soil respiration. Glob. Change Biol. Fierer, N., 2010. Soil bacterial and fungal communities across a pH gradient in
16, 1366e1379. an arable soil. Isme J. 4, 1340e1351.
Masek, O., Brownsort, P., Cross, A., Sohi, S., 2013. Influence of production conditions Rousk, J., Brookes, P.C., Baath, E., 2009. Contrasting soil pH effects on fungal and
on the yield and environmental stability of biochar. Fuel 103, 151e155. bacterial growth suggest functional redundancy in carbon mineralization. Appl.
Masiello, C.A., Chen, Y., Gao, X., Liu, S., Cheng, H.-Y., Bennett, M.R., Rudgers, J.A., Environ. Microbiol. 75, 1589e1596.
Wagner, D.S., Zygourakis, K., Silberg, J.J., 2013. Biochar and microbial signaling: Rutigliano, F.A., Romano, M., Marzaioli, R., Baglivo, I., Baronti, S., Miglietta, F.,
production conditions determine effects on microbial communication. Environ. Castaldi, S., 2014. Effect of biochar addition on soil microbial community in a
Sci. Technol. 47, 11496e11503. wheat crop. Eur. J. Soil Biol. 60, 9e15.
Mukherjee, A., Zimmerman, A.R., 2013. Organic carbon and nutrient release from a Sawyer, E.B., Claessen, D., Haas, M., Hurgobin, B., Gras, S.L., 2011. The assembly of
range of laboratory-produced biochars and biochar-soil mixtures. Geoderma individual chaplin peptides from Streptomyces coelicolor into functional amy-
193, 122e130. loid fibrils. PLoS One. http://dx.doi.org/10.1371/journal.pone.0018839.
Mukherjee, A., Zimmerman, A.R., Hamdan, R., Cooper, W.T., 2014. Physicochemical Seneviratne, M., Weerasundara, L., Ok, Y.S., Rinklebe, J., Vithanage, M., 2017.
changes in pyrogenic organic matter (biochar) after 15 months of field aging. Phytotoxicity attenuation in Vigna radiata under heavy metal stress at the
Solid earth. 5, 693e704. presence of biochar and N fixing bacteria. J. Environ. Manag. 186, 293e300.
Mukherjee, A., Zimmerman, A.R., Harris, W., 2011. Surface chemistry variations Silber, A., Levkovitch, I., Graber, E.R., 2010. pH-dependent mineral release and
among a series of laboratory-produced biochars. Geoderma 163, 247e255. surface properties of cornstraw biochar: agronomic implications. Environ. Sci.
Murray, J., Keith, A., Singh, B., 2015. The stability of low- and high-ash biochars in Technol. 44, 9318e9323.
acidic soils of contrasting mineralogy. Soil Biol. Biochem. 89, 217e225. Smebye, A., Alling, V., Vogt, R.D., Gadmar, T.C., Mulder, J., Cornelissen, G., Hale, S.E.,
Nannipieri, P., Giagnoni, L., Renella, G., Puglisi, E., Ceccanti, B., Masciandaro, G., 2016. Biochar amendment to soil changes dissolved organic matter content and
Fornasier, F., Moscatelli, M.C., Marinari, S., 2012. Soil enzymology: classical and composition. Chemosphere 142, 100e105.
molecular approaches. Biol. Fertil. Soils 48, 743e762. Spokas, K.A., Novak, J.M., Stewart, C.E., Cantrell, K.B., Uchimiya, M., DuSaire, M.G.,
Nielsen, S., Minchin, T., Kimber, S., van Zwieten, L., Gilbert, J., Munroe, P., Joseph, S., Ro, K.S., 2011. Qualitative analysis of volatile organic compounds on biochar.
Thomas, T., 2014. Comparative analysis of the microbial communities in agri- Chemosphere 85, 869e882.
cultural soil amended with enhanced biochars or traditional fertilisers. Agric. Stefaniuk, M., Oleszczuk, P., 2016. Addition of biochar to sewage sludge decreases
Ecosyst. Environ. 191, 73e82. freely dissolved PAHs content and toxicity of sewage sludge-amended soil.
Novak, J.M., Busscher, W.J., Laird, D.L., Ahmedna, M., Watts, D.W., Niandou, M.A.S., Environ. Pollut. 218, 242e251.
2009. Impact of biochar amendment on fertility of a southeastern coastal plain Steinbeiss, S., Gleixner, G., Antonietti, M., 2009. Effect of biochar amendment on soil
soil. Soil Sci. 174, 105e112. carbon balance and soil microbial activity. Soil Biol. Biochem. 41, 1301e1310.
Oh, S.Y., Son, J.G., Hur, S.H., Chung, J.S., Chiu, P.C., 2013. Black carbon-mediated Sun, D., Meng, J., Liang, H., Yang, E., Huang, Y., Chen, W., Jiang, L., Lan, Y., Zhang, W.,
reduction of 2,4-Dinitrotoluene by dithiothreitol. J. Environ. Qual. 42, 815e821. Gao, J., 2015. Effect of volatile organic compounds absorbed to fresh biochar on
Paz-Ferreiro, J., Fu, S., Mendez, A., Gasco, G., 2015. Biochar modifies the thermo- survival of Bacillus mucilaginosus and structure of soil microbial communities.
dynamic parameters of soil enzyme activity in a tropical soil. J. Soils Sed. 15, J. Soils Sed. 15, 271e281.
578e583. Sun, F., Lu, S., 2014. Biochars improve aggregate stability, water retention, and pore-
Paz-Ferreiro, J., Fu, S.L., Mendez, A., Gasco, G., 2014. Interactive effects of biochar space properties of clayey soil. J. Plant Nutr. Soil Sci. 177, 26e33.
and the earthworm Pontoscolex corethrurus on plant productivity and soil Truong, H., Lomnicki, S., Dellinger, B., 2010. Potential for misidentification of envi-
enzyme activities. J. Soils Sed. 14, 483e494. ronmentally persistent free radicals as molecular pollutants in particulate
Paz-Ferreiro, J., Gasco, G., Gutierrez, B., Mendez, A., 2012. Soil biochemical activities matter. Environ. Sci. Technol. 44, 1933e1939.
and the geometric mean of enzyme activities after application of sewage sludge Wang, S.S., Gao, B., Zimmerman, A.R., Li, Y.C., Ma, L.N., Harris, W.G., Migliaccio, K.W.,
and sewage sludge biochar to soil. Biol. Fertil. Soils 48, 511e517. 2015. Physicochemical and sorptive properties of biochars derived from woody
Purakayastha, T.J., Kumari, S., Pathak, H., 2015. Characterisation, stability, and mi- and herbaceous biomass. Chemosphere 134, 257e262.
crobial effects of four biochars produced from crop residues. Geoderma 239, Warnock, D.D., Lehmann, J., Kuyper, T.W., Rillig, M.C., 2007. Mycorrhizal responses
293e303. to biochar in soil - concepts and mechanisms. Plant Soil 300, 9e20.
Qian, L., Chen, B., 2013. Dual role of biochars as adsorbents for aluminum: the ef- Xiao, X., Chen, B., Zhu, L., 2014. Transformation, morphology, and dissolution of
fects of oxygen-containing organic components and the scattering of silicate silicon and carbon in rice straw-derived biochars under different pyrolytic
particles. Environ. Sci. Technol. 47, 8759e8768. temperatures. Environ. Sci. Technol. 48, 3411e3419.
Qian, L., Chen, B., 2014. Interactions of aluminum with biochars and oxidized bio- Xiao, X., Chen, Z., Chen, B., 2016. H/C atomic ratio as a smart linkage between py-
chars: implications for the biochar aging process. J. Agric. Food Chem. 62, rolytic temperatures, aromatic clusters and sorption properties of biochars
373e380. derived from diverse precursory materials. Sci. Rep. http://dx.doi.org/10.1038/
Qian, L., Chen, B., Hu, D., 2013. Effective alleviation of aluminum phytotoxicity by srep22644.
manure-derived biochar. Environ. Sci. Technol. 47, 2737e2745. Xu, S., Adhikari, D., Huang, R., Zhang, H., Tang, Y., Roden, E., Yang, Y., 2016. Biochar -
Qian, L., Chen, M., Chen, B., 2015. Competitive adsorption of cadmium and facilitated microbial reduction of hematite. Environ. Sci. Technol. 50,
aluminum onto fresh and oxidized biochars during aging processes. J. Soils Sed. 2389e2395.
15, 1130e1138. Xu, Y., Chen, B., 2013. Investigation of thermodynamic parameters in the pyrolysis
Qian, L.B., Chen, B.L., Chen, M.F., 2016. Novel alleviation mechanisms of aluminum conversion of biomass and manure to biochars using thermogravimetric anal-
phytotoxicity via released biosilicon from rice straw-derived biochars. Sci. Rep. ysis. Bioresour. Technol. 146, 485e493.
http://dx.doi.org/10.1038/srep29346. Xu, Y.L., Chen, B.L., 2015. Organic carbon and inorganic silicon speciation in rice-
Qin, G., Gong, D., Fan, M.-Y., 2013. Bioremediation of petroleum-contaminated soil bran-derived biochars affect its capacity to adsorb cadmium in solution.
by biostimulation amended with biochar. Int. Biodeterior. Biodegr. 85, 150e155. J. Soils Sed. 15, 60e70.
Quilliam, R.S., Glanville, H.C., Wade, S.C., Jones, D.L., 2013a. Life in the 'charosphere' Yanardag, I.H., Zornoza, R., Faz Cano, A., Yanardag, A.B., Mermut, A.R., 2015. Eval-
- does biochar in agricultural soil provide a significant habitat for uation of carbon and nitrogen dynamics in different soil types amended with
X. Zhu et al. / Environmental Pollution 227 (2017) 98e115 115

pig slurry, pig manure and its biochar by chemical and thermogravimetric soil nutrients, extracellular enzyme activities and microbial communities across
analysis. Biol. Fertil. Soils 51, 183e196. particle-size fractions in a long-term fertilizer experiment. Appl. Soil Ecol. 94,
Yang, J., Pan, B., Li, H., Liao, S., Zhang, D., Wu, M., Xing, B., 2016a. Degradation of p- 59e71.
Nitrophenol on biochars: role of persistent free radicals. Environ. Sci. Technol. Zheng, J., Chen, J., Pan, G., Liu, X., Zhang, X., Li, L., Sian, R., Cheng, K., JinweiZheng,
50, 694e700. 2016. Biochar decreased microbial metabolic quotient and shifted community
Yang, X., Liu, J., McGrouther, K., Huang, H., Lu, K., Guo, X., He, L., Lin, X., Che, L., Ye, Z., composition four years after a single incorporation in a slightly acid rice paddy
Wang, H., 2016b. Effect of biochar on the extractability of heavy metals (Cd, Cu, from southwest. China. Sci. Total Environ. 571, 206e217.
Pb, and Zn) and enzyme activity in soil. Environ. Sci. Pollut. Res. 23, 974e984. Zielinska, A., Oleszczuk, P., 2016. Bioavailability and bioaccessibility of polycyclic
Yu, L., Yuan, Y., Tang, J., Wang, Y., Zhou, S., 2015. Biochar as an electron shuttle for aromatic hydrocarbons (PAHs) in historically contaminated soils after lab in-
reductive dechlorination of pentachlorophenol by Geobacter sulfurreducens. cubation with sewage sludge-derived biochars. Chemosphere 163, 480e489.
Sci. Rep. http://dx.doi.org/10.1038/srep16221. Zimmerman, A.R., 2010. Abiotic and microbial oxidation of laboratory-produced
Yuan, H., Lu, T., Wang, Y., Chen, Y., Lei, T., 2016. Sewage sludge biochar: nutrient black carbon (biochar). Environ. Sci. Technol. 44, 1295e1301.
composition and its effect on the leaching of soil nutrients. Geoderma 267, Zimmerman, A.R., Ahn, M.-Y., 2010. Organo-mineral enzyme interactions and in-
17e23. fluence on soil enzyme activity. In: Shukla, G.C., Varma, A. (Eds.), Soil Enzymes.
Yuan, J.H., Xu, R.K., Zhang, H., 2011. The forms of alkalis in the biochar produced Springer, Berlin Heidelberg.
from crop residues at different temperatures. Bioresour. Technol. 102, Zimmerman, A.R., Gao, B., Ahn, M.-Y., 2011. Positive and negative carbon mineral-
3488e3497. ization priming effects among a variety of biochar-amended soils. Soil Biol.
Zhang, K., Chen, L., Li, Y., Brookes, P.C., Xu, J., Luo, Y., 2017. The effects of combi- Biochem. 43, 1169e1179.
nations of biochar, lime, and organic fertilizer on nitrification and nitrifiers. Biol. Zimmermann, M., Bird, M.I., Wurster, C., Saiz, G., Goodrick, I., Barta, J., Capek, P.,
Fertil. Soils 53, 77e87. Santruckova, H., Smernik, R., 2012. Rapid degradation of pyrogenic carbon.
Zhang, Q., Zhou, W., Liang, G.Q., Sun, J.W., Wang, X.B., He, P., 2015. Distribution of Glob. Change Biol. 18, 3306e3316.

You might also like