ArticlePDF Available

Abstract and Figures

Considering that even contaminated soils are a potential resource for agricultural production, it is essential to develop a set of cropping systems to allow a safe and sustainable agriculture on contaminated lands while avoiding any transfer of toxic trace elements to the food chain. In this review, three main strategies, i.e., phytoexclusion, phytostabilization, and phytoextraction, are proposed to establish cropping systems for production of edible and non-edible plants, and for extraction of elements for industrial use. For safe production of food crops, the selection of low-accumulating plants/cultivars and the application of soil amendments are of vital importance. Phytostabilization using non-food energy and fiber plants can provide additional renewable energy sources and economic benefit with minimum cost of agricultural measures. Phytoextracting trace elements (e.g., As, Cd, Ni, and Zn) using hyperaccumulator species is more suitable for slightly and moderately polluted sites, and phytomining of Ni from serpentine soils has shown a great potential to extract Ni-containing bio-ores of economic interests. We conclude that appropriate combinations of soil types, plant species/cultivars, and agronomic practices can restrict trace metal transfer to the food chain and/or extract energy and metals of industrial use and allow safe agricultural activities.
Content may be subject to copyright.
Pedosphere 22(4): 470–488, 2012
ISSN 1002-0160/CN 32-1315/P
c
2012 Soil Science Society of China
Published by Elsevier B.V. and Science Press
Designing Cropping Systems for Metal-Contaminated Sites: A Review1
TANG Ye-Tao1,2,3, DENG Teng-Hao-Bo1, WU Qi-Hang1, WANG Shi-Zhong1,2, QIU Rong-Liang1,2, WEI Ze-Bin4,
GUO Xiao-Fang4, WU Qi-Tang4,LEIMei
5, CHEN Tong-Bin5, G. ECHEVARRIA3, T. STERCKEMAN
3,
M. O. SIMONNOT6and J. L. MOREL
3,2
1School of Environmental Science and Engineering, Sun Yat-sen University, Guangzhou 510275 (China)
2Guangdong Provincial Key Laboratory of Environmental Pollution Control and Remediation Technology, Guangzhou 510275 (China)
3INPL(ENSAIA)/INRA, Laboratoire Sols et Environnement, 2 Avenue de la Forˆet de Haye, BP 172, 54505 Vandoeuvre-l`es-Nancy
Cedex (France)
4College of Natural Resource and Environment, South China Agricultural University, Guangzhou 510642 (China)
5Center for Environmental Remediation, Institute of Geographic Sciences and Natural Resources Research, Chinese Academy of
Sciences, Beijing 100101 (China)
6Laboratoire R´eactions et G´enie des Proc´ed´es (LRGP), Universit´e de Lorraine-CNRS (UPR 3349), 1 Rue Grandville BP20451, 54001
Nancy Cedex (France)
(Received February 4, 2012; revised May 10, 2012)
ABSTRACT
Considering that even contaminated soils are a potential resource for agricultural production, it is essential to develop a set of
cropping systems to allow a safe and sustainable agriculture on contaminated lands while avoiding any transfer of toxic trace elements
to the food chain. In this review, three main strategies, i.e., phytoexclusion, phytostabilization, and phytoextraction, are proposed
to establish cropping systems for production of edible and non-edible plants, and for extraction of elements for industrial use. For
safe production of food crops, the selection of low-accumulating plants/cultivars and the application of soil amendments are of vital
importance. Phytostabilization using non-food energy and fiber plants can provide additional renewable energy sources and economic
benefit with minimum cost of agricultural measures. Phytoextracting trace elements (e.g., As, Cd, Ni, and Zn) using hyperaccumulator
species is more suitable for slightly and moderately polluted sites, and phytomining of Ni from serpentine soils has shown a great
potential to extract Ni-containing bio-ores of economic interests. We conclude that appropriate combinations of soil types, plant
species/cultivars, and agronomic practices can restrict trace metal transfer to the food chain and/or extract energy and metals of
industrial use and allow safe agricultural activities.
Key Words: agronomic practices, food safety, hyperaccumulator, phytoremediation, trace element
Citation: Tang, Y. T., Deng, T. H. B., Wu, Q. H., Wang, S. Z., Qiu, R. L., Wei, Z. B., Guo, X. F., Wu, Q. T., Lei, M., Chen, T. B.,
Echevarria, G., Sterckeman, T., Simonnot, M. O. and Morel, J. L. 2012. Designing cropping systems for metal-contaminated sites: A
review. Pedosphere. 22(4): 470–488.
INTRODUCTION
The scarcity of soil resources as human population
continuously increases will inevitably force farmers to
cultivate on contaminated areas. In the long run, this
issue will be of most importance for land contami-
nated with heavy metals (e.g., Cd, Hg, Ni, Pb, and
Zn) and other trace elements (e.g., As and Se), as re-
gard to their persistence in soils. Soil contamination
is a widespread phenomenon, and soil-to-plant trans-
fer and effects of metals have been well documented
(e.g., Morel, 1997). It is now established that metals
can be phytotoxic, plants are unable to grow in some
circumstances (e.g., high metal concentration and low
pH), and the toxic metals can be transferred to edible
plant parts at rates sufficient to cause a threat to hu-
man health. Such phenomena have led governments to
ban food production in large surface areas of land con-
sidered too contaminated regarding the risks for popu-
lation health unless decontamination procedures are
previously conducted.
Hot spots of metal pollution can generally be mana-
ged and remediated by means of heavy technologies
based on excavation, treatment, and subsequent reuse
of treated materials for landscaping and soil construc-
tion (S´er´eet al., 2008; Simonnot and Croze, 2008), but
such technologies are neither economically nor environ-
mentally feasible for large surface areas. In general,
1Supported by the NSFC-Guangdong Joint Foundation of China (No. U0833004) and the National Natural Science Foundation of
China (No. 40901151).
2Corresponding author. E-mail: jean-louis.morel@ensaia.inpl-nancy.fr.
CROPPING SYSTEMS FOR METAL-CONTAMINATED SITES 471
vast sites are abandoned and under strong surveillance
regarding their possible negative impacts on human
health and environmental targets, such as water bod-
ies, air, and ecosystems. In principle, they are seldom
used for economical purposes, and even less for food
production.
However, in densely populated countries such as
China, some contaminated land is already exploited
for food production to compensate the loss of agri-
cultural surface due to urbanization. It has been re-
ported that 1/5 of agricultural land of the country is
currently contaminated (Chen, 2007), making agricul-
ture on metal-contaminated soils obligatory. Similarly,
in other industrialized countries, such as Europe (e.g.,
France), land has to be removed from traditional agri-
cultural use (i.e., food production) as a result of con-
tamination with metals (e.g., deposition from metal
smelters and mine sites and wastewater spreading over
long period of time). Moreover, with the increasing de-
mand for safer food, which will undoubtedly reduce
the quality standards, more care will be required for a
large portion of land with soil-to-plant transfer risks of
toxicants.
Therefore, there is no other alternative besides de-
veloping new agricultural scenarios suitable for such
environments, to allow safe and economically valuable
agriculture on areas that are currently considered inap-
propriate and toxic. The new strategies to be develo-
ped should be based on the designing of suitable agro-
ecological and even forest systems combining biomass
production and management of contaminants in or-
der to minimize the risks for adjacent environments,
crops, and human beings, and provide regular revenue
to the farmers. There are two main options: i) remo-
ving pollutants according to economically and techni-
cally reasonable means, applicable to large surfaces,
while using land for valuable plant production and ii)
managing contaminants, e.g., growing plants with spe-
cific methods to minimize transfer from soil to edible
plant part and avoid dispersion of contaminants in the
environment; in the latter case, designed cropping sys-
tems have to focus on production quality rather than
on soil remediation. Of course, in both options, plants
(e.g., species, variety, and agricultural practices) are
part of the challenge, and phytoremediation appears
as a relevant remediation process. Phytoremediation,
a concept which has emerged in the 1990s (Salt et al.,
1994), after having been suggested earlier (Chaney,
1983), covers a set of soil treatments based on pro-
cesses induced by plant roots on pollutants (i.e.,sta-
bilization, extraction, and degradation). Following the
study of Baker et al. (1991), which opened the way
for phytoextraction of Zn using hyperaccumulators,
much work has been conducted at the international
level demonstrating the potential of accumulating and
hyperaccumulating plants for the remediation of metal-
contaminated soils. Removal of metals by plants sig-
nificantly reduces the available pool (Chardot et al.,
2005), such as that for Ni (Echevarria et al., 2006), Cd
(G´erard et al., 2001), and Zn (Sinaj et al., 2004). Ano-
ther application of phytoremediation is stabilization
through the reduction of solubility and transport (solu-
ble, colloidal) of pollutants as a result of rhizosphere
interactions (e.g.,Morelet al., 1986) and root activity.
Phytostabilization is generally combined with amend-
ment additions to increase metal immobilization (Rut-
tens et al., 2010), and allows growth of edible plants.
However, phytoremediation is only part of the solu-
tion, and there is a great need for comprehensive agro-
nomic approaches to deal with contaminated soils. As
a remediation process, phytoremediation can only be
considered as an element to be included in sustainable
production systems adapted to contaminated environ-
ments.
Considering that even contaminated soils are a po-
tential resource for plant production, it is necessary to
propose sets of agricultural scenarios and recommenda-
tions to permit plant growth and avoid dissemination
of toxic elements, and to extract them for further eco-
nomical use. In this perspective, this paper is a contri-
bution to the future development of cropping systems
which would allow a safe and regular agricultural ac-
tivity on contaminated soils while avoiding any trans-
fer of metals to the food chain. From the knowledge
acquired in the management of contaminated land, we
have examined three main situations where agriculture
has to be conducted on metal contaminated soil: i) pro-
duction of edible plants, ii) production of non-edible
plants of industrial interests, and iii) extraction of el-
ements for industrial use. Then, we have attempted
to suggest appropriate cropping systems which would
require more scientific and technologic attention in the
near future.
CROPPING SYSTEMS FOR PRODUCTION OF
EDIBLE PLANTS
General principles and goals
There is increasing concern that cultivation of edi-
ble plants on contaminated soil can lead to the up-
take and accumulation of heavy metals in the edible
parts with their concentrations exceeding statutory or
advisory limits (e.g., legal limit set by FAO/WHO)
(Zhuang et al., 2009; Nabulo et al., 2010; Singh et al.,
472 Y. T. TANG et al.
2010; Luo et al., 2011). It is therefore imperative to
develop suitable strategies to restrict the metal accu-
mulation in such crops and its further transfer to the
food chain. The main principle is to use phytoexclu-
sion strategy, which is involved in the selection and
breeding of low-accumulating cultivars, reduction of
bioavailable metals in the soil, and restriction of their
potential uptake and translocation by plants.
Cereals
In the past decades, there have been reports that
the uptake and translocation of heavy metal pollutants
in plants varies greatly, not only among plant species
but also among cultivars within the same species
(John, 1973; Morishita et al., 1987; Wu et al., 1994,
1999; Cieslinski et al., 1996; Yu et al., 2006; Wu et al.,
2007a). Recently, there have been studies of significant
differences in the uptake of heavy metals among cul-
tivars of rice (Oryza sativa L.), maize (Zea mays L.),
wheat (Triticum aestivum L.), and soybeans (Glycine
max L.). Studies on the genotypic differences in heavy
metal concentrations in cereal crops are summarized in
Table I.
The selection of rice cultivars to minimize heavy
metal accumulation by the plant has been extensively
studied. Wu et al. (1999) evaluated 20 cultivars of rice
in South China on a contaminated soil and their re-
sults showed that the grain Cd concentrations ranged
from 0.48 to 1.17 mg kg1. The main reason causing
the differences was that the low-accumulating cultivar
exhibited a much lower Cd translocation rate to grains
than the high accumulating cultivar. In recent years,
most studies have been conducted on the molecular
mechanism of the low-accumulating rice cultivar. A
novel mapping population consisting of 39 chromosome
segment substitution lines (CSSLs) was used to locate
the putative quantitative trait loci (QTLs) for Cd con-
centration in brown rice (Ishikawa et al., 2005a). Xue
et al. (2009) found that 3 QTLs on chromosomes 6 and
7 controlled root and shoot Cd concentrations. Ueno
et al. (2010) reported that the gene (OsHMA3) was
responsible for low Cd accumulation in rice.
Heavy metal concentrations have also been shown
to vary among wheat cultivars (Oliver et al., 1995;
Cieslinski et al., 1996; Zhang et al., 2002a; Gao et
al., 2011). Erdman and Moul (1982) evaluated 7 cul-
tivars of durum wheat and reported that their grain
Cd, Cu, and Zn concentrations ranged from 0.13 to
0.25 mg kg1with an average of 0.22 mg kg1.Gaoet
al. (2011) also found that the wheat cultivars showed
significant genotypic differences in grain concentrations
of Cd and Zn. Greger and Lofstedt (2004) observed
that differences among the cultivars in the ability to ac-
cumulate Cd in grain were related to variations in the
translocation from root to shoot and within the shoot,
rather than to differences in root uptake. Similar trends
were observed by Harris and Taylor (2004). Clarke et
al. (1997) found that grain Cd concentrations were
largely controlled by a single gene. The low-Cd wheat
cultivar Strongfield (Clarke et al., 2005) was released in
2004 and is now sown on more than 25% of the durum
area in Canada (Grant et al., 2008).
Guo et al. (2010) observed that the grain Cd, Pb,
Zn, and Cu concentrations in sweet corn cultivars were
higher than those of normal corn cultivars. Analysis of
17 varieties of soybean showed that the soybean variety
En-b0-1-2 was the low-Cd cultivar in both field and pot
experiments (Arao et al., 2003b). Low-accumulating
maize and soybean cultivars appeared to retain more
heavy metals in the root and translocate less to the
grain than high-accumulating ones (Arao et al., 2003b;
Ishikawa et al., 2005b; Guo et al., 2010).
Vegetables
Similar to cereal crops, vegetables have a variety
of capacities in accumulating heavy metals among dif-
ferent types of species. The leafy vegetables tend to
accumulate higher amounts of heavy metals in edible
TAB L E I
Studies on genotypic differences in heavy metal concentrations in cereal crops
Crop References
Rice (Oryza sativa L.) Morishita et al., 1987; Wu et al., 1994, 1999; Arao and Ae, 2003a; Liu et al., 2003b, 2005a, 2007a;
Li et al., 2005b; Yu et al., 2006, 2008; Fu et al., 2008; Shi et al., 2009
Wheat (Triticum aestivum L.) Erdman and Moul, 1982; Oliver et al., 1995; Cieslinski et al., 1996; Clarke et al., 1997; Zhang et
al., 2000, 2002a; Greger and Lofstedt, 2004; Harris and Taylor, 2004; Clarke et al., 2005; Gao et al.,
2011
Maize (Zea may L.) Florijn and Beusichem, 1993; Gao et al., 1995; Kurz et al., 1999; Samake et al., 2003; Tudoreanu
and Phillips, 2004; Guo et al., 2010
Soybean (Glycine max L.) Reddy and Dunn, 1986; Arao et al., 2003b; Ishikawa et al., 2005b
CROPPING SYSTEMS FOR METAL-CONTAMINATED SITES 473
parts than fruit or even root vegetables (Zheng et al.,
2007; Wang et al., 2009a; Zhuang et al., 2009; Khan
et al., 2010), although this is not always the case. The
high metal accumulation in leafy vegetables is regarded
as a result of their high rates of transpiration and
translocation of metals, particularly Cd (Khan et al.,
2010). In addition, leafy vegetables are fairly suscepti-
ble to physical contamination by soil dust and splash
because of their high foliar surface area (Zhuang et
al., 2009). Accumulation of Pb and Cr in leafy vege-
tables resulting from aerial deposition is substantially
greater than that of Cd, Zn, Ni, and Cu, of which the
uptake occurs predominantly via the roots and subse-
quent transport to the edible shoot tissues. This high-
lights the importance of mulching vegetable gardens to
reduce exogenous metal contaminants on leaves (Na-
bulo et al., 2010).
Heavy metal accumulation may also differ greatly
within cultivars of an individual species when grown
on the same soil. Wide variation of metal accumulation
within cultivars has been documented in potato (Dun-
bar et al., 2003), carrot, pea (Alexander et al., 2006),
asparagus bean (Zhu et al., 2007), and several leafy
vegetables (Wang et al., 2007b, 2009a). Alexander et
al. (2006) tested five cultivars of six common vegetables
and found that carrot and pea cultivars exhibited sig-
nificant intra-species differences in Cd and Cu. Wang et
al. (2009a) observed that there were large genetic vari-
ances in water spinach Cd concentration, ranging from
0.593 to 1.824 mg kg1among 30 tested cultivars when
grown on Cd-treated soil. Among them, six cultivars
were identified as pollution-safe cultivars and the pre-
ferential sequestration of Cd in root and cell wall frag-
ment was one possible mechanism leading to low Cd ac-
cumulation. Using suppression subtractive hybridiza-
tion analysis, Huang et al. (2009) found tremendous
differences in Cd-induced gene expression in two water
spinach cultivars with contrasting Cd-accumulating ca-
pacity, suggesting that Cd-accumulating ability of this
species is a genotype-dependent and inherited trait.
Orchards
One of the most common elevated trace metal con-
taminants in fruit orchard soils is Cu. The long-term
use of Cu-based fungicides has resulted in elevated ac-
cumulation of Cu in soils of vineyard and apple or-
chard, typically ranging from 200 to 500 mg kg1
(Chopin et al., 2008). The values are over ten times
higher than the typical background Cu concentrations
in soils (5–30 mg kg1) (Wightwick et al., 2010).
There are increasing concerns of Cu availability
in orchard soils and its subsequent accumulation in
grapevines (Mirlean et al., 2005; Lai et al., 2010). The
adverse effects of Cu accumulation in apple orchard
and vineyard soils on enzymatic activity, microbial
community, and soil invertebrates have also been em-
phasized (Wang et al., 2009b; Zhou et al., 2011). In
some vineyard sites of central Taiwan, high accumu-
lation of Cu up to 86 mg kg1in grape berries was
reported, far exceeding the maximum residue level of
Cu in wine grapes (20 mg kg1) proposed by the Eu-
ropean Regulation (Lai et al., 2010). In contrast, much
lower Cu content (4.5 mg kg1) was observed in berries
grown in a calcareous contaminated Champagne vine-
yard area, France (Chopin et al., 2008). In fact, many
studies have demonstrated that the availability of Cu is
dominated by soil factors such as pH, cation exchange
capacity (CEC), clay content, and organic matter con-
tent. Cu availability in soil is generally lowest in neu-
tral soils and increases with increasing soil acidity and
alkalinity (Wightwick et al., 2010). This may have im-
portant implications in considering strategies (e.g., soil
pH management) to reduce Cd availability and toxi-
city in vineyards. On the other hand, several studies
suggested that organic carbon is responsible for a sub-
stantial retention of Cu mainly by increasing CEC, and
thus application of compost is another feasible option
to reduce Cu mobility in vineyard soils (Lai et al., 2010;
Wightwick et al., 2010). From the long run, it is wise
to restrict or even stop the use of Cu-based fungicides,
and develop environmentally safer alternatives to Cu-
based fungicides, which have been banned in some Eu-
ropean countries like the Netherlands (Lai et al., 2010).
Chemical immobilization of heavy metals in soil
In-situ chemical immobilization is a cost-effective
way to reduce metal concentrations in plants through
the addition of chemical amendments in soil, which
may provide a solution through the formation of low
solubility and/or precipitation of minerals (Basta and
McGowen, 2004). Kumpiene et al. (2008) reviewed the
published papers during 2000–2005 on immobilization
of As, Cr, Cu, Pb, and Zn in soils.
Research on chemical immobilization of heavy me-
tals has included alkaline substances, phosphate- and
silicon-based materials, adsorption agents, and organic
matters. Alkaline substances, such as lime, have been
used in traditional agricultural practice to raise soil pH
levels for plant growth (Simon, 2005). The addition of
alkaline amendments reduced heavy metal solubility
in soil by increasing soil pH and formed insoluble com-
pounds (Maddox and Soileau, 1991; Chen et al., 2000;
Khan and Jones, 2009). One of the disadvantages of
liming is that the gradual reduction of the liming agent
474 Y. T. TANG et al.
over time due to dissolution and leaching, which can
be accelerated by acid deposition, may have a poten-
tial risk that needed to be considered prior to long-term
phytoexclusion practices (Ruttens et al., 2010).
Silicon (Si) is not considered to be an essential
mineral element for plants, but has been shown to
ameliorate the deleterious effects of metals on plants
grown on contaminated soils (Liang et al., 2005, 2007;
da Cunha and do Nascimento, 2009). Silicon applied
in agriculture came mainly from by-products of fly-ash
(Dwivedi et al., 2007), steel sludge, furnace slag (Chen
et al., 2000), and Si chemicals (Shi et al., 2005; Li et
al., 2009c; Gu et al., 2011). Liu et al. (2009) reported
that foliar application of silica sols significantly in-
creased the dry weight of rice grains and shoots grown
on Cd-contaminated soils, whereas the Cd concentra-
tion apparently decreased in the grains and shoots. The
mechanism of Si foliar application to alleviate the to-
xicity and accumulation of Cd in rice grains may be
related to the co-precipitation of Cd with Si in stems
(Gu et al., 2011), and probable Cd sequestration in the
shoot cell walls (Liu et al., 2009).
Phosphate-based materials added to contamina-
ted soils reduced the Pb mobility by ionic excha-
nge and precipitation of pyromorphite-type minerals,
Pb5(PO4)3X, where X = F, Cl, or OH (Kumpiene
et al., 2008). The phosphate-based materials include
phosphate rock (Zwonitzer et al., 2003; Cao et al.,
2004; Brown et al., 2005), diammonium phosphate
(Basta and McGowen, 2004; Chen et al., 2007; Khan
and Jones, 2009), phosphate-based salts (Ownby et al.,
2005; Simon, 2005; Li et al., 2008), and phosphoric acid
(Cao et al., 2003).
Adsorption agents such as zeolites, iron and/or
manganese oxides, and clay minerals have also been
widely studied because of their high cation exchange
capacity and high specific surface. In a pot experi-
ment, natural zeolites decreased the available part
of Pb and restricted Pb uptake by rape (Li et al.,
2009a). Among 15 metal ions investigated, addition of
vermiculite significantly reduced the uptake of various
metal pollutants by lettuce and spinach (Malandrino
et al., 2011). A similar reduction effect on Cd and Zn
accumulation in radish and Japanese mustard spinach
wasreportedwhen5%lherzolitewasappliedinacon-
taminated soil (Kashem et al., 2010).
Organic matters such as peat, animal excrement,
and composts can also effectively be used in metal sta-
bilization. Li et al. (2008) reported that pig manure
and peat could increase grain yields and decrease Cu
and Cd concentrations in grain.
CROPPING SYSTEMS FOR PRODUCTION OF
NON-EDIBLE PLANTS
Energy crops
The technology development and utilization of phy-
toremediation using tolerant energy crops can pro-
vide additional economic benefits (Meers et al., 2005,
2010). These crop cultivars are cultivated on metal-
contaminated soils for energy production purposes,
and the sustainable development of resilient energy
plants has turned into a new trend in the developed
and developing countries.
Industrial non-food crops have been used for both
remediation of contaminated land and renewable ene-
rgy sources (Meers et al., 2007). Canola (Brassica na-
pus var. Hyola) plants grown under adverse soil condi-
tions in the west side of the San Joaquin Valley (SJV)
accumulated Se in various plant organs including seeds.
The Se-enriched economically viable phyto-products of
canola such as seed oil have been used as biofuel ad-
ditive and organic fertilizer (Ba˜nuelos, 2006; Ba˜nuelos
et al., 2010).
A number of studies have indicated that many ene-
rgy plants (e.g., sunflower, willow, and poplar) have
high tolerance to heavy metals such as Cu, Cd, Cr, and
Zn (Lin et al., 2003; Lun´ckov´aet al., 2003; Rosselli
et al., 2003). In addition, soybean (Glycine max) oil,
palm (Trachycarpus fortunei) oil, and coconut (Cocos
nucifera) oil have been considered as possible substi-
tutes for fossil diesel (Agarwal, 2007). Hence, metal-
tolerant industrial and energy crops are suitable can-
didates for long-term field remediation and sustainable
economic development. This process could be of par-
ticular importance in countries with increasingly less
arable land, such as China and India, due to the rapid
expansion of population and urbanization.
Jatropha curcas L. (Euphorbiaceae) is a perennial
oil-yielding plant cultivated in the tropical and sub-
tropical regions of the world. It has been widely used
for energy plantations and wasteland reclamation due
to the ability to tolerate extreme environmental condi-
tions such as drought, non-fertile, and heavily polluted
soils. Based on the accumulation of different trace ele-
ments (Cu, Cr, Mn, Ni, and Zn) and nutrients (Ca,
K, Mg, Na, and P) in J. curcas grown in the vicinity
of a mica belt, it could be used for phytostabilization
of mining areas (Nagaraju and Karimnlla, 2002). Re-
cent studies have mainly focused on the improvement
of growth performance and metal tolerance of J. curcas
grown on metal-contaminated calcareous soils with the
amendments of organic wastes (biosludge and dairy
CROPPING SYSTEMS FOR METAL-CONTAMINATED SITES 475
sludge) and a biofertilizer (Azotobacter chroococcum)
(Juwarkar et al., 2008; Kumar et al., 2008; Yadav et
al., 2009, 2010).
J. curcas can also grow well on the acid mine
drainage irrigated soils (pH 3.78) with addition of
1% limestone in extremely acidic soils (Li et al., 2009d;
Wu et al., 2011). The growth performance of J. cur-
cas was enhanced at 0.50% of lime addition due to the
increase of soil pH and the decrease of phytoavailable
Al (about 95%), Zn (about 75%), and Cu (about 65%)
in soil. The high ability to tolerate Al was suggested
to relate to cell wall binding and the organic acids se-
creted by the roots of J. curcas (Wu et al., 2011). Based
on the inherent tolerance ability of J. curcas in exis-
ting adverse environmental conditions, it could be used
as a suitable candidate for phytostabilization in acid
mine tailings. Effective stabilization of liming is an ad-
ditional possible practice to improve the soil quality.
Fiber crops
Bast fiber crops are widely grown in China and
throughout the world. The characteristics of fiber crops
include fast growth, strong adaptability, large biomass,
and high economic value. The bast fibers are widely
applied in textiles, adsorbents, and building materi-
als (Liu et al., 2003a; Hasfalina et al., 2010). These
non-wood fibers were also proved to be suitable for pa-
permaking (Saikia et al., 1997; Ververis et al., 2004).
Recent attention has been given to fiber crops as a
potential plant for heavy metal-contaminated soils due
to its tolerance (Angelova et al., 2004; Ho et al., 2008;
Yang et al., 2010). Hemp (Cannabis sativa) could grow
under 17 mg kg1Cd in soil and 1 mmol L1Cu in cul-
ture solution without showing any phytotoxicity symp-
toms (Citterio et al., 2003; Arru et al., 2004). Through
a large-scale field investigation and a hydroponic cul-
ture experiment, ramie (Boehmeria nivea) has been
revealed to possess a certain degree of constitutional
metal tolerance to As, Cd, Pb, and Zn (Yang et al.,
2010).
However, due to the presence of toxic metals, high
soil acidity, and poor physical properties with defi-
ciency of nutrients in the vicinity of mine tailings (Pich-
tel and Salt, 1998; Ernst, 2005), fiber crops are hard to
grow in these contaminated areas. Therefore, revege-
tation process should combine with application of soil
amendments, especially those that can neutralize soil
pH and enhance nutrition. A 3-year field experiment on
a stripmine land showed that the yield of kenaf (Hibis-
cus cannabinus) increased with sewage sludge applica-
tion (Carlson et al., 1982). Similar results were also
found in a pot trial by applying chicken manure to a
Pb-contaminated soil (Ho et al., 2008).
In another field trial conducted in Shangba Vil-
lage, Guangdong, China, where the farmland soil was
highly polluted by acid mine drainage from upstream
mine tailings, the ameliorative effects of lime, dolomite,
fly ash, and compost were tested (Table II) (Huang,
2009). The contaminated soil was characterized by its
low pH (3.36) and high concentrations of heavy me-
tals (1 500, 10, 660, and 480 mg kg1of Pb, Cd, Cu,
and Zn, respectively) (unpublished data). The soil pH
was adjusted to 7–8 by amendment application. The
yield of kenaf shoots could reach 20 t ha1or above
(Fig. 1) (Zou, 2008), which is similar to those on nor-
mal soils, indicating a successful and sustainable mode
of reclaiming the acid mine soils by using kenaf in com-
bination with soil amelioration. Besides, low levels of
heavy metals were detected in kenaf fibers. Meanwhile,
it was suggested that heavy metals had no significant
negative effects on fiber quantity and quality (Linger
et al., 2002; Wang et al., 2008). A highly contaminated
soil is always considered to be a wasteland. Howe-
ver, with a certain amendment applied into the soil,
fiber crops could grow healthily and rapidly. There-
fore, these crops are good candidates to cultivate on
highly contaminated soils for revegetation as well as
economic purposes.
Though the heavy metals in fibers may not easily
enter the food chain and do harm to human beings di-
TAB L E I I
Effects of different amendment treatments on yields and agricultural and economic characteristics of kenaf (Huang, 2009)
Treatment Agronomic trait Biomass
Yield Stem diameter Stem height Bast fiber Stem core
tha
1cm gplant
1
Lime (0.6%) 22.0aba) 4.4a 308a 24.7a 63.0a
Dolomite (5%) 40.5b 6.2b 374a 47.9b 129.0b
Fly ash (5%) 30.7ab 5.9ab 358a 37.1ab 93.3ab
Lime (0.6%) + compost (5%) 35.0ab 5.4ab 356a 40.2ab 110.0ab
Lime (0.6%) + fly ash (5%) 20.1a 5.6ab 351a 36.9ab 94.4ab
a)Means followed by the same letter(s) in a column are not significantly different at P<0.05.
476 Y. T. TANG et al.
Fig. 1 A field trial of acid mine contaminated soil amendment in Shangba Village, Guangdong, China: (a) before growth of kenaf
and (b) after growth of kenaf (Zou, 2008).
rectly, a minimum amount of metals is expected in
these industrial raw materials. So, more attention
should be paid to amendment selection and soil mana-
gement practices to achieve the best crop yield and
least potential hazard.
CROPPING SYSTEMS FOR EXTRACTION OF
ELEMENTS FOR INDUSTRIAL USE
Phytoextraction of Cd, Ni and Zn: perspectives for phy-
tomining and soil remediation
General principles and goals. Nickel is the most
accumulated metal among hyperaccumulator plants.
Approximately 390 plant species have been reported
worldwide (Verbruggen et al., 2009) to hyperaccumu-
late Ni in their shoot tissues (i.e., Ni concentrations of
above 1 000 mg kg1) and most of them easily reach 1%
Ni in their leaf tissues when growing on soils with suffi-
cient Ni availability (Massoura et al., 2004; Chardot et
al., 2005). Hyperaccumulation of Cd and Zn occurs to
be less than that of Ni (Peer et al., 2006; Verbruggen
et al., 2009). Also, only a few plant species are able
to take up both metals at high levels; these plants be-
long mainly to the family Brassicaceae (several species
of the Thlaspi/Noccaea genus and Arabidopsis hallerii)
as well as the species Sedum alfredii (Yang et al.,
2004). The co-hyperaccumulation is highly important
because most of the Cd-contaminated sites also display
high concentrations of Zn as these two metals are of-
ten present together in mineral ores and in industrial
wastes and by-products. Metal-rich soils and mainly
ultramafic soils (i.e., serpentine soils) are widespread
over different areas of the world including the Alps,
the Balkans, Turkey, New Caledonia, Cuba, Australia,
and many other regions. These soils are a major prob-
lem for the development of agriculture due to extreme
nutrient deficiency and toxicity of trace metals. The re-
covery of such metals from soils with hyperaccumulator
plant crops as an alternative to traditional agriculture
is a challenge for the rural development of these areas
as well as for the local mining and metallurgical indus-
try (Bani et al., 2007). Very often, ultramafic soils are
too poor in metal contents for viable traditional mining
operations; phytomining, i.e., recovery of metal hyper-
accumulated in aboveground plant parts, proves to be
profitable in these soils and becomes a true alterna-
tive to classical crops (Bani et al., 2007; Chaney et al.,
2007). Also, the prices of some metals (e.g., Ni, Co, and
Tl) have dramatically increased for the past five years
although a noticeable breakdown occurred during the
2009 crisis. During 2010, the prices of metals went up
again due to increasing demand of Asian industries; at
the time this text was written, the price of Ni at the
London Metal Exchange was USD $ 26 per kg. This
alone justifies their mining from naturally rich soils or
their recycling when remediating metal-contaminated
soils. Industrial brownfield and mining sites are also of-
ten contaminated with trace metals (Cd, Co, Cr, Cu,
Ni, and Zn), representing a major environmental con-
cern (De Kimpe and Morel, 2000). In the case of Cd
and Zn, the main objective of remediation is often re-
moval of the metal fractions that present a risk in co-
CROPPING SYSTEMS FOR METAL-CONTAMINATED SITES 477
ntaminated soils. Removal of total contents of metals in
contaminated soils is often impossible thermodynami-
cally speaking but has long been considered as an is-
sue by stakeholders (Schwartz et al., 2003; Keller and
Hammer, 2004). Now, it is commonly accepted that
the main goal of phytoextraction of Cd and Zn is to
deal with keeping low the bioavailable fractions of these
metals to prevent excessive uptake in following crops
or leaching to groundwater. However, the recovery of
metals in the produced biomass could be a way of re-
ducing the overall costs of contaminated soil treatment,
especially considering the expected increases in Cd and
Zn market values in the forthcoming years.
Phytomining of Ni: agronomic perspectives. Se-
veral field trials and research programs have led to
a better understanding of phytomining and proposed
several sets of agronomic practices to end-users (Li et
al., 2003; Bani et al., 2007; Chaney et al., 2007). It
is quite easy to grow Ni-hyperaccumulator plants in
pot conditions on a wide array of soils and treatments
(Robinson et al., 1999) and significant amounts of Ni
can be recovered after three months with Ni concentra-
tions in shoots reaching more than 3% and depletion of
total soil Ni reaching almost 5% (Kukier et al., 2004;
Massoura et al., 2004; Chardot et al., 2005). However,
field-scale phytomining requires approaches that inte-
grated species selection and breeding, phenology and
cropping calendar, fertilization practices, and pest con-
trol (Chaney et al., 2007). Two approaches have been
developed so far: one intensively designed that includes
all aspects of mechanized agriculture available in deve-
loped countries (Li et al., 2003; Chaney et al., 2007)
and the other extensively designed that is adapted
to ultramafic areas in the world where no modern
cropping tools are available (Bani et al., 2007; Bani,
2009). The latter needs the same scientific basic know-
ledge to optimize agronomic options. Some studies
on the phenology of Alyssum murale showed that the
curves of Ca, Mg, and Ni uptake during plant growth
and development defined the mid-flowering stage as the
best stage for harvesting with the highest Ni concen-
trations and the lowest Ca concentrations in shoots
(Bani, 2009). Chaney’s group led a breeding program
with various species of Alyssum (Li et al., 2003), among
which A. murale from eastern Albania proved to be
one of the best candidates (Bani et al., 2009). The se-
lected strains were able to accumulate 3% Ni in the
shoots (including stems) as compared to 2% as a best
value in the wild (Bani et al., 2009), with reasonably
increased yields (up to 20 t ha1in field conditions)
and a better retention of leaves that are the richest
parts in the plant (Chaney et al., 2007). Optimization
of soil fertility concerns mainly nitrogen (Bani et al.,
2007). A. murale is very efficient in taking up phospho-
rus (P) and soil P fertility is seldom a limiting factor
(Shallari et al., 2001; Bani et al., 2007). However, N
fertilization is seriously required and it was possible
to limit N inputs to 100 kg N ha1by fractioning its
application (Chaney et al., 2007; Bani, 2009). Since
most Ni hyperaccumulators take up large amounts of
Ca from Ca-deficient serpentine soils, it is necessary to
maintain Ca fertility of such soils at a reasonable level
(Kukier et al., 2004; Chaney et al., 2007).
Phytoextraction of Cd and Zn with hyperaccumula-
tor plants at the field scale. Several teams in the
past twenty years have carried out experiments on
Cd and Zn phytoextraction by hyperaccumulators at
the field scale and confirmed what had been studied
at the lab scale although the adaptation of species
such as Noccaea caerulescens to the field is not so
easy (Schwartz et al., 2003; Keller and Hammer, 2004;
Chaney et al., 2005, 2007). To date, most studies have
shown that although N. caerulescens can be used to
clean up soils slightly or moderately contaminated by
Cd and Zn (Ernst, 2005; McGrath et al., 2006), the
technology usually needs a long period (e.g.,10years)
for cleanup due to the limited biomass production
of hyperaccumulator species used (Koopmans et al.,
2008). Also, the phytoextraction duration can be un-
derestimated since the bioavailable metal pools in soil
decrease during phytoextraction, which leads to a de-
crease of metal uptake by plants and lower metal re-
moval rates (Koopmans et al., 2008). Thus, phytoex-
traction appears possible for Ni and As, but for Cd and
Zn, the strategy is still far from large-scale practice
(Van Nevel et al., 2007). One of the keys for successful
phytoextraction operations based on N. caerulescens
requires an increased biomass production.
Metal recovery in hyperaccumulator biomass. Ni
is used in a wide variety of alloys and superalloys but
the largest needs concern the production of stainless
steel. For this reason, there is a wide variety of pro-
cesses for Ni recovery that mainly depend on the type
of and the concentrations of Ni in ores. Currently,
about twenty processes are known, among which 6
are significantly used at the industrial scale, e.g.,
pyrometallurgy (INCO and Falconbridge, Canada) and
hydrometallurgy (Eramet-SLN, France). Some scien-
tists have shown that pyrometallurgical processes can
be used to produce pure Ni from hyperaccumulator
plant residues (Boominathan et al., 2004). This process
is however difficult to promote in developed countries
because of the strict environmental regulations. Other
scientists promote co-incineration of hyperaccumula-
478 Y. T. TANG et al.
tor plants along with municipal wastes (Keller et al.,
2005). This also causes problems in terms of social ac-
ceptability and difficulty to recover metals that may
escape with fly ashes. These ashes would then need a
special treatment as dangerous wastes. One alterna-
tive to pyrometallurgy to valorize phytoextracted me-
tals could be a hydrometallurgical process that better
responds to clean technology standards (Barbaroux et
al., 2009, 2011, 2012). Hydrometallurgy consists in a
set of techniques used for a long time (Habashi et al.,
2005). The yearly world production of Ni (over one
million tons) is mainly based on pyrometallurgy but
hydrometallurgy is also used because of the possibili-
ties it offers for the elaboration of end products and for
the recovery of metals in complex systems. Therefore,
it is essential to continue development and adaptation
of hydrometallurgical methods, in particular lixiviation
methods, lixiviate purification, and production routes
of end products, in order to valorize metals present in
various environmental materials. Biomass produced by
phytomining is a substantial resource (i.e., 2% Ni), for
which a specific separation process should be develo-
ped.
Phytoextraction of As
Several plant species have been reported as As hy-
peraccumulators, primarily represented by terrestrial
Pteris species (Ma et al., 2001; Chen et al., 2002b;
Wang et al., 2007a) and aquatic seaweeds (Nguyen et
al., 2011). Some angiosperms, e.g., Brassicaceae Isatis
cappadocica (Karimi et al., 2009) and Hesperis per-
sica (Karimi et al., 2010), and ferns, e.g., marsh fern
(Anderson et al., 2010), have also been reported as
As hyperaccumulators. Pteris vittata has the merits of
large biomass, high As concentration, and wide distri-
bution; thus, it has been widely used in phytoextrac-
tion projects (Chen et al., 2007; Ebbs et al., 2009).
P. vittata has a wide distribution. Diverse habi-
tats result in different genotypes. Although all ecotypes
of P. vittata could hyperaccumulate As, the ability of
hyperaccumulating As was very variable. The above-
ground As concentration of P. vittata ranged from 643
to 3 009 mg kg1(Cai et al., 2007). It means that using
different genotypes in phytoremediation projects could
lead to several times of variations in the remediation
efficiency. Therefore, differences in As accumulation by
P. vittata genotypes were investigated and the 3–4 fold
As accumulation differences were confirmed (Wang et
al., 2002; Zhao et al., 2002; Wu et al., 2009). It was
found that some P. vittata ecotypes also had huge po-
tentials in phytoextracting other toxic trace elements,
e.g., Zn, Cd, Cr, and Hg (Anet al., 2006; Su et al.,
2008b; Xiao et al., 2008; Roccotiello et al., 2010; Feng
et al., 2011; Kalve et al., 2011). The reasons behind
such ecotypic variations have not been identified yet,
but these research results clearly showed the necessity
to select appropriate genotypes according to the situ-
ation of contaminated fields.
Besides plant selection, complementary approaches
to optimize plant growth and increase As concentration
in P. vittata have been adopted. The uses of fertilizers
(phosphate and nitrogen) and mycorrhizal fungi have
received focused attention and both were able to in-
crease the remediation efficiency (Chen et al., 2002a;
Liu et al., 2005c; Wei and Chen, 2006; Huang et al.,
2007; Liao et al., 2007; Leung et al., 2010; Muthuku-
mar and Bagyaraj, 2010; Xiong et al., 2010). The ef-
fect of phosphate fertilizers is still unclear (Huang et
al., 2007; Santos et al., 2010).
In addition, the optimization of environmental fac-
tors including soil properties, seedling age, and har-
vesting regimes has also been reported (Gonzaga et
al., 2007; Wan et al., 2010; Xu et al., 2010; Natarajan
et al., 2011). Younger seedlings have been found to ex-
tract As more quickly, but whether we should adopt
young plants or apply more frequent harvest regimes
still depends on the balance between larger biomass
and higher concentrations.
A field experiment conducted showed that P. vit-
tata can achieve maximum extraction efficiency by
mowing 3 times a year, and the stubble height had
a great impact on the P. vittata growth and extraction
efficiency: a stubble height of 7.5 cm benefited the re-
generation of P. vittata and the extraction efficiency of
arsenic (Li et al., 2005a).
In the plant tissues of arsenic hyperaccumulators
(Pteris species), inorganic arsenic species arsenate
(As(V)) and arsenite (As(III)) account for more than
95% of the total arsenic, while organic species have
been detected as minor compounds (Ma et al., 2001;
Tu et al., 2002, 2004; Zhang et al., 2002b; Mathews et
al., 2010). Regardless of As(III) or As(V) supplied, As
exists predominantly as As(V)–O in roots while mainly
as As(III)–O in the rhizomes and fronds (Huang et al.,
2008; Mathews et al., 2010). Arsenite was firstly oxi-
dized in the growing medium and then further oxidized
in the roots of P. vittata (Huang et al., 2004; Mat-
hews et al., 2010). Arsenate was mainly reduced in the
rhizomes (68%–71% as As(III)) and pinnae (>90%
as As(III)), and limitedly reduced in the roots (7%–
8% as As(III)) (Huang et al., 2004; Mathews et al.,
2010). Monomethylarsonic acid (MMA) can be methy-
lated to dimethylarsinic acid (DMA), while DMA can-
not be demethylated to MMA. Both MMA and DMA
CROPPING SYSTEMS FOR METAL-CONTAMINATED SITES 479
can be oxidized to arsenate (As(V)) and arsenite
(As(III)) (Chen et al., 2004). Arsenic transformation
mechanism has not yet been clearly explored and still
leaves much to be resolved.
Phytoextraction faces the challenge of safe dis-
posal of large quantities of contaminant-rich biomass
after harvesting from contaminated sites (Cao et al.,
2010). Incineration was considered to be suitable for
hyperaccumulators after harvest. During incineration
of P. vittata containing high As content (1 170 mg
kg1), about 24% of total As was emitted at 800 C,
and 62.5% of the total emitted As was volatilized below
400 C. Simultaneously, below 400 C, As(V) reduced
to As(III) (Yan et al., 2008). The ash weight was about
4.4%–7.7% of the dried weight and the concentration of
heavy metals presented regular variations in different
stages of incineration. As concentration in the bottom
ash was lower than that in the fly ash.
CO-CROPPING FOR REDUCING HEAVY METAL
CONTENTS IN CROPS
Co-cropping of a hyperaccumulator (N. caerule-
scens) with a non-accumulator (Thlaspi arvense)in-
creased the growth of the non-accumulator and re-
duced its Zn uptake, while in contrast, increased
Zn uptake of the hyperaccumulator (Whiting et al.,
2001). Comparing with monoculture, intercropping de-
creased the Cd content in the aboveground parts of al-
falfa (a normal crop) by 57.1%, while increased the Cd
content in the aboveground parts of Brassica juncea (a
high Cd accumulator) by 14.5% for the soils with Cd
contents in the range of 0.37–5.37 mg kg1(Li et al.,
2009e). The Cd contents in the aboveground parts of
alfalfa under monoculture and intercropping were 0.21
and 0.09 mg kg1, respectively, neither of them being
over the feed safety standard (Li et al., 2009e). High
Cd-accumulating varieties of oilseed rape could reduce
the Cd uptake of co-cropped cabbage, but the Cd con-
centration in cabbage was still high (Liu et al., 2007b;
Su et al., 2008a). Therefore, a low-metal crop is needed
to produce safe agricultural products on contaminated
agricultural lands in association with a hyperaccumu-
lator. Growing a metal hyperaccumulator plant S. al-
fredii alongside a low metal-accumulating cultivar of
maize was introduced to simultaneously phytoextract
heavy metals from contaminated sewage sludge and to
produce an agricultural crop (Liu et al., 2005b). Re-
sults showed that this co-planting system effectively
removed heavy metals from the sewage sludge by the
hyperaccumulator while the harvested corn grains or
Alocasia tubers were conformed to the Chinese stan-
dard for animal feeds (Liu et al., 2005a; Wu et al.,
2007b).
However, the beneficial effect on the co-cropped
normal crop was not always observed from the co-
cropping with a hyperaccumulator. Combined crop-
ping of N. caerulescens and Hordeum vulgare, though
caused a decrease in Zn concentration in barley bio-
mass, had a significant increase in Cd accumula-
tion (Gove et al., 2002). Similar results were also
obtained by Jiang et al. (2010). The average Cd
concentration in ryegrass shoots was increased by
co-cropped N. caerulescens, but the Zn concentra-
tion and total uptake by ryegrass were decreased
by co-cropping with N. caerulescens compared with
mono-cropping or co-cropping with T. arvense (non-
hyperaccumulator). These observations suggest that
the hyperaccumulator competes for Zn against the co-
planted crop, but not always true for Cd, probably be-
cause the hyperaccumulator does not need Cd for the
physiological functioning of the plant and may mobilize
too much Cd which may be absorbed by the co-planted
crop.
Co-cropping of two normal crops may also alter the
availability of metals. Wheat/rice intercropping could
reduce Cd uptake by wheat and rice in shoots and re-
duce Cd concentration in the wheat grain; however, Cd
concentration in rice grain was increased, compare to
that in the monocultured plant (Wu et al., 2003). In
the intercropping system of Cunninghamia lanceolata
in tea gardens, the concentrations of Pb, Mn, Cu, and
Zn in the tea leaves were significantly lower than those
in the mono-cropping one, indicating that intercrop-
ping C. lanceolata in tea gardens can improve the quali-
ty of the tea (Xue and Fei, 2006).
Continued agricultural production of safe animal
feeds or energy crops on contaminated soils is necessary
considering the limited arable land in China and time
required for phytoremediating metal-contaminated
soils. Co-cropping or intercropping may alter condi-
tions in shared rhizosphere and thereby affect the avail-
ability of selected metals to neighboring plants. There-
fore, it is possible that planting some low-metal crops
in association with hyperaccumulators or other ap-
propriate plants may allow agricultural production on
heavy metal-contaminated soils.
CO-CROPPING FOR ENHANCING PHYTOEX-
TRACTION
Co-cropping N. caerulescens with T. arvense in-
creased Zn uptake by the hyperaccumulator (Whiting
et al., 2001). Combined cropping of N. caerulescens
and Hordeum vulgare caused a higher concentration
of heavy metals (Cd, Pb, and Zn) in the biomass of
480 Y. T. TANG et al.
the hyperaccumulator N. caerulescens (Gove et al.,
2002). It may indicate a potential approach to increase
the remediation efficiency of metal-contaminated soils
in a natural way.
A co-planting system, comprised of a Zn and Cd
hyperaccumulator (S. alfredii) and a low-accumulating
crop (Z. mays), was established in a rice field that was
historically irrigated with water affected by Pb and Zn
mining activities (Wu et al., 2007c). The co-planting
system enhanced not only the removal of heavy metals
from the contaminated soils but could also produce
safe corn for animal feed, allowing farmers to continue
their agricultural activities. The co-cropping of a Cd
hyperaccumulator Brassica napus with B. parachinen-
sis or Zea mays increased the Cd concentration and
accumulation in the shoot of the hyperaccumulator
plant (Selvam and Wong, 2009). Co-cropping system
could combine with chemical enhancement, including
chelator-enhanced phytoextraction (Wu et al., 2007c)
and chemical washing (Huang et al., 2010), to further
increase the phytoextraction efficiency.
For the non-hyperaccumulators, Li et al. (2009b)
studied the effect of 7 intercrops on Cd uptake by
maize for naturally enhancing the phytoremediation
with high biomass plants and the results showed that
legumes substantially increased Cd uptake by maize.
However, maize is not a hyperaccumulator so that its
phytoextraction efficiency is low, and the disposal of
the produced biomass with metal contents generally
too low to be recovered should be considered.
In a pot experiment conducted to study the in-
teraction mechanisms between the roots of two co-
cropped plants with the corn roots separated from S.
alfredii with a mesh barrier, co-cropping decreased pH
of the sludge solution, increased dissolved organic car-
bon (DOC), and resulted in higher Zn/Cd concentra-
tion than the S. alfredii mono-cropping, which caused
more Zn/Cd transported to the S. alfredii side and then
enhanced the uptake of the heavy metals by the hyper-
accumulator (Hei et al., 2007). The concentrations of
Cu, Pb, Zn, Cd, and Fe in the shoots of peas grown in
a mixed culture with barley in pots were 1.5, 1.8, 1.4,
1.4, and 1.3 times higher than those of the monocul-
ture (Luo et al., 2008). The root exudates from barley
in the mixed culture system played an important role
in the process of solubilizing metals in soil and facili-
tating the uptake of metals by peas.
The enhancing effect on the phytoextraction de-
pends on plant species in co-cropping. For example,
intercropping of Salix caprea with the hyperaccumula-
tor Arabidopsis halleri did not enhance total Zn and Cd
extraction compared to single plantings (Wieshammer
et al., 2007), which is potentially due to competition
for water or nutrients and possibly metal sequestration
into plants. Athyrium yokoscens, a fern that hyperac-
cumulates Zn and Cd, is often accompanied by another
Zn and Cd-hyperaccumulating plant, Arabis flagellosa;
a mixed planting of two plants of each species did not
largely changed the levels of Zn and Cd accumulated
by these two plants (Chen et al., 2009), which may be
related to competition for metals, and indicated that
it would be better to intercrop plants that hyperaccu-
mulate different metals.
CONCLUSIONS
This review presents a broad summary and evalu-
ation of cropping systems that are available for trace
element-contaminated sites. Three major cropping sys-
tems are identified, each of which targets on one of
the following crop production: edible plants, non-edible
plants, and bio-ores for industrial use (Fig. 2). The pri-
mary strategies include phytoexclusion, phytostabiliza-
tion, and phytoextraction/phytomining. Each strategy
has benefits and limitations, depending on the remedial
goals targeted and on-site and species-specific factors.
For safe production of edible food crops such as
cereals and vegetables, further research is needed to
select low-accumulating species/cultivars by means of
agronomic and genetic breeding. Chemical amend-
ments (e.g., alkaline agents, adsorption agents, and
fertilizers) should be optimized according to their effi-
ciency of metal immobilization and the economic feasi-
bility. In this case, the designed cropping system needs
to be evaluated mainly by production quality rather
than soil remediation.
At contaminated sites with harsh soil conditions
(e.g., acidity, low fertility, and multiple metals), crop-
ping industrial non-food energy and fiber plants will
likely be preferred. However, more attention should
be paid to amendment selection and soil management
practices to achieve a best crop yield and least poten-
tial hazard. On the other hand, for large-scale sites
with low to moderate levels of trace elements (e.g.,
As, Cd, Ni, and Zn), a plant-based phytoextraction
will be a possible option for decontamination of met-
als and/or extraction of metals with high price (e.g.,
Ni). However, the technology can in some cases be
time-consuming and it is anyway limited to the soil
depth prospected by roots of hyperaccumulators. Thus,
high-biomass hyperaccumulating species should be se-
lected, and further research is required for better un-
CROPPING SYSTEMS FOR METAL-CONTAMINATED SITES 481
482 Y. T. TANG et al.
derstanding the mechanisms involved in metal hyper-
accumulation. Further, biomass produced by phytomi-
ning is a substantial resource (i.e., 2% Ni), for which
a specific separation process should be developed.
Co-cropping two plant species with contrasting
metal-accumulating abilities could improve phytore-
mediation efficiency and remediate metal-contami-
nated soils in a naturally enhanced way. The intera-
ction mechanisms between the two plants deserve fur-
ther investigation and the combination ratios of the
two plants are warranted to be optimized.
In summary, the majority of these cropping sys-
tems are still far from maturity at present, but they
have shown the potential for a safe and regular agri-
culture on large-scale sites contaminated with trace
elements in a cost-effective manner. With appropriate
combination of soil types, plant species/cultivars, and
agronomic managements, toxic metals into the food
chain will be restricted and the potential benefits of-
fered by these cropping systems can be substantial in
the long run.
ACKNOWLEDGEMENT
Tang Ye-Tao gratefully acknowledges the grant
from the China Scholarship Council (CSC).
REFERENCES
Agarwal, A. K. 2007. Biofuels (alcohols and biodiesel) applica-
tions as fuels for internal combustion engines. Prog. Energ.
Combust.33: 233–271.
An, Z. Z., Huang, Z. C., Lei, M., Liao, X. Y., Zheng, Y. M.
and Chen, T. B. 2006. Zinc tolerance and accumulation in
Pteris vittata L. and its potential for phytoremediation of
Zn- and As-contaminated soil.Chemosphere. 62: 796–802.
Alexander, P. D., Alloway, B. J. and Dourado, A. M. 2006. Geno-
typic variations in the accumulation of Cd, Cu, Pb and Zn
exhibited by six commonly grown vegetables. Environ. Pol-
lut. 144: 736–745.
Anderson, L. L., Walsh, M., Roy, A., Bianchetti, C. M. and Mer-
chan, G. 2010. The potential of Thelypteris palustris and As-
paragus sprengeri in phytoremediation of arsenic contami-
nation. Int. J. Phytorem. 13: 177–184.
Angelova, V., Ivanova, R., Delibaltova, V. and Ivanov, K. 2004.
Bio-accumulation and distribution of heavy metals in fibre
crops (flax, cotton and hemp). Ind. Crop Prod. 19: 197–
205.
Arao, T. and Ae, N. 2003a. Genotypic variations in cadmium
levels of rice grain. Soil Sci. Plant Nutr. 49: 473–479.
Arao, T., Ae, N., Sugiyama, M. and Takahashi, M. 2003b. Geno-
typic differences in cadmium uptake and distribution in soy-
beans. Plant Soil. 251: 247–253.
Arru, L., Rognoni, S., Baroncini, M., Bonatti, P. and Perata, P.
2004. Copper localization in Cannabis sativa L. grown in a
copper-rich solution. Euphytica. 140: 33–38.
Baker, A. J. M., Reeves, R. D. and McGrath, S. P. 1991. In situ
decontamination of heavy metal polluted soils using crops of
metal accumulating plants: A feasibility study. In Hinchee,
R. E. and Olfenbuttel, R. F. (eds.) In situ Bioremediation.
Butterworth-Heinemann, Stoneham. pp. 539–544.
Bani, A. 2009. Phytoextraction du Ni dans les sols ultramafiques
d’Albanie. Ph.D. Thesis, Institut National Polytechnique de
Lorraine (INPL).
Bani, A., Echevarria, G., Mullaj, A., Reeves, R. D., Morel, J.
L. and Sul¸ce, S. 2009. Nickel hyperaccumulation by Brassi-
caceae in serpentine soils of Albania and Northwest Greece.
NE Nat.16: 385–404.
Bani, A., Echevarria, G., Sul¸ce, S., Morel, J. L. and Mullai, A.
2007. In-situ phytoextraction of Ni by a native population
of Alyssum murale on an ultramafic site (Albania). Plant
Soil.293: 79–89.
Ba˜nuelos, G. 2006. Multi-faceted considerations for sustainable
phytoremediation under field conditions. For. Snow Landsc.
Res.80: 235–245.
Ba˜nuelos, G., Da Roche, J. and Robinson, J. 2010. Develo-
ping selenium-enriched animal feed and biofuel from canola
planted for managing Se-laden drainage waters in the West-
side of Central California. Int. J. Phytorem.12:243–254.
Barbaroux, R., Meunier, N., Mercier, G., Taillard, V., Morel, J.
L., Simonnot, M. O. and Blais, J. F. 2009. Chemical leac-
hing of nickel from the seeds of the metal hyperaccumulator
plant Alyssum murale.Hydrometallurgy. 100: 10–14.
Barbaroux, R., Mercier, G., Blais, J. F., Morel, J. L. and Simon-
not, M. O. 2011. A new method for obtaining nickel metal
from the hyperaccumulator plant Alyssum murale.Separ.
Purif. Technol. 83: 57–65.
Barbaroux, R., Plasari, E., Mercier, G., Simonnot, M. O., Bani,
A., Morel, J. L. and Blais, J. F. 2012. A new process for
nickel ammonium disulphate production from ash of the hy-
peraccumulating plant Alyssum murale.Sci. Total Environ.
in press.
Basta, N. T. and McGowen, S. L. 2004. Evaluation of chemical
immobilization treatments for reducing heavy metal trans-
port in a smelter-contaminated soil. Environ. Pollut. 127:
73–82.
Boominathan, R., Saha-Chaudhury, N. M., Sahajwalla, V. and
Doran, P. M. 2004. Production of nickel bio-ore from hyper-
accumulator plant biomass: Applications in phytomining.
Biotechnol. Bioeng. 86: 243–250.
Brown, S., Christensen, B., Lombi, E., McLaughlin, M., Mc-
Grath, S., Colpaert, J. and Vangronsveld, J. 2005. An inter-
laboratory study to test the ability of amendments to reduce
the availability of Cd, Pb, and Zn in situ. Environ. Pollut.
138: 34–45.
Cai, B. S., Zhang, G. P. and Chen, T. B. 2007. Genotypic dif-
ference in growth and As accumulation in Pteris vittata.J.
Zhejiang Uni. Agr. Life Sci. 33: 473–478.
Cao, R. X., Ma, L. Q., Chen, M., Singh, S. P. and Harris, W.
G. 2003. Phosphate-induced metal immobilization in a con-
taminated site. Environ. Pollut. 122: 19–28.
Cao, X. D., Ma, L. Q., Rhue, D. R. and Appel, C. S. 2004.
Mechanisms of lead, copper, and zinc retention by phos-
phate rock. Environ. Pollut. 131: 435–444.
Cao, X. D., Ma, L. N., Shiralipour, A. and Harris, W. 2010.
Biomass reduction and arsenic transformation during com-
posting of arsenic-rich hyperaccumulator Pteris vittata L.
Environ. Sci. Pollut. Res. 17: 586–594.
Carlson, K. D., Cunningham, R. L., Garcia, W. J., Bagby, M.
O. and Kwolek, W. F. 1982. Performance and trace metal
content of crambe and kenaf grown on sewage sludge-treated
stripmine land. Environ. Pollut. A. Ecol. Biol. 29: 145–
161.
CROPPING SYSTEMS FOR METAL-CONTAMINATED SITES 483
Chaney, R. L. 1983. Plant uptake of inorganic waste constitunts.
In Parr, J. F., Marsh, P. B. and Kla, J. M. (eds.) Land
Treatment of Hazardous Wastes. Noyes Data Corp., Park
Ridge. pp. 50–76.
Chaney, R. L., Angle, J. S., Broadhurst, C. L., Peters, C. A.,
Tappero, R. V. and Donald, L. S. 2007. Improved under-
standing of hyperaccumulation yields commercial phytoex-
traction and phytomining technologies. J. Environ. Qual.
36: 1429–1443.
Chaney, R. L., Angle, J. S., McIntosh, M. S., Reeves, R. D.,
Li, Y. M., Brewer, E. P., Chen, K. Y., Roseberg, R. J.,
Perner, H., Synkowski, E. C., Broadhurst, C. L., Wang, S.
and Baker, A. J. M. 2005. Using hyperaccumulator plants
to phytoextract soil Ni and Cd. Zeitschrift Naturforschung
C.60: 190–198.
Chardot, V., Massoura, S. T., Echevarria, G., Reeves, R. D. and
Morel, J. L. 2005. Phytoextraction potential of the nickel
hyperaccumulators Leptoplax emarginata and Bornmuellera
tymphaea.Int. J. Phytorem.7: 323–335.
Chen, H. M., Zheng, C. R., Tu, C. and She, Z. G. 2000. Chemi-
cal methods and phytoremediation of soil contaminated with
heavy metals. Chemosphere. 41: 229–234.
Chen, J. 2007. Rapid urbanization in China: a real challenge to
soil protection and food security. Catena.69: 1–15.
Chen, R. X., Smith, B. W., Winefordner, J. D., Tu, M.
S., Kertulis, G. and Ma, L. Q. 2004. Arsenic speciation
in Chinese brake fern by ion-pair high-performance liquid
chromatography-inductively coupled plasma mass spectros-
copy. Anal. Chim. Acta. 504: 199–207.
Chen, T. B., Fan, Z. L., Lei, M., Huang, Z. C. and Wei, C.
Y. 2002a. Effect of phosphorus on arsenic accumulation in
As-hyperaccumulator Pteris vittata L. and its implication.
Chin. Sci. Bull.47: 1876–1879.
Chen, T. B., Liao, X. Y., Huang, Z. C., Lei, M. and Li, W.
X. 2007. Phytoremediation of arsenic-contaminated soil in
China. In Willey, N. (ed.) Phytoremediation: Methods and
Reviews. Humana Press, Totowa. pp. 393–404.
Chen, T. B., Wei, C. Y., Huang, Z. C., Huang, Q. F., Lu, Q.
G. and Fan, Z. L. 2002b. Arsenic hyperaccumulator Pteris
vittata L. and its arsenic accumulation. Chin.Sci.Bull.
47: 902–905.
Chen, Z., Setagawa, M., Kang, Y., Sakurai, K., Alkawa, Y. and
Iwasaki, K. 2009. Zinc and cadmium uptake from a metal-
liferous soil by a mixed culture of Athyrium yokoscense and
Arabis flagellosa. Soil Sci. Plant Nut. 55: 315–324.
Chopin, E. I. B., Marin, B., Mkoungafoko, R., Rigaux, A., Hop-
good, M. J., Delannoy, E., Canc`es, B. and Laurain, M. 2008.
Factors affecting distribution and mobility of trace elements
(Cu, Pb, Zn) in a perennial grapevine (Vitis vinifera L.)
in the Champagne region of France. Environ. Pollut.156:
1092–1098.
Cieslinski, G., Van Rees, K. C. J., Huang, P. M., Kozak, L. M.,
Rostad, H. P. W. and Knott, D. R. 1996. Cadmium uptake
and bioaccumulation in selected cultivars of durum wheat
and flax as affected by soil type. Plant Soil.182: 115–124.
Citterio, S., Santagostino, A., Fumagalli, P., Prato, N., Ranalli,
P. and Sgorbati, S. 2003. Heavy metal tolerance and accu-
mulation of Cd, Cr and Ni by Cannabis sativa.Plant Soil.
256: 243–252.
Clarke, J. M., Leisle, D. and Kopytko, G. L. 1997. Inheritance of
cadmium concentration in five durum wheat crosses. Crop
Sci. 37: 1722–1726.
Clarke, J. M., McCaig, T. N., DePauw, R. M., Knox, R. E.,
Clarke, F. R., Fernandez, M. R. and Ames, N. P. 2005.
Strongfield durum wheat. Can. J. Plant Sci. 85: 651–654.
da Cunha, K. P. V. and do Nascimento, C. W. A. 2009. Silicon
effects on metal tolerance and structural changes in maize
(Zea mays L.) grown on a cadmium and zinc enriched soil.
Water Air Soil Pollut. 197: 323–330.
De Kimpe, C. R. and Morel, J. L. 2000. Urban soil management:
A growing concern. Soil Sci. 165: 31–40.
Dunbar, K. R., McLaughlin, M. J. and Reid, R. J. 2003. The up-
take and partitioning of cadmium in two cultivars of potato
(Solanum tuberosum L.). J. Exp. Bot. 54: 349–354.
Dwivedi, S., Tripathi, R. D., Srivastava, S., Mishra, S., Shukla,
M. K., Tiwari, K. K., Singh, R. and Rai, U. N. 2007. Growth
performance and biochemical responses of three rice (Oryza
sativa L.) cultivars grown in fly-ash amended soil. Chemo-
sphere. 67: 140–151.
Echevarria, G., Massoura, S., Sterckeman, T., Becquer, T.,
Schwartz, C. and Morel, J. L. 2006. Assessment and con-
trol of the bioavailability of nickel in soils. Environ. Toxicol.
Chem.25: 643–651.
Ebbs, S., Hatfield, S., Nagarajan, V. and Blaylock, M. 2009. A
comparison of the dietary arsenic exposures from ingestion
of contaminated soil and hyperaccumulating Pteris ferns
used in a residential phytoremediation project. Int. J. Phy-
torem. 12: 121–132.
Erdman, J. A. and Moul, R. C. 1982. Mineral composition
of small-grain cultivars from a uniform test plot in South
Dakota. J. Agr . Food Chem. 30: 169–174.
Ernst, W. H. O. 2005. Phytoextraction of mine wastes—options
and impossibilities. Chem. Erde Geochem.65: 29–42.
Feng, R. W., Wei, C. Y., Tu, S. X., Tang, S. R. and Wu, F. C.
2011. Simultaneous hyperaccumulation of arsenic and anti-
mony in Cretan brake fern: Evidence of plant uptake and
subcellular distributions. Microchem. J. 97: 38–43.
Florijn, P. J. and Van Beusichem, M. L. 1993. Uptake and distri-
bution of cadmium in maize inbred lines. Plant Soil. 150:
25–32.
Fu, J., Zhou, Q., Liu, J., Liu, W., Wang, T., Zhang, Q. and
Jiang, G. 2008. High levels of heavy metals in rice (Oryza
sativa L.) from a typical E-waste recycling area in southeast
China and its potential risk to human health. Chemosphere.
71:1269–1275.
Gao, X. P., Mohr, R. M., McLaren, D. L. and Grant, C. A. 2011.
Grain cadmium and zinc concentrations in wheat as affected
by genotypic variation and potassium chloride fertilization.
Field Crops Res. 122: 95–103.
Gao, Y. and Marschner, H. 1995. Uptake, distribution, and bin-
ding of cadmium and nickel in different plant species. J.
Plant Nutr. 18: 2691–2706.
erard, E., Echevarria, G., Sterckeman, T. and Morel, J. L.
2001. Cadmium availability to three plant species varying
in cadmium accumulation pattern. J. Environ. Qual.29:
1117–1123.
Grant, C. A., Clarke, J. M., Duguid, S. and Chaney, R. L. 2008.
Selection and breeding of plant cultivars to minimize cad-
mium accumulation. Sci. Total Environ. 390: 301–310.
Greger, M. and Lofstedt, M. 2004. Comparison of uptake and
distribution of cadmium in different cultivars of bread and
durum wheat. Corp Sci. 44: 501–507.
Gove, B., Hutchinson, J. J., Young, S. D., Craigon, J. and Mc-
Grath, S. P. 2002. Uptake of metals by plants sharing a rhi-
zosphere with the hyperaccumulator Thlaspi caerulescens.
Int. J. Phytorem. 4: 267–281.
Gonzaga, M., Ma, L. and Santos, J. 2007. Effects of plant age
on arsenic hyperaccumulation by Pteris vittata L. Water Air
Soil Pollut. 186: 289–295.
484 Y. T. TANG et al.
Gu, H. H., Qiu, H., Tian, T., Zhan, S. S., Deng, T. H. B., Chaney,
R. L., Wang, S. Z., Tang, Y. T., Morel, J. L. and Qiu, R. L.
2011. Mitigation effects of fly ash and steel slag on heavy
metal accumulation in Oryza sativa L. planted on multi-
metal contaminated soil. Chemosphere.83: 1234–1240.
Guo, X. F., Wei, Z. B., Qiu, J. R., Wu, Q. T. and Zhou, J.
L. 2010. Differences between corn cultivars in accumulation
and translocation of heavy metals. J. Ecol. Rural Environ.
26: 367–371.
Habashi, J. P., Holm, T., Loeys, B. L., Bedja, D., Neptune, E. R.,
Holmes, K. W., Judge, D. P. and Dietz, H. C. 2005. Losar-
tan arrests aortic root enlargement in Marfan syndrome in-
dependent of hemodynamic effects. Circulation.112: U32.
Harris, N. S. and Taylor, G. J. 2004. Cadmium uptake and
translocation in seedlings of near isogenic lines of du-
rum wheat that differ in grain cadmium accumulation.
BMC Plant Biol. 4:4.
Hasfalina, C. M., Maryam, R. Z., Luqman, C. A. and Rashid, M.
2010. The potential use of kenaf as a bioadsorbent for the
removal of copper and nickel from single and binary aqueous
solution. J. Natl. Fiber.7: 267–275.
Hei, L., Wu, Q. T., Long, X. X. and Hu, Y. M. 2007. Ef-
fect of co-planting of Sedum alfredii and Zea mays on Zn-
contaminated sewage sludge. Environ. Sci. (in Chinese).
28: 852–858.
Ho, W. M., Ang, L. H. and Lee, D. K. 2008. Assessment of Pb
uptake, translocation and immobilization in kenaf (Hibiscus
cannabinus L.) for phytoremediation of sand tailings. J. En-
viron. Sci. 20: 1341–1347.
Huang, B., Xin, J., Yang, Z., Zhou, Y., Yuan, J. and Gong,
Y. 2009. Suppression subtractive hybridization (SSH)-based
method for estimating Cd-induced differences in gene ex-
pression at cultivar level and identification of genes induced
by Cd in two water spinach cultivars. J . Agr . Food Ch em .
57: 8950–8962.
Huang, S. H. 2009. Study on immobilization of heavy metals
polluted soils in Dabao Mountain by amendments and their
mechanism. M. S. Thesis, Sun Yat-Sen University.
Huang, X. H, Wei, Z. B., Guo, X. F., Shi, X. F. and Wu, Q. T.
2010. Metal removal from contaminated soil by co-planting
phytoextraction and soil washing. Environ. Sci. (in Chi-
nese). 31: 3067–3074.
Huang, Z. C., An, Z. Z., Chen, T. B., Lei, M., Xiao, X. Y. and
Liao, X. Y. 2007. Arsenic uptake and transport of Pteris
vittata L. as influenced by phosphate and inorganic arsenic
species under sand culture. J. Environ Sci. 19: 714–718.
Huang, Z. C., Chen, T. B., Lei, M., Liu, Y. R. and Hu, T. D.
2008. Difference of toxicity and accumulation of methylated
and inorganic arsenic in arsenic-hyperaccumulating and -
hypertolerant plants. Environ. Sci. Technol. 42: 5106–5111.
Huang, Z. C., Chen, T. B., Lei, M. and Hu, T. D. 2004. Direct
determination of arsenic species in arsenic hyperaccumula-
tor Pteris vittata by EXAFS. Acta Bot. Sin. 46: 46–50.
Ishikawa, S., Ae, N. and Yano, M. 2005a. Chromosomal regions
with quantitative trait loci controlling cadmium concentra-
tion in brown rice (Oryza sativa). New Phytol. 168: 345–
350.
Ishikawa, S., Ae, N., Sugiyama, M., Murakami, M. and Arao, T.
2005b. Genotypic variation in shoot cadmium concentration
in rice and soybean in soils with different levels of cadmium
contamination. Soil Sci. Plant Nutr. 51: 101–108.
Jiang, C. A., Wu, Q. T., Sterckeman, T., Schwartz, C., Sirguey,
C., Ouvrard, S., Perriguey, J. and Morel, J. L. 2010. Co-
planting can phytoextract similar amounts of cadmium and
zinc to mono-cropping from contaminated soils. Ecol. Eng.
36: 391–395.
John, M. K. 1973. Cadmium uptake by eight food crops as in-
fluenced by various soil levels of cadmium. Environ. Pollut.
4: 7–15.
Juwarkar, A. A., Yadav, S. K., Kumar, P. and Singh, S. K. 2008.
Effect of biosludge and biofertilizer amendment on growth
of Jatropha curcas in heavy metal contaminated soils. Envi-
ron. Monit. Assess. 145: 7–15.
Kalve, S., Sarangi, B. K., Pandey, R. A. and Chakrabarti, T.
2011. Arsenic and chromium hyperaccumulation by an eco-
type of Pteris vittata—prospective for phytoextraction from
contaminated water and soil. Curr. Sci. 100: 888–894.
Karimi, N., Ghaderian, S. M., Maroofi, H. and Schat, H. 2010.
Analysis of arsenic in soil and vegetation of a contaminated
area in Zarshuran, Iran. Int. J. Phytorem. 12: 159–173.
Karimi, N., Ghaderian, S. M., Raab, A., Feldmann, J. and
Meharg, A. A. 2009. An arsenic-accumulating, hypertole-
rant brassica, Isatis capadocica.New Phytol.184: 41–47.
Kashem, M., Kawai, S., Kikuchi, N., Takahashi, H., Sugawara,
R. and Singh, B. 2010. Effect of lherzolite on chemical frac-
tions of Cd and Zn and their uptake by plants in contami-
nated soil. Water Air Soil Pollut. 207: 241–251.
Keller, C. and Hammer, D. 2004. Metal availability and soil to-
xicity after repeated croppings of Thlaspi caerulescens in
metal contaminated soils. Environ. Pollut. 131: 243–254.
Keller, C., Ludwig, C., Davoli, F. and Wochele, J. 2005. Thermal
treatment of metal-enriched biomass produced from heavy
metal phytoextraction. Environ. Sci. Technol. 39: 3359–
3367.
Khan, M. J. and Jones, D. L. 2009. Effect of composts, lime and
diammonium phosphate on the phytoavailability of heavy
metals in a copper mine tailing soil. Pedosphere. 19: 631–
641.
Khan, S., Rehman, S., Khan, A. Z., Khan, M. A. and Shah, M.
T. 2010. Soil and vegetables enrichment with heavy metals
from geological sources in Gilgit, northern Pakistan. Ecotox.
Environ. Safe. 73: 1820–1827.
Koopmans, G. F., R¨omkens, P. F. A. M., Fokkema, M. J., Song,
J., Luo, Y. M., Japenga, J. and Zhao, F. J. 2008. Feasibility
of phytoextraction to remediate cadmium and zinc contam-
inated soils. Environ. Pollut.156: 905–914.
Kukier, U., Peters, C. A., Chaney, R. L., Angle, J. S. and Rose-
berg, R. J. 2004. The effect of pH on metal accumulation
in two Alyssum species. J. Environ. Qual.33:2090–2102.
Kumar, G. P., Yadav, S. K., Thawale, P. R., Singh, S. K. and
Juwarkar, A. A. 2008. Growth of Jatropha curcas on heavy
metal contaminated soil amended with industrial wastes and
Azotobacter—A greenhouse study. Bioresour. Technol. 99:
2078–2082.
Kumpiene, J., Lagerkvist, A. and Maurice, C. 2008. Stabiliza-
tion of As, Cr, Cu, Pb and Zn in soil using amendments—a
review. Waste Manage. 28: 215–225.
Kurz, H., Schulz, R. and Romheld, V. 1999. Selection of culti-
vars to reduce the concentration of cadmium and thallium
in food and fodder plants. J. Plant Nutr. Soil Sci. 162:
323–328.
Lai, H. Y., Juang, K. W. and Chen, B. C. 2010. Copper concen-
trations in grapevines and vineyard soils in central Taiwan.
Soil Sci. Plant Nut.56: 601–606.
Leung, H. M., Wu, F. Y., Cheung, K. C., Ye, Z. H. and Wong,
M. H. 2010. The effect of arbuscular mycorrhizal fungi and
phosphate amendment on arsenic uptake, accumulation and
growth of Pteris vittata in As-contaminated soil. Int. J. Phy-
torem. 12: 384–403.
CROPPING SYSTEMS FOR METAL-CONTAMINATED SITES 485
Li, Y. M., Chaney, R., Brewer, E., Roseberg, R., Angle, J. S.,
Baker, A., Reeves, R. and Nelkin, J. 2003. Development of
a technology for commercial phytoextraction of nickel: eco-
nomic and technical considerations. Plant Soil.249: 107–
115.
Li, H., Shi, W. Y., Shao, H. B. and Shao, M. A. 2009a. The re-
mediation of the lead-polluted garden soil by natural zeolite.
J. Hazard. Mater. 169: 1106–1111.
Li, N. Y., Li, Z. A., Zhuang, P., Zou, B. and McBride, M. 2009b.
Cadmium uptake from soil by maize with intercrops. Water
Air Soil Pollut.199: 45–56.
Li, P., Wang, X. X., Zhang, T. L., Zhou, D. M. and He, Y. Q.
2008. Effects of several amendments on rice growth and up-
take of copper and cadmium from a contaminated soil. J.
Environ. Sci. 20: 449–450.
Li, P., Wang, X. X., Zhou, T. L., Zhou, D. M. and He, Y. Q.
2009c. Distribution and accumulation of copper and cad-
mium in soil-rice system as affected by soil amendments.
Water Air Soil Pollut. 196: 29–40.
Li, Q. F., Qiu, R. L., Shi, N., Zhou, X. Y., Senthilkumar, P. and
Huang, S. H. 2009d. Remediation of strongly acidic mine
soils contaminated by multiple metals by plant reclamation
with Jatropha curcas L. and addition of limestone. Acta Sci.
Circumst. (in Chinese). 29: 1733–1739.
Li, W. X., Chen, T. B. and Liu, Y. R. 2005a. Effects of har-
vesting on As accumulation and removal efficiency of As by
Chinese brake (Pteris vittata L.). Acta Ecol. Sin. (in Chi-
nese). 25: 538–542.
Li, X. B., Xie, J. Z., Li, B. W. and Wang, W. 2009e. Ecolo-
gical responses of Brassica juncea-alfalfa intercropping to
cadmium stress. Chin. J. Appl. Ecol. (in Chinese). 20:
1711–1715.
Li, Z. W., Li, L. Q., Pan, G. X. and Chen, J. 2005b. Bioavaila-
bility of Cd in a soil-rice system in China: soil type versus
genotype effects. Plant Soil.271: 165–173.
Liang, Y. C., Sun, W. C., Zhu, Y. G. and Christie, P.
2007. Mechanisms of silicon-mediated alleviation of abiotic
stresses in higher plants: a review. Environ. Pollut. 147:
422–428.
Liang, Y. C., Wong, J. W. C. and Wei, L. 2005. Silicon-mediated
enhancement of cadmium tolerance in maize (Zea mays L.)
grown in cadmium contaminated soil. Chemosphere. 58:
475–483.
Liao, X. Y., Chen, T. B., Xiao, X. Y., Xie, H., Yan, X. L., Zhai,
L. M. and Wu, B. 2007. Selecting appropriate forms of ni-
trogen fertilizer to enhance soil arsenic removal by Pteris
vittata: A new approach in phytoremediation. Int. J. Phy-
torem. 9: 269–280.
Lin, J., Jiang, W. and Liu, D. 2003. Accumulation of copper by
roots, hypocotyls and leaves of sunflower (Helianthus an-
nuus L.). Biores. Technol.86:151–155.
Linger, P., M¨ussig, J., Fischer, H. and Kobert, J. 2002. Indus-
trial hemp (Cannabis sativa L.) growing on heavy metal
contaminated soil: fibre quality and phytoremediation po-
tential. Ind. Crop. Prod. 16: 33–42.
Liu, C. P., Li, F. B., Lou, C. L., Liu, X. M., Wang, S. H., Liu, T.
X. and Li, X. D. 2009. Foliar application of two silica sols
reduced cadmium accumulation in rice grains. J. Hazard.
Mater. 161: 1466–1472.
Liu, F. H., Li, Z., Liu, Q., He, H., Liang, X. and Lai, Z. 2003a.
Introduction to the wild resources of the genus Boehmeria
Jacq. in China. Genet. Resour. Crop Evol. 50: 793–797.
Liu,J.G.,Li,K.Q.,Xu,J.K.,Liang,J.S.,Lu,X.L.,Yang,J.
C. and Zhu, Q. S. 2003b. Interaction of Cd and five mine-
ral nutrients for uptake and accumulation in different rice
cultivars and genotypes. Field Crop Res. 83: 271–281.
Liu, J. G., Qian, M., Cai, G. L., Yang, J. C. and Zhu, Q. S.
2007a. Uptake and translocation of Cd in different rice cul-
tivars and the relation with Cd accumulation in rice grain.
J. Hazard. Mater. 143: 443–447.
Liu, J. G., Zhu, Q. S., Zhang, Z. J., Xu, J. K., Yang, J. C. and
Wong, M. H. 2005a. Variations in cadmium accumulation
among rice cultivars and types and the selection of culti-
vars for reducing cadmium in the diet. J. Sci. Food Agr. 85:
147–153.
Liu, X. M., Wu, Q. T. and Banks, M. K. 2005b. Effect of simu-
ltaneous establishment of Sedum alfredii and Zea mays on
heavy metal accumulation in plants. Int. J. Phytorem.7:
41–53.
Liu, Y. G., Ye, F., Zeng, G. M., Fan, T., Meng, L. and Yuan,
H. S. 2007b. Effects of added Cd on Cd uptake by oilseed
rape and pai-tsai co-cropping. Trans. Nonferr. Metal Soc.
China.17: 846–852.
Liu, Y., Zhu, Y. G., Chen, B. D., Christie, P. and Li, X. L.
2005c. Influence of the arbuscular mycorrhizal fungus Glo-
mus mosseae on uptake of arsenate by the As hyperaccu-
mulator fern Pteris vittata L. Mycorrhiza. 15: 187–192.
Lun´ckov´a, L., Masaroviˇcov´a, E., Kr´al’ov´a, K. and Streˇsko, V.
2003. Response of fast growing woody plants from family
Salicaceae to cadmium treatment. Bull. Environ. Contam.
Tox i co l.70: 576–585.
Luo, C., Liu, C., Wang, Y., Liu, X., Li, F., Zhang, G. and Li,
X. 2011. Heavy metal contamination in soils and vegetables
near an e-waste processing site, south China. J. Hazard.
Mater. 186: 481–490.
Luo, C. L., Shen, Z. G. and Li, X. D. 2008. Root exudates in-
crease metal accumulation in mixed cultures: implications
for naturally enhanced phytoextraction. Water Air Soil Pol-
lut.193: 147–154.
Ma, L. Q., Komar, K. M., Tu, C., Zhang, W. H., Cai, Y. and
Kennelley, E. D. 2001. A fern that hyperaccumulates arse-
nic. Nature.409: 579.
Maddox, J. J. and Soileau, J. M. 1991. Effects of phosphate
fertilization, lime amendments and inoculation with VA-
mycorrhizal fungi on soybeans in an acid soil. Plant Soil.
134: 83–93.
Malandrino, M., Abollino, O., Buoso, S., Giacomino, A., La
Gioia, C. and Mentasti, E. 2011. Accumulation of heavy
metals from contaminated soil to plants and evaluation of
soil remediation by vermiculite. Chemosphere.82: 169–178.
Massoura, S. T., Echevarria, G., Leclerc-Cessac, E. and Morel,
J. L. 2004. Response of excluder, indicator and hyperaccu-
mulator plants to nickel availability in soils. Aust. J. Soil
Res. 42: 933–938.
Mathews, S., Ma, L. Q., Rathinasabapathi, B., Natarajan, S. and
Saha, U. K. 2010. Arsenic transformation in the growth
media and biomass of hyperaccumulator Pteris vittata L.
Bioresour. Technol. 101: 8024–8030.
McGrath, S. P., Lombi, E., Gray, C. W., Caille, N., Dunham,
S. J. and Zhao, F. J. 2006. Field evaluation of Cd and Zn
phytoextraction potential by the hyperaccumulators Thlaspi
caerulescens and Arabidopsis halleri.Environ. Pollut.141:
115–125.
Meers, E., Ruttens, A., Hopgood, M., Lesage, E. and Tack, F.
M. G. 2005. Potential of Brassica rapa,Cannabis sativa,
Helianthus annuus and Zea mays for phytoextraction of
heavy metals from calcareous dredged sediment derived
soils. Chemosphere.61: 561–572.
Meers, E., Vandecasteele, B., Ruttens, A., Vangronsveld, J. and
Tack, F. M. G. 2007. Potential of five willow species (Salix
486 Y. T. TANG et al.
spp.) for phytoextraction of heavy metals. Environ. Exp.
Bot.60: 57–68.
Meers, E., Van Slycken, S., Adriaensen, K., Ruttens, A., Van-
gronsveld, J., Du Laing, G., Witters, N., Thewys, T. and
Tack, F. M. G. 2010. The use of bio-energy crops (Zea
mays) for ‘phytoattenuation’ of heavy metals on moderately
contaminated soils: A field experiment. Chemosphere. 78:
35–41.
Mirlean, N., Roisenberg, A. and Chies, J. O. 2005. Copper-based
fungicide contamination and metal distribution in Brazilian
grape products. Bull. Environ. Contam. Toxicol. 75: 968–
974.
Morel, J. L. 1997. Bioavailability of trace elements to terrestrial
plants. Chapter 6. In Tarradellas, J., Bitton, G. and Rossel,
D. (eds.) Soil Ecotoxicology. Lewis Publishers, CRC Press,
Boca Raton. pp. 141–176.
Morel, J. L., Mench, M. and Guckert, A. 1986. Measurement of
Pb2+,Cu
2+ and Cd2+ binding with mucilage exudates from
maize (Zea mays L.) roots. Biol. Fertil. Soils.2: 29–34.
Morishita, T., Fumoto, N., Yoshizawa, T. and Kagawa, K.
1987. Varietal differences in cadmium levels of rice grains of
Japonica, Indica, Javanica and hybrid varieties produced in
the same plot of a field. Soil Sci. Plant Nutr. 33: 629–637.
Muthukumar, T. and Bagyaraj, D. J. 2010. Use of arbuscular
mycorrhizal fungi in phytoremediation of heavy metal con-
taminated soils. Proc.Natl.A.Sci.IndiaB.5: 103–121.
Nabulo, G., Young, S. D. and Black, C. R. 2010. Assessing risk to
human health from tropical leafy vegetables grown on con-
taminated urban soils. Sci. Total Environ. 408: 5338–5351.
Nagaraju, A. and Karimulla, S. 2002. Accumulation of elements
in plants and soils in and around Nellore mica belt, Andhra
Pradesh, India—a biogeochemical study. Environ. Geol. 41:
852–860.
Natarajan, S., Stamps, R. H., Ma, L. N. Q., Saha, U. K., Hernan-
dez, D., Cai, Y. and Zillioux, E. J. 2011. Phytoremediation
of arsenic-contaminated groundwater using arsenic hyper-
accumulator Pteris vittata L.: Effects of frond harvesting
regimes and arsenic levels in refill water. J. Hazard. Mater.
185: 983–989.
Van Nevel, L., Mertens, J., Oorts, K. and Verheyen, K. 2007.
Phytoextraction of metals from soils: How far from prac-
tice? Environ. Pollut. 150: 34–40.
Nguyen, T. H. H., Sakakibara, M., Sano, S. and Mai, T. N. 2011.
Uptake of metals and metalloids by plants growing in a lead-
zinc mine area, Northern Vietnam. J. Hazard. Mater. 186:
1384–1391.
Oliver, D. P., Gartrell, J. W., Tiller, K. G., Correll, R., Coz-
ens, G. D. and Youngberg, B. L. 1995. Differential responses
of Australian wheat cultivars to cadmium concentration in
wheat grain. Aus. J. Agr. Res. 46: 873–886.
Ownby, D. R., Galvan, K. A. and Lydy, M. J. 2005. Lead and zinc
bioavailability to Eisenia fetida after phosphorus amend-
ment to repository soils. Environ. Pollut. 136: 315–321.
Peer, W. A., Mahmoudian, M., Freeman, J. L., Lahner, B.,
Richards, E. L., Reeves, R. D., Murphy, A. S. and Salt, D.
E. 2006. Assessment of plants from the Brassicaceae family
as genetic models for the study of nickel and zinc hyperac-
cumulation. New Phytol. 172: 248–260.
Pichtel, J. and Salt, C. A. 1998. Vegetative growth and trace
metal accumulation on metalliferous wastes. J. Environ.
Qual. 27: 618–624.
Reddy, M. R. and Dunn, S. J. 1986. Heavy-metal absorption by
soybean on sewage sludge treated soil. J. Agr . Food Che m.
34: 750–753.
Robinson, B. H., Brooks, R. R. and Clothier, B. E. 1999. Soil
amendments affecting nickel and cobalt uptake by Berkheya
coddii: Potential use for phytomining and phytoremediation.
Ann. Bot. 84: 689–694.
Roccotiello, E., Manfredi, A., Drava, G., Minganti, V., Mariotti,
M. G., Berta, G. and Cornara, L. 2010. Zinc tolerance and
accumulation in the ferns Polypodium cambricum L. and
Pteris vittata L. Ecotox. Environ. Safe. 73: 1264–1271.
Rosselli, W., Keller, C. and Boschi, K. 2003. Phytoextraction ca-
pacity of trees growing on a metal contaminated soil. Plant
Soil.256: 265–272.
Ruttens, A., Adriaensen, K., Meers, E., De Vocht, A., Gee-
belen, W., Carleer, R., Mench, M. and Vangronsveld, J.
2010. Long-term sustainability of metal immobilization by
soil amendments: Cyclonic ashes versus lime addition. En-
viron. Pollut. 158: 1428–1434.
Saikia, C. N., Goswami, T. and Ali, F. 1997. Evaluation of
pulp and paper making characteristics of certain fast gro-
wing plants. Wood Sci. Technol. 31: 467–475.
Salt, D. E., Kumar, P. B. A. N., Dushenkov, S. and Raskin, I.
1994. Phytoremediation: a new technology for the environ-
mental cleanup of toxic metals. In Mahant, P., Pickles, C.
and Lu. W.-K. (eds.) Proceedings of the International Sym-
posium on Resource Conservation and Environmental Tech-
nologies in Metallurgical Industries. Canadian Institute of
Mining, Metallurgy and Petroleum, Qu´ebec. pp. 381–384.
Samake, M., Wu, Q. T., Mo, C. H. and Morel, J. L. 2003. Plants
grown on sewage sludge in South China and its relevance to
sludge stabilization and metal removal. J. Environ. Sci. 15:
622–627.
Santos, J. A. G., Gonzaga, M. I. S. and Ma, L. Q. 2010.
Optimum P levels for arsenic removal from contaminated
groundwater by Pteris vittata L. of different ages. J. Ha-
zard. Mater. 180: 662–667.
Schwartz, C., Echevarria, G. and Morel, J. L. 2003. Phytoex-
traction of cadmium with Thlaspi caerulescens.Plant Soil.
249: 27–35.
Selvam, A. and Wong, J. W. C. 2009. Cadmium uptake potential
of Brassica napus cocropped with Brassica parachinensis
and Zea mays.J. Hazard. Mater.167: 170–178.
er´e, G., Schwartz, C., Ouvrard, S., Sauvage, C., Renat, J. C.
and Morel, J. L. 2008. Soil construction: A step for eco-
logical reclamation of derelict lands. J. Soils Sediment. 8:
130–136.
Shallari, S., Echevarria, G., Schwartz, C. and Morel, J. L.
2001. Availability of nickel in soils for the hyperaccumulator
Alyssum murale (Waldst. & Kit.). South Afr. J. Sci. 97:
568–570.
Shi, J., Li, L. Q. and Pan, G. X. 2009. Variation of grain Cd
and Zn concentrations of 110 hybrid rice cultivars grown in
a low-Cd paddy soil. J. Environ. Sci. 21: 168–172.
Shi, X. H., Zhang, C. C., Wang, H. and Zhang, F. S. 2005. Effect
of Si on the distribution of Cd in rice seedlings. Plant Soil.
272: 53–60.
Simon, L. 2005. Stabilization of metals in acidic mine spoil with
amendments and red fescue (Festuca rubra L.) growth. En-
viron. Geochem. Health.27:289–300.
Simonnot, M. O. and Croze, V. 2008. Proc´ed´es de traitement
physique et chimique des sols pollu´es. Tec h .Ing´enieur.JB5:
J3981.
Sinaj, S., Dubois, A. and Frossard, E. 2004. Soil isotopically
exchangeable zinc: a comparison between E and L values.
Plant Soil.261: 17–28.
Singh, A., Sharma, R. K., Agrawal, M. and Marshall, F. M.
2010. Health risk assessment of heavy metals via dietary in-
CROPPING SYSTEMS FOR METAL-CONTAMINATED SITES 487
take of foodstuffs from the wastewater irrigated site of a dry
tropical area of India. Food Chem. Toxicol. 48: 611–619.
Su, D. C., Lu, X. X. and Wong, J. W. C. 2008a. Could cocrop-
ping or successive cropping with Cd accumulator oilseed
rape reduce Cd uptake of sensitive Chinese Cabbage? Pract.
Period. Hazard. Toxic Radioact. Waste Manage.12: 224–
228.
Su, Y., Han, F. X., Chen, J., Sridhar, B. B. M. and Monts,
D. L. 2008b. Phytoextraction and accumulation of mercury
in three plant species: Indian mustard (Brassica juncea),
beard grass (Polypogon monospeliensis), and Chinese brake
fern (Pteris vittata). Int. J. Phytorem. 10: 547–560.
Tu, C. and Ma, L. Q. 2002. Effects of arsenic concentrations
and forms on arsenic uptake by the hyperaccumulator lad-
der brake. J. Environ. Qual. 31: 641–647.
Tu, S., Ma, L. Q., MacDonald, G. E. and Bondada, B. 2004.
Effects of arsenic species and phosphorus on arsenic absorp-
tion, arsenate reduction and thiol formation in excised parts
of Pteris vittata L. Environ. Exp. Bot.51: 121–131.
Tudoreanu, L. and Phillips, C. J. C. 2004. Empirical models of
cadmium accumulation in maize, rye grass and soya bean
plants. J. Sc i. Food Agr. 84: 845–852.
Ueno, D., Yamaji, N., Kono, I., Huang, C. F., Ando, T., Yano,
M. and Ma, J. F. 2010. Gene limiting cadmium accumula-
tion in rice. Proc. Natl. Acad. Sci. USA. 107: 16500–
16505.
Verbruggen, N., Hermans, C. and Schat, H. 2009. Molecu-
lar mechanisms of metal hyperaccumulation in plants. New
Phytol. 181: 759–776.
Ververis, C., Georghiou, K., Christodoulakis, N., Santas, P. and
Santas, R. 2004. Fiber dimensions, lignin and cellulose con-
tent of various plant materials and their suitability for paper
production. Ind. Crop. Prod. 19: 245–254.
Wan, X. M., Lei, M., Huang, Z. C., Chen, T. B. and Liu, Y. R.
2010. Sexual propagation of Pteris Vittata L. influenced by
pH, calcium, and temperature. Int. J. Phytorem.12: 85–95.
Wang, H. B., Wong, M. H., Lan, C. Y., Baker, A. J. M., Qin,
Y. R., Shu, W. S., Chen, G. Z. and Ye, Z. H. 2007a. Uptake
and accumulation of arsenic by 11 Pteris taxa from southern
China. Environ Pollut. 145: 225–233.
Wang, J., Fang, W., Yang, Z., Yuan, J., Zhu, Y. and Yu, H.
2007b. Inter- and intraspecific variations of cadmium accu-
mulation of 13 leafy vegetable species in a greenhouse ex-
periment. J. Agr. Food Chem. 55: 9118–9123.
Wang, J., Yuan, J., Yang, Z., Huang, B., Zhou, Y., Xin, J., Gong,
Y. and Yu, H. 2009a. Variation in cadmium accumulation
among 30 cultivars and cadmium subcellular distribution
in 2 selected cultivars of water spinach (Ipomoea aquatica
Forsk.). J. Agr. Food Chem. 57: 8942–8949.
Wang, J. R., Zhao, F. J., Meharg, A. A., Raab, A., Feldmann, J.
and McGrath, S. P. 2002. Mechanisms of arsenic hyperac-
cumulation in Pteris vittata. Uptake kinetics, interactions
with phosphate, and arsenic speciation. Plant Physiol.130:
1552–1561.
Wang, Q. Y., Zhou, D. M. and Cang, L. 2009b. Microbial and
enzyme properties of apple orchard soil as affected by long-
term application of copper fungicide. Soil Biol. Biochem.
41: 1504–1509.
Wang, X., Liu, Y., Zeng, G., Chai, L., Song, X., Min, Z. and
Xiao, X. 2008. Subcellular distribution and chemical forms
of cadmium in Bechmeria nivea (L.) Gaud. Environ. Exp.
Bot. 62: 389–395.
Wei, C. Y. and Chen, T. B. 2006. Arsenic accumulation by two
brake ferns growing on an arsenic mine and their potential
in phytoremediation. Chemosphere. 63: 1048–1053.
Whiting, S. N., Leake, J. R., McGrath, S. P. and Baker, A. J.
M. 2001. Hyperaccumulation of Zn by Thlaspi caerulescens
can ameliorate Zn toxicity in the rhizosphere of co-cropped
Thlaspi arvense. Environ. Sci. Technol. 35: 3237–3241.
Wieshammer, G., Unterbrunner, R., Garcia, T. B., Zivkovic, M.
F., Puschenreiter, M. and Wenzel, W. W. 2007. Phytoex-
traction of Cd and Zn from agricultural soils by Salix ssp.
and intercropping of Salix caprea and Arabidopsis halleri.
Plant Soil.298: 255–264.
Wightwick, A. M., Salzman, S. A., Reichman, S. M., Allinson,
G. and Menzies, N. W. 2010. Inter-regional variability in en-
vironmental availability of fungicide derived copper in vine-
yard soils: An Australian case study. J. Agr . Food Chem.
58: 449–457.
Wu, F. B., Zhang, G. P., Dominy, P., Wu, H. X. and Bachir,
D. M. L. 2007a. Differences in yield components and kernel
Cd accumulation in response to Cd toxicity in four barley
genotypes. Chemosphere. 70: 83–92.
Wu, F. Y., Leung, H. M., .Wu, S. C., Ye, Z. H. and Wong, M. H.
2009. Variation in arsenic, lead and zinc tolerance and accu-
mulation in six populations of Pteris vittata L. from China.
Environ. Pollut. 157: 2394–2404.
Wu, H. J., Li, L. and Zhang, F. X. 2003. The influence of inter-
specific interactions on Cd uptake by rice and wheat inter-
cropping. Review China Agr. Sci. Technol. (in Chinese). 5:
43–47.
Wu, Q. H., Wang, S. Z., Palaniswamy, T., Li, Q. F., Zheng, H.
and Qiu, R. L. 2011. Phytostabilization potential of Jatro-
pha curcas L. in polymetallic acid mine tailings. Int. J.
Phytorem.13: 788–804.
Wu, Q. T., Wang, G. S., Tan, X. F., Chen, L. and He, L. C.
1994. Effect of crop cultivars and chemical fertilizers on the
cadmium accumulation in plants. J. South China Agr. Univ.
15: 1–6.
Wu, Q. T., Chen, L. and Wang, G. S. 1999. Differences on Cd
uptake and accumulation among rice cultivars and its mech-
anism. Acta Ecol. Sini. (in Chinese). 19: 104–107.
Wu, Q. T., Hei, L., Wong, J. W. C., Schwartz, C. and Morel, J. L.
2007b. Co-cropping for phyto-separation of zinc and potas-
sium from sewage sludge. Chemosphere.68:1954–1960.
Wu, Q. T., Wei, Z. B. and Ouyang, Y. 2007c. Phytoextraction
of metal-contaminated soil by hyperaccumulator Sedum al-
fredii H: effects of chelator and co-planting. Water Air Soil
Pollut.180: 131–139.
Xiao, X. Y., Chen, T. B., An, Z. Z., Lei, M., Huang, Z. C.,
Liao, X. Y. and Liu, Y. R. 2008. Potential of Pteris vit-
tata L. for phytoremediation of sites co-contaminated with
cadmium and arsenic: The tolerance and accumulation. J.
Environ. Sci. 20: 62–67.
Xiong, J. B., Wu, L. Y., Tu, S. X., Van Nostrand, J. D., He, Z. L.,
Zhou, J. Z. and Wang, G. J. 2010. Microbial communities
and functional genes associated with soil arsenic contamina-
tion and the rhizosphere of the arsenic-hyperaccumulating
plant Pteris vittata L. Appl. Environ. Microbiol. 76: 7277–
7284.
Xu, W. H., Kachenko, A. G. and Singh, B. 2010. Effect of soil
properties on arsenic hyperaccumulation in Pteris Vittata
and Pityrogramma Calomelanos var. Aus troa me ricana.Int.
J. Phytorem. 12: 174–187.
Xue, D. W., Chen, M. C. and Zhang, G. P. 2009. Mapping of
QTLs associated with cadmium tolerance and accumulation
during seedling stage in rice (Oryza sativa L.). Euphytica.
165: 587–596.
Xue, J. H. and Fei, Y. X. 2006. Effects of intercropping Cunning-
hamia lanceolata in tea garden on contents and distribution
488 Y. T. TANG et al.
of heavy metals in soil and tea leaves. J. Ecol. Rural Envi-
ron (in Chinese). 22: 71–73, 83.
Yadav, S. K., Dhote, M., Kumar, P., Sharma, J., Chakrabarti,
T. and Juwarkar, A. A. 2010. Differential antioxidative en-
zyme responses of Jatropha curcas L. to chromium stress.
J. Hazard. Mater. 180: 609–615.
Yadav, S. K., Juwarkar, A. A., Kumar, G. P., Thawale, P. R.,
Singh, S. K. and Chakrabarti, T. 2009. Bioaccumulation
and phyto-translocation of arsenic, chromium and zinc by
Jatropha curcas L.: Impact of dairy sludge and biofertilizer.
Bioresour. Technol. 100: 4616–4622.
Yan, X. L., Chen, T. B., Liao, X. Y., Huang, Z. C., Pan, J.
R., Hu, T. D., Nie, C. J. and Xie, H. 2008. Arsenic trans-
formation and volatilization during incineration of the hy-
peraccumulator Pteris vittata L. Environ. Sci. Technol. 42:
1479–1484.
Yang, B., Zhou, M., Shu, W. S., Lan, C. Y., Ye, Z. H., Qiu, R. L.,
Jie, Y. C., Cui, G. X. and Wong, M. H. 2010. Constitutional
tolerance to heavy metals of a fiber crop, ramie (Boehme-
ria nivea), and its potential usage. Environ. Pollut. 158:
551–558.
Yang, X. E., Long, X. X., Ye, H. B., He, Z. L., Calvert, D. V. and
Stoffella, P. J. 2004. Cadmium tolerance and hyperaccumu-
lation in a new Zn hyperaccumulating plant species (Sedum
alfredii Hance). Plant Soil.259: 181–189.
Yu, H., Wang, J. L., Fang, W., Yuan, J. G. and Yang, Z. Y.
2006. Cadmium accumulation in different rice cultivars and
screening for pollution-safe cultivars of rice. Sci. Total En-
viron. 370: 302–309.
Yu, H., Yang, Z. Y., Yang, Z. J. and Xiang, Z. X. 2008. Chemical
forms and subcellular and molecular distribution of Cd in
two Cd-accumulation rice genotypes. Chin.J.Appl.Ecol.
19: 2221–2226.
Zhang, G. P., Motohiro, F. and Hitoshi, S. 2002a. Influence of
cadmium on mineral concentrations and yield components
in wheat genotypes differing in Cd tolerance at seedling
stage. Field Crop. Res.77: 93–98.
Zhang, G. P., Motohiro, F. and Hitoshi, S. 2000. Genotypic
differences in effects of cadmium on growth and nutrient
compositions in wheat. J. Plant Nutr.23: 1337–1350.
Zhang, W. H., Cai, Y., Tu, C. and Ma, L. Q. 2002b. Arsenic
speciation and distribution in an arsenic hyperaccumulating
plant. Sci. Total Environ. 300: 167–177.
Zhao, F. J., Dunham, S. J. and McGrath, S. P. 2002. Arsenic
hyperaccumulation by different fern species. New Phytol.
156: 27–31.
Zheng, N., Wang, Q. and Zheng, D. 2007. Health risk of Hg, Pb,
Cd, Zn, and Cu to the inhabitants around Huludao Zinc
Plant in China via consumption of vegetables. Sci. Total
Environ. 383: 81–89.
Zhou, D. M., Wang, Q. Y. and Cang, L. 2011. Free Cu2+ ions,
Cu fractionation and microbial parameters in soils from ap-
ple orchards following long-term application of copper fungi-
cides. Pedosphere. 21: 139–145.
Zhu, Y., Yu, H., Wang, J., Fang, W., Yuan, J. and Yang, Z.
2007. Heavy metal accumulations of 24 asparagus bean cul-
tivars grown in soil contaminated with Cd alone and with
multiple metals (Cd, Pb, and Zn). J. Agr. Food Chem. 55:
1045–1052.
Zhuang, P., McBride, M. B., Xia, H., Li, N. and Li, Z. 2009.
Health risk from heavy metals via consumption of food crops
in the vicinity of Dabaoshan mine, South China. Sci. Total
Environ. 407: 1551–1561.
Zou, X. J. 2008. Immobilization and re-vegetation of heavy
metals polluted soils in Dabao Mountain by amendments.
Ph.D. Dissertation, Sun Yat-Sen University.
Zwonitzer, J., Pierzynski, G. M. and Hettiarachchi, G. M. 2003.
Effects of phosphorus additions on lead, cadmium, and zinc
bioavailabilities in a metal-contaminated soil. Water Air
Soil Pollut.143: 193–209.
... However, much amount of them is stored in roots compared to shoots [30], this finding was consistent with our result. The reason for it could be attributable to the roots have many transition metal transport proteins that improve the ability of Cd 2+ extraction, but the process of Cd transport to the shoot was affected by factors such as Cd concentration in the culture medium, and the Cd transport to the shoot was much lower than the absorption amount of root [31], Previously study has shown that intercropping between heavy metal hyperaccumulators and cash crops may reduce the concentration of Cd in food crops [32]. This finding was consistent with our result, which intercropping with Solanum nigrum L. inhibited the absorption of Cd by the roots of wheat and reduced the Cd contents of wheat shoots and roots. ...
Article
Full-text available
In order to evaluate the effect of intercropping of the hyperaccumulator, Solanum nigrum L. and wheat on the absorption and accumulation of cadmium (Cd) in wheat. The experiment was conducted with three replicates, which conducted on four Cd concentrations (0, 20, 40, 60 μmol L−1) in the Hoagland solution and using two planting patterns [monoculture wheat (MW), intercropping wheat and Solanum nigrum L. (IW,IS)]. The results showed that the addition of Cd in the solutions reduced the total root length by 19.08–55.98%, total root area by 12.35–44.48%, and total root volume by 16.01–46.00% of wheat plants. Intercropping with Solanum nigrum L. significantly reduced Cd contents by 28.3–47.2% and Cd accumulations by 10.08–32.43% in the roots of wheat. Transmission electron microscope (TEM) revealed that the root-tip cells of the monoculture wheat treated with Cd exhibited swollen spheres of intracellular mitochondria, disorderly arranged inner ridges of mitochondria, some damaged mitochondrial membranes, and deformed nuclear membranes. Many dense electron particles in the form of Cd were deposited in the cell gap, and the cell nucleus became smaller or even disappeared. Under the same Cd concentrations, root-tip cells of intercropped wheat showed much less density of electron particles, starch granules, and the damage to the nucleus and nuclear membrane by Cd.These results indicated that intercropping with Solanum nigrum L. reduced the Cd toxicity to wheat roots and decreased Cd uptake and accumulation in both the shoots and roots of the wheat.
... The current trend is that the number of polluted sites that require urgent action doubles every 25 years [40]. The scarcity of soil resources as the human population continuously increases will inevitably force more farmers to cultivate in contaminated areas [41]. The development of effective, low-cost soil remediation strategies is, thus, an urgent matter, especially in developing countries where few affordable options exist for the remediation of the vast extensions of polluted agricultural land. ...
Article
Full-text available
Soil pollution is a threat to food security and ecological and human health. Cd is one of the most common pollutants in agricultural soil and, due its human toxicity, one of the most hazardous. Amaranth is a documented hyperaccumulator of Cd and other pollutants, and it is commonly grown in Asia and South America. A considerable amount of amaranth is grown in suboptimal conditions, including nutrient-poor acidic soils. The objective of this experimental study was to examine the capacity of Amaranthus hypochondriacus to extract Cd from a nutrient-poor, acidic substrate that was spiked with different concentrations of Cd (2 and 20 mg kg−1 dw) during a period of 180 days. The plants grown in the substrate that was spiked with 20 mg Cd kg−1 dw did not develop into mature plants, but the plants grown in substrate that was spiked with 2 mg Cd kg−1 dw extracted a significant amount of Cd from the substrate by accumulating it into the above-ground biomass. The Cd levels varied from 113 to 176 mg kg−1 in the stems at the four measuring points, and from 64 to 94 mg kg−1 in the leaves. The concentrations in the plants increased with time and reached a maximal concentration of 176 ± 45 mg kg−1 dw for stems and 94 ± 41 mg kg−1 dw for leaves after 180 days. The mean bioaccumulation factor in the plants was 86 ± 15 after 90 days, 72 ± 12 after 120 days, 105 ± 37 after 150 days, and 99 ± 31 after 180 days, which confirms the previously reported capacity of Amaranthus hypochondriacus to hyperaccumulate Cd. Amaranthus hypochondriacus may, thus, be used to improve ecological and human health by remediating moderately Cd-polluted soils, even in nutrient-poor acidic soils.
Article
Full-text available
In recent years, there has been an increase in interest in the accumulation of heavy metals and metal(loid)s (HMM) in areas where agriculture and mining exist side by side. As a contribution to this body of knowledge, we report the first dataset into HMM concentrations in food crops and agricultural soils in Moquegua, which is a typical mining region and contains one of Peru’s largest copper deposits. Thanks to its geographic diversity, samples were taken in different agroecological regions at altitudes between 9 and 3,934 m. For food crops, 31 elements were measured using inductively coupled plasma mass spectrometry and atomic absorption spectrometry. For soils, 23 elements were measured using inductively coupled plasma optical emission spectrometry. Thus, the dataset includes a total of 13,828 observations from 341 sampling sites. We hope that this dataset will facilitate a wide range of agricultural and food safety studies, as well as serving as a reference for monitoring changes in pollution over time or comparing HMM levels with other farmlands influenced by mining activities.
Article
Full-text available
Cadmium (Cd) contamination of soils is an environmental concern, as cadmium harms food crops and can therefore impact human health. The use of combinations of biochar (seeded with Rhizobium leguminosarum) and Vigna radiata (as an intercrop) has the potential to reduce the mobilization of Cd from soil via mustard plants (Brassica juncea). Mustard plants are grown as a food and oil production crop that is consumed worldwide. However, this plant has the property of hyperaccumulation; thus, it bioaccumulates Cd in its tissues, which in turn, if eaten, can become part of the human food chain. Hence, reducing Cd bioaccumulation in mustard plants is crucial to making these plants a reliable and safe source of food for consumption. To improve soil sorption capacity and immobilization efficiency, biochar (in the form of wheat husk) was mixed with R. leguminosarum and intercropped (using V. radiata) with mustard plants for further investigation. Sampling was performed at an early growth stage (i.e., at 30 days) and at maturity (i.e., at 60 days) to determine the impact of Cd on a plant’s morphophysiological attributes. Data were analyzed in two ways: first by analysis of variance (ANOVA) and then by the post hoc Tukey’s honestly significant difference (HSD) test. The statistical analysis concluded that combinations effectively improved plant traits by 65%–90% in the early growth stage and by 70%–90% in the maturity stage. The T6 treatment combination [i.e., biochar + R. leguminosarum + V. radiata (BC + RL + VR)] provided the most effective results in terms of growth, biomass, pod yield, and pigmentation content. In addition, this combination reduced the translocation of Cd in mustard plants by 70%–95%. The combination of BC + RL + VR effectively reduced Cd contamination of mustard tissue and provided a suitable growing environment for the plants. A post-harvesting soil analysis using X-ray diffraction (XRD) found that Cd was undetectable in soil. This provides clear confirmation that these approaches can lead to Cd soil remediation. Moreover, this study provided insight into the responses of different morphophysiological attributes of mustard plants to Cd stress and could aid in developing Cd stress tolerance in mustard plants.
Article
Effects of intercropping Cunninghamia lancealata in tea gardens on contents and distribution of heavy metals in the soil an tea leaves were analyzed, through comparison study of a mono-cropping tea garden and an intercropping tea garden. Results show (1) contents of Pb, Ni, Mn and Zn decreased in the intercropping tea plantation, and the absorption capacity of Cunninghamia lanceolata varied from element to element, displaying an decreasing order of Mn > Ni > Pb > Zn; (2) concentrations of the heavy metals in the tea leaves of the intercropping tea garden were significantly lower than in the mono-cropping one. So intercropping of Cunninghamia lanceolata in tea gardens can significantly decrease contents of Pb, Ni, Mn and Zn and improve quality of the tea.
Article
The effects of chemical amendments on heavy metals immobilization of contaminated soils (agro-discard soil and acid mine-tailing soil) were determined. The effectiveness of limestone, zeolite and fly ash to reduce the metal availability to pea on the two types of heavily contaminated soils was evaluated. The growth and uptake of heavy metals by pea in amended soils were studied in pot experiments under greenhouse conditions. The application of both limestone and fly ash increased the soil pH. All tested amendments decreased the metal availability in both the soils and decreased the heavy metals uptake by pea compared with the unamended control. Pea failed to establish in acid mine tailing control and the zeolite treatments, but grew in agro-discard control and other treatments. The biomass of pea increased with a factor 2.4 (soil amended with limestone), 2.6 (soil amended with zeolite), 2.9 (soil amended with zeolite and limestone), 2.7 (soil amended with fly ash) and 2.9 (soil amended with fly ash and limestone) with respected to unamended soil. Fly ash appeared more efficient than other amendments in reducing plant metal concentrations, and the mixture of amendments showed more efficient than they were applied alone. Cd, Cu, Pb and Zn concentration in pea under fly ash plus limestone treatment decreased by 94.6%, 69.3%, 80.6% and 87.3%, respectively, compared with the control. Soil chemical treatment improved the performance of crops by increasing soil pH, releasing nutrient elements to soils, and reducing bioavailability of metals in the soils. The results suggest that fly ash and limestone can be used to remediate heavily contaminated acid soils to allow re-vegetation.