Anda di halaman 1dari 14

Journal of Geochemical Exploration 81 (2004) 1 14

www.elsevier.com/locate/jgeoexp

Arsenic in soil and vegetation of contaminated areas in


southern Tuscany (Italy)
F. Baroni a, A. Boscagli b, L.A. Di Lella a, G. Protano a, F. Riccobono a,*
a

Dipartimento di Scienze Ambientali, Sezione di Geochimica Ambientale, University of Siena, Via del Laterino 8, I-53100 Siena, Italy
b
Dipartimento di Scienze Ambientali, Sezione di Botanica, University of Siena, Via Mattioli 4, I-53100 Siena, Italy
Received 27 February 2003; accepted 16 June 2003

Abstract
Arsenic contents of soils and higher plants were surveyed in two former Sb-mining areas and in an old quarry once used for
ochre extraction. Total As in the soils ranged from 5.3 to 2035.3 mg kg 1, soluble and extractable As from 0.01 to 8.5 and from
0.04 to 35.8 mg kg 1, respectively. The As concentrations in the different fractions of soil were correlated significantly or very
significantly. Sixty-four plant species were analyzed. The highest As contents were found in roots and leaves of Mentha
aquatica (540 and 216 mg kg 1, respectively) and in roots of Phragmites australis (688 mg kg 1). In general, the As contents
of plants were low, especially in crops and in the most common wild species. In the analyzed species, roots usually showed the
highest content followed by leaves and shoots. Arsenic levels in soils and plants were positively correlated, while the ability of
the plants to accumulate the element (expressed by their Biological Accumulation Coefficients and Concentration Factors) was
independent of the soil As content. Comparison with the literature data, relationships between the As contents in plants and
soils, and biogeochemical and environmental aspects of these results are discussed.
D 2003 Elsevier B.V. All rights reserved.
Keywords: Arsenic; Plant accumulation; Mining area; Soil contamination

1. Introduction
The average arsenic content in the Earths crust
(clarke) was estimated by Greenwood and Earnshaw
(1984) to be as high as 1.8 mg kg 1. A rather similar
value of 1.5 mg kg 1 was suggested by Wedepohl
(1970) for igneous rocks on the basis of the average
values of granites, basalts and gabbros. Decidedly
higher As values were detected in sedimentary rocks
and a value as high as 13 mg kg 1 (Wedepohl, 1970)

* Corresponding author. Fax: +39-577-233945.


E-mail address: riccobono@unisi.it (F. Riccobono).
0375-6742/$ - see front matter D 2003 Elsevier B.V. All rights reserved.
doi:10.1016/S0375-6742(03)00208-5

appears appropriate for clayey rocks such as shales.


Since As accumulates during weathering and translocation in colloid fractions, its concentration is usually
higher in soil than in parent rocks (Yan-Chu, 1994). In
nature, the element is a fundamental constituent of the
sulfide mineral arsenopyrite (FeAsS), as well as the
minerals lollingite (FeAs), realgar (AsS) and orpiment
(As2S3).
The anthropogenic contribution to As contents of
superficial geochemical environments is also important (Nriagu and Pacyna, 1988; Pacyna et al., 1995).
In some cases, mining activity is directly involved in
the release of huge stocks of arsenic into superficial
environments (Murdoch and Clair, 1986).

F. Baroni et al. / Journal of Geochemical Exploration 81 (2004) 114

Arsenic exists in soil mainly as pentavalent (AsV)


arsenate (AsO43 ), which makes up 90% of dissolved
As species in aerobic soils (ONeill, 1995), or as
trivalent (AsIII) arsenite (AsO2 ). The latter represents
the most environmentally dangerous form of As.
The chemical similarity of arsenate to phosphate
and the high affinity of arsenite with sulfhydryl groups
of enzymatic and structural proteins are the ultimate
reasons for arsenic toxicity to living organisms.
The arsenate anion is rather easily chemisorbed by
soil colloids and, deriving from the strong arsenic acid
H3AsO4, adsorbs most effectively at low pH. Consequently, arsenate mobility is quite low in acidic soils,
especially where high contents of clay or metal oxide
are involved. Conversely, in alkaline soils, As may be
mobile in the soluble Na-arsenate form (McBride,
1994). Thus, adsorption of arsenate onto soil particles
is dependent on various parameters, but mostly pH.
According to Elkhatib et al. (1984), the pH, the
redox conditions and the Fe-oxide content in soil are
the most important features controlling AsIII adsorption. The element has a rather long residence time in
soils (from 1000 to 3000 years; Bowen, 1979) and
tends to be enriched into top horizons by cycling
through vegetation, atmospheric deposition and sorption by soil organic matter (Alloway, 1990). Its
availability for uptake by plants is affected by several
factors, such as the source, chemical speciation and
soil parameters (pH, Eh, organic matter and colloid
contents, soil texture and drainage conditions; Eisler,
1994; Mitchell and Barr, 1995).
The mobility of arsenic in aqueous solutions increases in the trivalent oxidation state (Hermann and
Neumann-Mahlkau, 1985). However, the conversion
of AsV to AsIII appears rather slow, even under
strongly reducing conditions, as suggested by observations on a range of redox and pH environments
(Masscheleyn et al., 1991).
In terrestrial plants, arsenic uptake is largely species specific and arsenic concentrations in plant tissues generally cannot be related to those in the soils
(ONeill, 1995). In this regard, it is relevant for human
health that As levels in edible plants are usually low,
even when they grow on contaminated soils (NRC,
1977; MAFF, 1982).
Kabata-Pendias and Pendias (1984) reported that
the As background for terrestrial plants growing on
uncontaminated soils ranges from 0.009 to 1.5 mg

kg 1 on a d.w. basis. Some species of the genus


Agrostis growing on contaminated soils have been
found to accumulate and tolerate high As levels: up to
6640 mg kg 1 d.w. in the old leaves of A. canina and
A. tenuis (Porter and Peterson, 1975), 1350 mg kg 1
in Agrostis stolonifera (Porter and Peterson, 1977a),
1900 mg kg 1 in Agrostis castellana and 1800 mg
kg 1 in Agrostis delicatula (de Koe et al., 1991; de
Koe, 1994). Pseudosuga taxifolia, Pityrogramma calomelanos and Pteris vittata growing on soils of
mineralized or contaminated areas were even more
able to accumulate As, showing contents of 8200,
8350 and 7526 mg kg 1, respectively (Warren et al.,
1968; Ma et al., 2001; Visoottiviseth et al., 2002).
Aquatic plants such as Ceratophyllum demersum,
Egeria densa and Potamogeton pectinatus accumulated arsenic up to 1160, 1120 and 4990 mg kg 1,
respectively, without any apparent damage (Dushenko
et al., 1994; Robinson et al., 1995). In contrast,
wetland plants appear unable to accumulate As to
the same extent (Otte et al., 1990; Qian et al., 1999).
The highest concentrations have been found in Spartina alterniflora (550 mg kg 1) by Carbonell et al.
(1998) and in Eichornia crassipes (500 mg kg 1) by
Zhu et al. (1999).
Arsenic is not an essential element for plants, and
once it is taken up, usually only a small proportion is
translocated to the epigeal parts. The result is the
following concentration gradient: roots>stems>leaves.
Nevertheless, concentrations up to 2000, 22,630 and
8350 mg kg 1 were found in foliage of Agrostis
ecotypes, P. vittata and P. calomelanos, respectively,
by Porter and Peterson (1975, 1977a,b), Ma et al.
(2001) and Visoottiviseth et al. (2002).
Inside the plant cell, the two most common chemical species (arsenite and arsenate) strongly induce
phytochelatin synthesis, which has an important role
in detoxification (Schmoger et al., 2000).
This paper deals with the impact on vegetation of
As diffusion into the environment related to different
forms of historical mining activities.

2. The study areas


Since the pre-Roman Age, southern Tuscany has
been one of the few important mining districts in Italy.
Epithermal deposits of Hg and Sb were intensely

F. Baroni et al. / Journal of Geochemical Exploration 81 (2004) 114

exploited until the 1970s. These intensive mining and


smelting activities produced huge quantities of waste
materials which, in the absence of any reclamation,
still release toxic heavy elements into the surrounding
environment (Protano and Riccobono, 1997; Protano
et al., 1998). Since As is a minor but ubiquitous
constituent of the epithermal mineral assemblage of
these ores (Riccobono, 1993), this element also contributes to the overall pollution of the mining areas.
High resolution geochemical maps of southern Tuscany (Protano et al., 1998, 1999) have shown that
strong, extensive As anomalies are present in various
areas of this region where base metal and Sb Hg
epithermal deposits were exploited in the past.
In addition to the abandoned mining and smelting
areas, other situations with severe arsenic contamination are known from this region. In the Mt. Amiata
volcanic massif area in the southeastern corner of the
region, there are numerous deposits of yellow-brown
ochre, employed as a dye since the Etruscan Age and
intensely exploited until very recently. These deposits
consist of horizons (3 4 m thick) belonging to the
sedimentary sequence of Quaternary lacustrine basins.
The ochre exhibits extremely high arsenic contents,
ranging (if expressed as As2O3) from 0.59% to 9.04%
by weight (Cipriani et al., 1971). Such high figures are,
at least in part, due to the diffuse presence of iron
arsenates (most probably poorly crystalline FeAsO4
2H2O or scorodite).
Fig. 1 shows the location of the three areas of
southern Tuscany (with expected anomalous As contents in the soil) where we chose to investigate the
transfer of this element from soil to plant species.
Area A is in the Sb-mining district of the Tafone
Valley in the Monti Romani area (Protano and Riccobono, 1997; Baroni et al., 2000a). The most important
antimony mines of the region were active there,
together with a plant for Sb-ore smelting and the
production of antimony trioxide.
Area B refers to the old antimony mine of Cetine di
Cotorniano near the town of Siena, which was mainly
exploited in the period between the two World Wars.
More than half a century after closure, the mine dumps
and the roasting-plant area are now largely colonized
by several herbaceous plant and shrub species.
Area C is located in the volcanic massif of Mt.
Amiata near the town of Castel del Piano (Grosseto
County). A quarry was active there for the exploita-

Fig. 1. Location of the study areas in southern Tuscany (see text).

tion of an extensive ochre horizon. The central and


deepest part of the quarry is now occupied by a very
small lake, while the surroundings are colonized by
wild vegetation.
The general geology and lithological features of
areas A and B are not identical but rather similar. A
Paleozoic metamorphic basement, mainly composed
of quartzites and metasandstones, is unconformably
covered by a Triassic succession. The lowest carbonate member (Cavernous Limestone) of the Mesozoic
sequence was widely affected by strong hydrothermal
alterations, which mostly produced limestone silicization, and hosts stibnite (Sb2S3)-rich ore bodies.
Area C lies on the lava flows, mostly riodacitic in
composition, present on the western slope of the
inactive Mt. Amiata volcano. The ochre horizons are

F. Baroni et al. / Journal of Geochemical Exploration 81 (2004) 114

often interlayered with very thin clayey and diatomite


levels, sometimes bearing lenses of yellow opal
(Cipriani et al., 1971). Analyses of ochre samples
revealed very high contents of iron (up to 65% Fe by
weight) and arsenic (up to 9% As expressed as As2O3
by weight; Cipriani et al., 1971). The genesis of these
As-rich ochre horizons has still not been satisfactorily
explained. However, according to Carobbi and Rodolico (1976), both inorganic and organic processes
were involved in the formation of this quite peculiar
type of rock.

3. Sampling surveys
Plant and soil samples were collected in 1996,
1997 and 1998 from the three localities described
above. Wild plant species were usually sampled,
except in area A where specimens of cultivated plants
were also collected.
Sampling was carried out in area A in the surroundings of the small lake in the open pit of the old
Tafone Mine and along a length of the Chiarone
stream where mining works were intensive. In this
area, six sampling sites were established (S1 S6). S1
and S2 were located in cultivated fields, S3 in a
vegetable garden, S4 in an old field, S5 on mine
dumps and S6 at a mineral processing tailing pond.
In area B, plant specimens were collected from the
dumps and from the area used to roast sulfide minerals. In area C, plant species were sampled from the
slopes of a quarry and from adjacent pastures.
Plant and soil samples were collected according to
the following criteria. Cultivated plant species were
selected to represent the ones most commonly sown in
the sampling areas. Wild species were selected on the
basis of their potential use as phytoremediators,
according to their:

adaptation to adverse edaphic conditions


large biomass allocation in the above-ground part
fast maximization of above-ground biomass
frequency in the field

At least three specimens of each selected species


were collected for analysis at each sampling site. In
species where roots were difficult to collect, only the
shoots were collected.

From four to nine soil samples, representative of


the top 20 cm, were collected at each site.
The nomenclature of the plant species is according
to Pignatti (1982).

4. Materials and methods


4.1. Analysis of plant tissues and soil
Plant material was carefully washed in tap water
and then processed in an ultrasonic cleaner to remove
soil particles. This was followed by a rinse with
acidified deionized water (HCl 3%) and a final rinse
with ultra-pure water.
Absorbing (roots) and non-absorbing parts of
plants (leaves, shoots, inflorescence, etc.) were usually separated in order to obtain information about the
species ability to transfer As. In some species,
especially those of Gramineae, the leaves and shoots
were analyzed together because of the difficulty in
separating them. The plant material was oven-dried at
40 jC to constant weight, then pulverized. About 0.5
g of dry matter was digested with a Milestone microwave lab-station (Ethos 900) after addition of 5 ml
HNO3 and 2 ml H2O2 (Baker ultra-pure reagents).
Soil samples were air dried, sieved through a 2-mm
mesh and pulverized in an agate mortar. They were
then subjected to acid digestion for determination of
total As content: 1 ml HF + 2 ml HNO3 + 2 ml HCl + 1
ml HClO4 were added to 200 mg of powdered soil and
the mixture was processed with an Ethos 900 labstation. An estimate of the phytoavailable As content
was obtained by both pure water and gentle acidic
extraction. Water extraction was performed by adding
100 ml of ultra-pure water to 40 g of soil and shaking it
for 24 h; acidic extraction was carried out by adding
200 ml of a 0.43 mol solution of acetic acid to 5 g of
soil and shaking it for 16 h (according to the procedure
of Ure et al., 1993). Polyethylene bottles were used to
collect and store the solution after the extraction.
Atomic absorption determination of As contents
was carried out with a Perkin-Elmer 5000 AAS,
equipped with FIAS, employing the hydride generation method. Working standards for chemical analyses
were prepared from Perkin-Elmer stock solutions.
Reference standards were SV-M (soil) from the Institute of Radioecology and Applied Nuclear Techniques

F. Baroni et al. / Journal of Geochemical Exploration 81 (2004) 114

(IRANT), Kosice (Slovakia), 2709 San Joaquin soil


from NIST, BCR 60 (aquatic plant) from the Bureau
of Reference of the EU and GBW-07603 (bush
branches and leaves) from the Institute of Geophysical
and Geochemical Exploration of Langfang (China).
Accuracy of the analytical results was within 7%.
The cation exchange capacity (CEC) was determined by the compulsive exchange procedure suggested by Gillman and Sumpter (1986); extractable
phosphorus was measured by UV Visible spectrophotometry (Hach DR-4000) following the Hach
8190 method; the pH was measured in deionized
water (soil/water ratio = 1:2.5 w/v). Organic matter
contents were estimated through the percentage loss
in weight after ignition at 375 jC for 16 h in a furnace
(Storer, 1984).
4.2. Data analysis
The analytical results were used to check the
relationships between: (i) total, soluble and extractable soil arsenic; (ii) soil and plant As contents.
Two coefficients, Biological Absorption Coefficient (BAC) and Concentration Factor (CF), were
also considered. The former is the ratio of plant
arsenic to total soil arsenic (Edwards et al., 1998),
while the latter is the ratio of plant arsenic to soluble
or extractable arsenic in the soil. The CFs for soluble
and extractable soil arsenic were designated as CFsol

and Cfextr, respectively. Relationships were identified


with the non-parametric Spearman Rank Correlation
Coefficient.
The ability of the plant species to translocate
arsenic from the roots to shoots was also tested. The
palatability of each plant species to livestock (primarily to sheep, the most common livestock in the study
areas) was considered according to Daget and Poissonet (1971) and Sostaric-Pisacic and Kovacevic
(1974).

5. Results
5.1. Soils
In the Tafone Chiarone district (area A), the total
arsenic contents in soils of the cultivated and uncultivated fields averaged from 5 to 40 mg kg 1 d.w. The
mine dumps and the tailing ponds showed mean values
of 266 and 1226 mg kg 1, respectively (Table 1).
However, the arsenic content was highly variable,
ranging from 1.3 to 55 mg kg 1 in the fields, from
38 to 899 mg kg 1 in the dumps and from 2 to 2466 mg
kg 1 in the tailings. Similar contents were found in the
soils above the mine dumps of area B and high values
were recorded in the quarry slopes and the pastures of
area C, where a mean content exceeding 2000 mg kg 1
and a range from 1037 to 3133 mg kg 1 were found.

Table 1
As content and some edaphic parameters of soils (mean F S.E.). In each column, the values followed by the same letter are not significantly
different (Tukey test; p < 0.05)
Sampling
sites

As (total)
(mg kg 1)

As (soluble)
(mg kg 1)

As (extractable)
(mg kg 1)

Organic matter
(%)

pH

Pavailable
(mg kg 1)

CEC
(meq/100 g)

Area A
S1 (n = 4)
S2 (n = 4)
S3 (n = 4)
S4 (n = 5)
S5 (n = 9)
S6 (n = 6)

14.60 F 1.00a
5.30 F 2.17b
39.90 F 2.85c
40.00 F 3.81c
265.60 F 88.26cd
1225.60 F 351.70d

0.01 F 0.01a
0.01 F 0.01a
0.02 F 0.01a
0.02 F 0.01a
0.01 F 0.01a
0.04 F 0.02a

0.10 F 0.03a
0.08 F 0.03a
0.04 F 0.02a
0.08 F 0.02a
0.66 F 0.23b
1.50 F 0.33c

4.37 F 0.66a
4.82 F 0.80a
5.25 F 1.27a
4.59 F 0.30a
3.03 F 0.40ab
1.68 F 0.39b

5.8 F 0.2a
5.5 F 0.1a
5.4 F 0.3a
5.0 F 0.1a
7.3 F 0.2ab
7.2 F 0.5ab

0.13 F 0.07a
0.15 F 0.05a
0.18 F 0.12a
0.11 F 0.10a
0.14 F 0.03a
0.11 F 0.02a

5.34 F 1.39a
5.51 F 1.08a
5.97 F 0.42a
5.19 F 1.54a
5.83 F 1.54a
4.06 F 0.96a

372.83 F 77.25

0.03 F 0.02

0.52 F 0.18

3.06 F 0.55ab

8.2 F 0.6b

0.08 F 0.04a

5.49 F 1.63a

753.82 F 34.55a
2035.32 F 297.12b

1.82 F 0.27a
8.48 F 1.40b

7.60 F 0.91a
35.80 F 2.98b

2.80 F 0.22ab
5.05 F 1.41a

5.8 F 0.7a
6.2 F 0.9ab

0.97 F 0.89bb
0.83 F 0.39bb

6.04 F 0.59a
6.66 F 1.04a

Area B
(n = 4)
Area C
Quarry slopes (n = 4)
Pasture (n = 4)

F. Baroni et al. / Journal of Geochemical Exploration 81 (2004) 114

Water-soluble arsenic gave figures from 2 to 5


orders of magnitude lower than the total As values.
In area A, the mean value (for each sampling site)
varied from 10 to 40 Ag kg 1 (see Table 1), while the
total range was from 2 to 70 Ag kg 1. In area C, the
water-soluble arsenic content was very high, especially in soils of pastures (mean: 8.5 mg kg 1).
There was a rather similar pattern for acid-extractable arsenic. The As contents reached the highest values in C but the extractable fraction still
represented a very small aliquot of the total As
(0.1 1.8%).
A relationship between As contents was only
evident in the case of fields (cultivated and old),
particularly between total and soluble As (r = 0.921;
p < 0.01; n = 17). However, when the high variability

of the data (the coefficients of variation averaged


53%) was smoothed by considering the mean of each
sampling site, the soluble and extractable fractions
were correlated with the total contents (r = 0.887 and
r = 0.862, respectively; p < 0.01; n = 9), and a relationship between the soluble and extractable fractions also
appeared (r = 0.680; p < 0.05; n = 9).
According to SISS (1985) criteria, only the soil of
the area A tailing ponds (S6) can be considered poor
in organic matter (Table 1). The pH was acidic in the
fields of area A (S1 S4) and area C, while it was
neutral or alkaline in the mine wastes of areas A (S5
S6) and B, respectively.
All sampling sites showed very low levels of available phosphorus in the soil, as well as low cation
exchange capacities (CEC). The correlation analysis

Table 2
Sampling area A. Arsenic content in crops and vegetables of cultivated sites (S1, S2, S3) and in plant species growing in the old field S4
(mean F S.E.; mg kg 1)
Sampling sites and
plant species

Leaves

S1
Zea mays
Triticum aestivum

< 0.02

S2
Helianthus annus
Medicago sativa

0.04 F 0.01

S3
Lactuca sativa
Solanum melangena
Cucurbita pepo
Capsicum annuum
Lycopersicon esculentum
S4
Rubus ulmifolius
Sonchus asper
Medicago hispida
Bromus hordeaceus
Bromus madritensis
Helichrysum italicum
Phalaris coerulescens
Avena fatua
Achillea ageratum
Brassica napus
Lupinus albus
Urospermum dalechampii
Coleostephus myconis
Rumex crispus
Anchus italica

Leaves and
shoots

Shoots

Roots

Flowers and
inflorescences

Fruits and
seeds
0.03 F 0.02
< 0.02

0.03 F 0.02
0.04 F 0.03

0.13 F 0.11
0.11 F 0.09
0.23 F 0.12
0.27 F 0.18
0.07 F 0.03

0.23 F 0.11
< 0.02
< 0.02

0.86 F 0.19
0.11 F 0.05
0.62 F 0.14
< 0.02
0.04 F 0.02
2.53 F 1.99
0.54 F 0.09
0.44 F 0.15
6.95 F 1.72
0.19 F 0.13
2.54 F 1.79
0.07 F 0.06
0.62 F 0.21
0.48 F 0.33
1.39 F 0.74

0.21 F 0.19
1.24 F 0.73
1.14 F 0.27

0.02 F 0.01
0.81 F 0.13
0.04 F 0.03
< 0.02
0.09 F 0.02

1.05 F 0.05
5.14 F 1.42
6.21 F 2.37
0.47 F 0.31
0.34 F 0.26
2.89 F 0.95
2.20 F 1.14
13.25 F 3.71
0.05 F 0.01
0.79 F 0.26

0.53 F 0.17
0.22 F 0.09
0.31 F 0.12
< 0.02
0.04 F 0.01

0.44 F 0.17
0.24 F 0.81
0.59 F 0.09
0.29 F 0.11
0.69 F 0.30

F. Baroni et al. / Journal of Geochemical Exploration 81 (2004) 114

showed that the As contents of soil (total, soluble and


extractable) were independent of CEC, pH, content of
available phosphorus and organic matter. All the
edaphic parameters were also independent of each
other.
5.2. Arsenic contents of plant species
In total, 64 plant species were sampled, 59 of them
in the Tafone Chiarone district (area A). The plant
species were mostly Dicotyledons (54), Compositae
(15), Leguminosae (12) and Gramineae (9), tap-rooted
(42), perennial (41) and herbaceous (54), largely as a
result of the sampling criteria. Some species, such as

Achillea ageratum, Dactylis hispanica, Helichrysum


italicum, Inula viscosa, Medicago sativa, Plantago
lanceolata and Silene vulgaris, were collected in two
or more sampling sites.
In area A, crops and vegetables sampled at sites S1,
S2 and S3 showed very low As concentrations,
sometimes below the limit of instrumental detection
(Table 2). In the old field (S4), plants had As contents
of 1 2 mg kg 1 or less and only the weed Coleostephus myconis showed slight accumulation, with 13
mg kg 1 of arsenic in its roots (Table 2). Only 5 (P.
lanceolata, Mentha aquatica, Galactites tomentosa,
C. myconis and A. stolonifera) of the 34 plant species
sampled on the mine dumps (S5) had at least 10 mg

Table 3
Sampling area A. Arsenic content in plant species growing on mine waste dumps S5 (mean F S.E.; mg kg
Plant species

Leaves

Achillea ageratum
Plantago lanceolata
Mentha aquatica
Galactites tomentosa
Silene vulgaris
Coleostephus myconis
Sylibum marianum
Trifolium pratense
Conyza bonariensis
Dorycnium hirsutum
Melilotus officinalis
Lepidium campestre
Dipsacus fullonum
Reichardia picroides
Ranunculus velutinus
Sinapis arvensis
Hedysarum coronarium
Inula viscosa
Hypericum perforatum
Ulmus minor
Sambucus ebulus
Cistus salvifolius
Trifolium incarnatum
Rosa canina
Dactylis hispanica
Medicago sativa
Lotus corniculatus
Sanguisorba minor
Spartium junceum
Foeniculum vulgare
Quercus ilex
Quercus cerris
Holoschoenus vulgaris
Agrostis stolonifera

1.81 F 0.52
9.35 F 1.70
216.35 F 19.36
4.40 F 1.62
5.82 F 1.09
10.85 F 4.52
5.19 F 2.07
0.50 F 0.13
7.70 F 3.28
3.11 F 0.09
0.67 F 0.27
1.44 F 0.35
0.24 F 0.09
1.51 F 0.68
2.14 F 0.88
1.67 F 0.41
0.47 F 0.15
2.97 F 1.31
1.23 F 0.08
0.16 F 0.05
0.59 F 0.16
1.29 F 0.37
0.38 F 0.09
0.39 F 0.21
0.16 F 0.04

Shoots

Leaves and shoots

0.07 F 0.04
37.44 F 9.26
1.15 F 0.07
2.54 F 1.74
2.40 F 0.08
2.16 F 0.49
0.08 F 0.03
2.54 F 0.72
3.24 F 1.04
0.13 F 0.05
0.15 F 0.07
0.09 F 0.02
0.72 F 0.28
0.44 F 0.13
0.06 F 0.02
0.37 F 0.11
0.06 F 0.04
0.87 F 0.09
0.14 F 0.03
0.97 F 0.26
0.22 F 0.07
0.89 F 0.39
0.19 F 0.05
< 0.02

Roots

Inflorescences

0.70 F 0.19
62.18 F 9.74
540.16 F 23.08
16.18 F 3.21
6.68 F 2.76
22.83 F 5.29
5.53 F 0.99
4.46 F 1.72
4.00 F 1.37
2.81 F 0.15
0.86 F 0.17
1.22 F 0.29
5.36 F 0.77
0.83 F 0.25
2.35 F 0.86
0.44 F 0.08
0.66 F 0.24
0.24 F 0.06
1.12 F 0.09

3.32 F 0.73

Seeds

4.50 F 1.08

0.12 F 0.04

0.79 F 0.16
0.21 F 0.04
0.09 F 0.04
0.34 F 0.05
2.81 F 1.04
0.09 F 0.07
0.09 F 0.01

< 0.02
0.08 F 0.03

3.34 F 0.58
0.99 F 0.15
0.17 F 0.02
3.32 F 0.35

0.10 F 0.03
0.05 F 0.01

2.55 F 0.27
< 0.02

6.87 F 1.73
10.13 F 1.08

F. Baroni et al. / Journal of Geochemical Exploration 81 (2004) 114

Table 4
Sampling area A. Arsenic content in plant species growing on mineral processing tailing ponds S6 (mean F S.E.; mg kg
Plant species

Leaves

Shoots

Roots

Achillea ageratum
Silene vulgaris
Plantago lanceolata
Phragmites australis
Dorycnium hirsutum
Aster squamatus
Atriplex patula
Inula viscosa
Melilotus alba

15.29 F 3.94
82.77 F 25.07
24.00 F 7.02
3.71 F 0.95
33.95 F 16.82

6.89 F 1.64
73.64 F 25.09

12.27 F 2.71
52.46 F 5.27
56.06 F 21.03
688.24 F 64.00
16.11 F 3.72
9.63 F 2.19
21.91 F 7.16
5.77 F 1.49
3.58 F 0.85

37.63 F 14.67
47.33 F 9.32
5.74 F 1.07

1.52 F 0.08
54.23 F 3.49
1.49 F 0.07
10.44 F 3.47
5.14 F 0.94
1.05 F 0.05

kg 1 of arsenic in their tissues, mostly in the roots


(Table 3). The M. aquatica specimens collected at the
shores of the small lake had rather high arsenic levels,
both in the leaves (mean: 216 mg kg 1) and roots
(mean: 540 mg kg 1).
Almost all the plant species growing in the tailing
ponds were sampled because of the small number
growing at this site. They generally had higher arsenic
contents than plants at the other sampling sites (compare Table 4 with Tables 2, 3 and 5). However, in five
of the nine collected species, the arsenic concentrations were < 50 mg kg 1 and only the roots of
Phragmites australis (mean: 688 mg kg 1) showed
arsenic accumulation.

Leaves

Shoots

Area B
Dactylis hispanica
Helichrysum italicum
Plantago lanceolata
Cichorium intybus
Calluna vulgaris
Hypericum perforatum
Reichardia picroides
Area C
Quarry slopes
Cytisus scoparius
Euphorbia cyparissias
Medicago sativa
Plantago lanceolata
Pasture
Silene dioica
Medicago sativa

)
Inflorescences
34.58 F 7.03

5.07 F 1.09
2.92 F 0.81

In the intensively sampled area A, there were low


or very low arsenic contents in most plant species
growing on the mine dumps, slopes and neighboring
dry badlands. This was especially true for the aboveground biomass (Table 6). In this area, I. viscosa and
Dorycnium hirsutum showed slightly higher arsenic
contents than other species.
Plant species common in wet habitats, such as M.
aquatica and P. australis, showed high arsenic contents mostly in roots but also in leaves (M. aquatica).
The above-ground tissues of the plant species most
palatable to livestock showed very low arsenic concentrations, except for the leaves of P. lanceolata (9
24 mg kg 1 on d.w.; Table 7).

Table 5
Sampling areas B and C. Arsenic content in plant species (mean F S.E.; mg kg
Sampling area and site

Rhizomes

)
Leaves and shoots
0.06 F 0.04
0.33 F 0.18
0.26 F 0.10
0.55 F 0.21
0.38 F 0.12
< 0.02
0.77 F 0.25

0.96 F 0.17

Roots
1.74 F 0.59
0.05 F 0.04
0.06 F 0.02

61.3 F 10.62
1.14 F 0.48
3.61 F 0.83

0.67 F 0.22
41.42 F 7.31
2.21 F 0.97
37.79 F 9.18

2.23 F 0.74

3.07 F 0.73
5.51 F 1.46

176.27 F 21.31

F. Baroni et al. / Journal of Geochemical Exploration 81 (2004) 114


Table 6
Arsenic content (mean or range; mg kg
Habitats and species

) in the more common plant species of the main habitats occurring in the sampling area A

Roots

Shoots

Mine dumps and slopes and neighbouring badlands


Bromus hordeaceus
Bromus madritensis
Dactylis hispanica
Inula viscosa
0.20 6.00
Helichrysum italicum
1.00
Spartium junceum
Rubus ulmifolius
Dorycnium hirsutum
3.00 16.00
Quercus cerris
Quercus ilex
Ulmus minor

Leaves

< 0.02
0.06 5.00

0.20
3.00 47.00
2.53
0.90

3.00 54.00

0.10

Table 7
Sampling area A. Arsenic content (mean or range; mg kg
leaves and shoots of the most pabular species for livestock

Avena fatua
Bromus hordeaceus
Dactylis hispanica
Hedysarum coronarium
Lotus corniculatus
Medicago sativa
Plantago lanceolata
Sanguisorba minor
Trifolium incarnatum
Trifolium pratense
Sonchus asper
Medicago hispida
Phalaris coerulescens
Brassica napus
Reichardia picroides
Cichorium intybus

Shoots

< 0.02
0.10
3.00

9.00 24.00
0.10
0.40
0.50

0.90
0.08
0.11
0.62
0.54

0.19
1.51

0.02
0.72

) in

Shoots and
leaves
0.40
< 0.02

0.20
0.50

Fruits

< 0.02

Very low arsenic contents were found in plant


species collected in area B ( < 1 mg kg 1 in almost
all the plants; Table 5) and area C. Exceptions were
Euphorbia cyparissias and P. lanceolata sampled on
the quarry slopes and notably Silene dioica from
pastures, which contained 176 mg kg 1 As in its
leaves.

Leaves

Shoots and leaves


< 0.02
0.04

3.00 34.00
0.05
0.10
0.20

Lake (ex open mine pit) shores and wet surfaces neighbouring to the tailing ponds
Agrostis stolonifera
Mentha aquatica
540.16
37.44
216.35
Phragmites australis
688.00
2.00
4.00

Species

0.77
0.55

< 0.02
3.00

10.00

5.3. Arsenic accumulation and translocation in plants


In general, the ability of the plants to accumulate
arsenic, as expressed by the Biological Accumulation
Coefficients (BACs) and Concentration Factors (CFs),
was independent of the As contents in the soils. These
indices were not correlated with total, soluble or
extractable arsenic in soil.
Nevertheless, when the plant part with the highest
As content was considered for each species, there
were significant relationships between arsenic in the
plant tissue and the total, soluble and extractable
arsenic in the soil (r = 0.635, 0.439 and 0.951, respectively; p < 0.001; n = 80).
With the exception of M. aquatica (BAC = 2.03),
the BACs of all the plant species were < 1; in 91% of
them, the BAC was < 10 1 and in 55% < 10 2. The
CFs for extractable arsenic were generally low: < 5 in
69% and < 20 in 86% of cases. The highest values
referred to the roots of the two wetland plants M.
aquatica (818) and P. australis (459).
Very high CFsol values were found in M. aquatica
(90,026), P. lanceolata (10,363) and P. australis
(17,206). They were also high in A. stolonifera
(1688), C. myconis (3805), Conyza bonariensis
(1283), D. hirsutum (1356), G. tomentosa (2697),
Holoschoenus vulgaris (1145), I. viscosa (1183) and
S. vulgaris (2069).

10

F. Baroni et al. / Journal of Geochemical Exploration 81 (2004) 114

In most of the sampled species, the roots had the


highest arsenic concentration (Tables 2 5). The translocation of As to shoots was low or very low, with
translocation coefficients (TC) < 2 in 76% of cases.
Three species showed enhanced As transport to shoots:
A. ageratum collected in the old field (TC = 14.8), I.
viscosa from the tailing ponds (TC = 8.2) and dumps
(TC = 12.4) of area A and S. dioica collected in the
pastures of area C (TC = 57.4).

6. Discussion
6.1. Relationships between As contents of soils and
plants
When compared to the As contents of soils at
several mining sites, where total As ranged from 2
to 17,000 mg kg 1 and available As from < 1 to 390
mg kg 1 (de Koe, 1994; Bech et al., 1997; Flynn et
al., 2002; Jung et al., 2002; Madejon et al., 2002;
Visoottiviseth et al., 2002), the values of our soils can
be considered intermediate or moderately low. Nevertheless, in some agricultural soils (pasture of area C
and part of the fields of area A), the arsenic contents
exceeded the Italian legal limits, even for garden or
park (20 mg kg 1) and industrial sites (50 mg kg 1;
DM 471/99, 1999).
The highest values of soluble and extractable As in
area C can be explained by the highest values of total
As in the soils. However, they may also be influenced
by the likely presence of the iron arsenate scorodite
and its stability relationships driven by Eh pH conditions (Vink, 1996).
Our results clearly showed that when the concentration of arsenic in the soil was not particularly high
(as in the fields and vegetable garden of area A), its
availability to plants strongly depended on the total
soil As content. Soluble and extractable As in soils
were positively correlated to the As contents in plants.
This relationship could be perceived throughout the
entire study areas when the mean concentrations for
each sampling site were considered.
This finding agrees with previous results for grass
growing near smelters (Temple et al., 1977), for
Urtica dioica and P. australis growing on experimental soils (Otte et al., 1990) and for Agrostis species
growing at mining sites (de Koe, 1994). We must

observe, however, that this relationship was not found


in other cases (ONeill, 1995; Pitten et al., 1999).
Perhaps a parallel change of As contents in plants and
As contents in soils through a gradient can be more
easily detected when several plant species are sampled
(as in our study).
In contrast, As translocation from the roots to the
above-ground biomass appeared to be under stronger
biological control than As uptake. This could explain
the lack of correlation between As contents of soils
and As concentrations in the epigeal parts. In this
respect, our results strongly agree with the literature
data (Otte and Ernst, 1994; ONeill, 1995).
When we compared the As contents in our plants
with those sampled at other mining sites by de Koe
(1994), Bech et al. (1997), Jung et al. (2002), Madejon et al. (2002) and Visoottiviseth et al. (2002), we
found much lower maximum values (from less than
one-half to less than one-tenth). Since As contents in
our soils were also lower in the most contaminated
situations, a weaker selective pressure can be inferred
or, simply, more effective accumulator plant species
may have escaped our sampling.
Neither the arsenic availability in soils nor its
concentration in plants reflected the differences in
organic matter content, pH and available phosphorus
in soils of the sampling sites. All these factors are
known to affect both bioavailability and plant uptake
of As (Otte et al., 1990; Bhumbla and Keefer, 1994).
Yet in our study, these factors had no apparent
influence.
6.2. Biogeochemical and environmental aspects
Plants contributed very little to the arsenic diffusion, cycling and transfer from soil to biosphere in the
study areas. Both the As concentrations in plant
tissues and the ability of plants to accumulate the
element were low or very low. This was especially so
in the most common crops (corn, lucerne, sunflower,
wheat), in common wild herbs (Bromus hordeaceus,
Bromus madritensis, D. hispanica, I. viscosa), in the
main chamaephytes (D. hirsutum and H. italicum) and
shrubby colonizers (Rubus ulmifolius and Spartium
junceum), as well as in the main components of the
local forest (Quercus cerris and Quercus ilex).
However, species such as M. aquatica, P. lanceolata, P. australis and some others growing on soils

F. Baroni et al. / Journal of Geochemical Exploration 81 (2004) 114

11

very rich in arsenic (especially in dumps and tailing


ponds of study area A) appeared to be good pumps for
soil As. Their CFs were high or very high, especially
for soluble soil As.
A low As concentration in the above-ground plant
biomass decreases the risk of food chain contamination through grazing. Indeed, with the exception of P.
lanceolata (9 24 mg kg 1 of As in its leaves), this
was the general case for the species most palatable to
livestock. However, P. lanceolata also had As contents below the maximum level (50 mg kg 1) tolerated by cattle, sheep and swine (Chaney, 1989). Thus,
plant contamination by As-rich soil particles (or soil
ingestion) could be the main route of As intake by
wild herbivores and livestock, as found elsewhere
(Thornton and Abrahams, 1983; Li and Thornton,
1993).
The As contents found in the edible parts of crops
and vegetables are not cause for concern, since they
were close to or below the instrumental detection
limit.

P. australis, with a mean As content of 688 mg


kg 1 in the roots, 1.5 mg kg 1 in the shoots and 3.7
mg kg 1 in the leaves, clearly confirmed its low
ability to translocate arsenic from roots to the
above-ground biomass (Otte et al., 1990). Reduced
metal transport from roots to shoots appears to be the
usual behavior of the species (see for instance Ye et
al., 1997).
A. stolonifera has been found to be an As accumulator in other mining areas, with contents up to
1350 mg kg 1 (Porter and Peterson, 1975), but this
was not the case in our study (maximum concentrations of 21 mg kg 1 were found). Nevertheless, it
must be stressed that Porter and Peterson also found
extremely high As contents in the soils they studied
(from 8510 to 26,530 mg kg 1).
The ability of M. aquatica to concentrate As and
translocate it to the leaves is rather interesting. Therefore, its potential use as a phytoremediator could be
assessed.

6.3. Remarks on some plant species

7. Conclusions

One of the sampled species, S. vulgaris, is well


known for its tolerance to several trace elements, such
as antimony, cadmium, cobalt, copper, lead, nickel
and zinc, which are even accumulated or hyperaccumulated (Harmens et al., 1993; de Knecht et al., 1995;
Wenzel and Jockwer, 1999; Baroni et al., 2000b).
With regard to the ability of S. vulgaris to accumulate
arsenic, our data agree with literature reports. Mean
values such as 261 mg kg 1 (Paliouris and Hutchinson, 1991) or 637 mg kg 1 (Sneller et al., 1999)
have been reported for soils with soluble As 25 37
times higher than the highest value (tailing ponds) in
our soils. Moreover, our plants had higher As contents
in the shoots than in the roots; this is an important
aspect since a substantial As depletion occurs at the
end of the growing cycle when most of the aboveground biomass is shed.
Rather high As contents were also found in S.
dioica growing in pastures of study area C. However,
in this case, the levels of soluble and extractable As in
the soil were high: 8.5 and 35.8 mg kg 1, respectively.
M. sativa sampled in area C, as well as in the old
fields and mine dumps of area A, behaved as an
excluder.

The soil levels of organic matter, available phosphorus, pH and CEC had no effect on soil As content
and its bioavailability to plants. Tissues of the 64 plant
species generally exhibited an As content positively
correlated to that of the soil. Nevertheless, the As
content in plants was always low, even in the most
contaminated conditions, with two exceptions: M.
aquatica and P. australis. In spite of the long contamination history of the surveyed areas, there is an evident
lack of effective pressure toward As tolerance by the
plant species through accumulation of the element.
With few exceptions, the As concentration was
higher in roots than in leaves and shoots. This
decidedly decreases the risk of food chain contamination via herbivores.
Arsenic concentrations were also low in the most
common herbaceous species (crops and wild plants),
in the main chamaephytic and shrubby colonizers and
in the main forest trees. This means that it is likely
that plants play a minor role in superficial geochemical cycling of arsenic. Nevertheless, the arsenic levels
above the legal limits in agricultural soils suggest that
a wider survey of As contents in crops, fodders and
vegetables should be carried out.

12

F. Baroni et al. / Journal of Geochemical Exploration 81 (2004) 114

Acknowledgements
This paper is a contribution to the research project:
Dispersion and transfer of metals to the biosphere in
mining areas, supported by Ministero dellUniversita`
e della Ricerca Scientifica e Tecnologica (MURST)
and the University of Siena.
References
Alloway, B.J., 1990. Soil processes and the behaviour of metal. In:
Alloway, B.J. (Ed.), Heavy Metals in Soils. Chapman & Hall,
Glasgow, pp. 7 28.
Baroni, F., Boscagli, A., Protano, G., Riccobono, F., 2000a. Antimony contents in plant species growing in an Sb-mining district
(Tuscany, Italy). In: Markert, B., Friese, K. (Eds.), Trace Elements in the EnvironmentTheir Distribution and Effects.
Elsevier, Amsterdam, pp. 341 361.
Baroni, F., Boscagli, A., Protano, G., Riccobono, F., 2000b. Antimony accumulation in Achillea ageratum, Plantago lanceolata
and Silene vulgaris growing in an old Sb-mining area. Environmental Pollution 109, 347 352.
Bech, J., Poschenrieder, C., Llugany, M., Barcelo, J., Tume, P.,
Tobias, F.J., Barranzuela, J.L., Vasquez, E.R., 1997. Arsenic
and heavy metal contamination of soil and vegetation around
a copper mine in Northern Peru. The Science of the Total Environment 203, 83 91.
Bhumbla, D.K., Keefer, R.F., 1994. Arsenic mobilization and bioavailability in soils. In: Nriagu, J.O. (Ed.), Arsenic in the Environment (Part I). Wiley, New York, pp. 51 82.
Bowen, H.J.M., 1979. Environmental Chemistry of the Elements.
Academic Press, London.
Carbonell, A.A., Aarabi, M.A., DeLaune, R.D., Gambrell, R.P.,
Patrick Jr., W.H., 1998. Arsenic in wetland vegetation:
availability, phytotoxicity, uptake and effects on plant growth
and nutrition. The Science of the Total Environment 217,
189 199.
Carobbi, G., Rodolico, F., 1976. I minerali della Toscana. Olschki,
Firenze.
Chaney, R.L., 1989. Toxic element accumulation in soils and crops:
protecting soil fertility and agricultural food-chains. In: BarYosef, B., Barrow, N.J., Goldshmid, J. (Eds.), Inorganic Contaminants in the Vadose Zone. Springer, Berlin, pp. 140 158.
Cipriani, C., Franzini, M., Malesani, P., Sabatini, G., 1971. I giacimenti di materiali litoidi. Rendiconti Societa` Italiana Mineralogia e Petrologia 27, 317 355 (special issue).
Daget, P., Poissonet, J., 1971. Une metode danalyse phytologique
des prairies. Annales Agronomiques 22, 5 41.
de Knecht, J.A., van Baren, N., Ten Bookum, W.T., Wong Fong
Sang, H.W., Koevoets, P.L.M., Schat, H., Verkleij, J.A.C.,
1995. Synthesis and degradation of phytochelatins in cadmiumsensitive and cadmium-tolerant Silene vulgaris. Plant Science
106, 9 18.
de Koe, T., 1994. Agrostis castellana and Agrostis delicatula on

heavy metal and arsenic enriched sites in NE Portugal. The


Science of the Total Environment 145, 103 109.
de Koe, T., Beek, M.A., Haarsma, M.S., Ernst, W.H.O., 1991.
Heavy metals and arsenic in grasses and soils of mine spoils
in North East Portugal, with particular reference to some Portuguese goldmines. In: Nath, B. (Ed.), Environmental Pollution.
Proc. Int. Conf., ICEP-1, vol. 1, pp. 373 380.
DM 471/99 (Decreto Ministeriale 25 ottobre 1999 n. 471). Regolamento recante criteri, procedure e modalita` per la messa in sicurezza, la bonifica e il ripristino ambientale dei siti inquinati ai
sensi dellarticolo 17 del D.L. 5 febbraio 1997 n. 22 e successive
modifiche e integrazioni. Gazzetta Ufficiale della Repubblica
Italiana, 15 dicembre 1999, 293, Supplemento Ordinario n.
293. 67 pp.
Dushenko, W.T., Bright, D.A., Reimer, K.J., 1994. Arsenic bioaccumulation and toxicity in aquatic macrophytes exposed to
gold-mine effluent: relationships with environmental partitioning, metal uptake and nutrients. Aquatic Botany 50,
141 158.
Edwards, S.C., MacLeod, C.L., Lester, J.N., 1998. The bioavailability of copper and mercury to the common nettle (Urtica
dioica) and the earthworm Eisenia foetida from contaminated
dredge spoil. Water Air Soil Pollution 102, 75 90.
Eisler, R., 1994. A review of arsenic hazards to plants and animals
with emphasis on fishery and wildlife resources. In: Nriagu, J.O.
(Ed.), Arsenic in the Environment (Part II). Wiley, New York,
pp. 185 259.
Elkhatib, E.A., Bennet, O.L., Wright, R.J., 1984. Kinetics of arsenite sorption in soils. Soil Science Society of America Journal
48, 758 762.
Flynn, H.C., McMahon, V., Diaz, G.C., Demergasso, C.S., Corbi
sier, P., Meharg, A.A., Paton, G.I., 2002. Assessment of bioavailable arsenic and copper in soils and sediments from the
Antofagasta region of northern Chile. The Science of the Total
Environment 286, 51 59.
Gillman, G.P., Sumpter, E.A., 1986. Modification to the compulsive
exchange method for measuring exchange characteristics of
soils. Australian Journal Soil Research 24, 61 66.
Greenwood, N.N., Earnshaw, A., 1984. Chemistry of the Elements.
Pergamon, Oxford.
Harmens, H., Gusmao, N.G.C.P.B., Den Hartog, P.R., Verkleij,
J.A.C., Ernst, W.H.O., 1993. Uptake and transport of zinc in
zinc-sensitive and zinc-tolerant Silene vulgaris. Journal of Plant
Physiology 141, 309 315.
Hermann, R., Neumann-Mahlkau, P., 1985. The mobility of zinc,
cadmium, copper, lead, iron and arsenic in ground water as a
function of redox potential and pH. The Science of the Total
Environment 43, 1 12.
Jung, M.C., Thornton, I., Chon, H.-T., 2002. Arsenic, Sb and Bi
contamination of soils, plants, waters and sediments in the vicinity of the Dalsung Cu-W mine in Kores. The Science of the
Total Environment 295, 81 89.
Kabata-Pendias, A., Pendias, H., 1984. Trace Elements in Soil and
Plants. CRC Press, Boca Raton.
Li, X., Thornton, I., 1993. Arsenic, antimony and bismuth in soil and
pasture herbage in some old metalliferous mining areas in England. Environmental Geochemistry and Health 15, 135 144.

F. Baroni et al. / Journal of Geochemical Exploration 81 (2004) 114


Ma, L.Q., Komar, K.M., Tu, C., Zhang, W., Cai, Y., Kennelley,
E.D., 2001. A fern that hyperaccumulates arsenic. Nature 409,
579.
Madejon, P., Murillo, J.M., Maranon, T., Cabrera, F., Lopez, R.,
2002. Bioaccumulation of As, Cd, Cu, Fe and Pb in wild grasses
affected by the Aznalcollar mine spill (SW Spain). The Science
of the Total Environment 290, 105 120.
MAFF Ministry of Agriculture, Fisheries and Food, 1982. Survey
of Arsenic in Food. HMSO, London.
Masscheleyn, P.H., Delaume, R.D., Patrick Jr, W.H., 1991. Effect of
redox potential and pH on arsenic speciation and solubility in a
contaminated soil. Environmental Science and Technology 25,
1414 1419.
McBride, M.B., 1994. Environmental Chemistry of Soils. Oxford
Univ. Press, New York.
Mitchell, P., Barr, D., 1995. The nature and significance of public
exposure to arsenic: a review of its relevance to South West
England. Environmental Geochemistry and Health 17, 57 82.
Murdoch, A., Clair, T.A., 1986. Transport of arsenic and mercury
from gold mining activities through an aquatic system. The
Science of the Total Environment 57, 432 437.
NRC National Research Council, 1977. Distribution of arsenic in
the environment. In: Committee on Medical and Biological Effects of Environmental Pollutants (Ed.), Arsenic. National Academy of Sciences, Washington D.C., pp. 16 79.
Nriagu, J.O., Pacyna, J.M., 1988. Quantitative assessment of worldwide contamination of air, water and soils by trace metals. Nature 338, 134 139.
ONeill, P., 1995. Arsenic. In: Alloway, B.J. (Ed.), Heavy Metals in
Soils. Chapman & Hall, Glasgow, pp. 105 121.
Otte, M.L., Ernst, W.H.O., 1994. Arsenic in vegetation of wetlands.
In: Nriagu, J.O. (Ed.), Arsenic in the Environment (Part I).
Wiley, New York, pp. 365 379.
Otte, M.L., Rozema, J., Beek, M.A., Kater, B.J., Broekman, R.A.,
1990. Uptake of arsenic by estuarine plants and interactions with
phosphate, in the field (Rhine estuary) and under outdoor experimental conditions. The Science of the Total Environment
97/98, 839 854.
Pacyna, J.M., Trevor Scholz, M., Li, Y.F., 1995. Global budget of
trace metal sources. NRC Environmental Reviews 3/2, 145 159.
Paliouris, G., Hutchinson, T., 1991. Arsenic, cobalt and nickel tolerances in two populations of Silene vulgaris (Moench) Garcke
from Ontario, Canada. New Phytologist 117, 449 459.
Pignatti, S., 1982. Flora dItalia. Edagricole, Bologna.
Pitten, F.A., Muller, G., Konig, P., Schmidt, D., Thurow, K.,
Kramer, A., 1999. Risk assessment of a former military base
contaminated with organoarsenic-based warfare agents: uptake
of arsenic by terrestrial plants. The Science of the Total Environment 226, 237 245.
Porter, E.K., Peterson, P.J., 1975. Arsenic accumulation by plants
on mine waste (United Kingdom). The Science of the Total
Environment 4, 365 371.
Porter, E.K., Peterson, P.J., 1977a. Arsenic tolerance in grasses
growing on mine waste. Environmental Pollution 14, 255 265.
Porter, E.K., Peterson, P.J., 1977b. Biogeochemistry of arsenic on
polluted sites in SW England. Trace Substances in Environmental Health 9, 89 99.

13

Protano, G., Riccobono, F., 1997. Environmental levels of antimony, arsenic and mercury in the Tafone mining area (southern Tuscany, Italy). Atti Societa` Toscana Scienze Naturali 104,
75 83.
Protano, G., Riccobono, F., Sabatini, G., 1998. La cartografia
geochimica della Toscana meridionale. Criteri di realizzazione
e rilevanza ambientale attraverso gli esempi di Hg, As, Sb,
Pb e Cd. Memorie Descrittive Carta Geologica dItalia 55,
pp. 119 140. With English Abstr.
Protano, G., Riccobono, F., Sabatini, G., 1999. Geochemical
anomalies for toxic heavy elements in southern Tuscany (Italy):
anthropogenic and natural sources. Proc. 2nd Conf. Environmental Geochemical Baseline Mapping in Europe, Vilnius, Lithuania, pp. 71 75.
Qian, J.H., Zayed, A., Zhu, Y.L., Yu, M., Terry, N., 1999. Phytoaccumulation of trace elements by wetland plants: III. Uptake and
accumulation of ten trace elements by twelve plant species.
Journal of Environmental Quality 28, 1448 1455.
Riccobono, F., 1993. I giacimenti minerari. In: Giusti, F. (Ed.), La
Storia Naturale della Toscana Meridionale. Amilcare Pizzi, Milano, pp. 107 139.
Robinson, B., Outred, H., Brooks, R., Kirkman, J., 1995. The distribution and fate of arsenic in the Waikato River system, North
Island, New Zealand. Chemical Speciation and Bioavailability
7, 89 96.
Schmoger, M.E.V., Oven, M., Grill, E., 2000. Detoxification of
arsenic by phytochelatins in plants. Plant Physiology 122,
793 801.
SISS Societa` Italiana per la Scienza del Suolo, 1985. Metodi normalizzati per lanalisi del suolo. Edagricole, Bologna.
Sneller, F.E.C., Van Heerwaarden, L.M., Kraaijeveld-Smit, F.J.L.,
Ten Bookum, W.M., Koevoets, P.L.M., Schat, H., Verkleij,
J.A.C., 1999. Toxicity of arsenate in Silene vulgaris, accumulation and degradation of arsenate-induced phytochelatins. New
Phytologist 144, 223 232.
Sostaric-Pisacic, K., Kovacevic, J., 1974. Evaluation of quality and
total value of grassland and leys by the Complex Method.
Agriculturae Conspectus Scientificus, University of Zagreb,
14 102 (in Croatian, with English Abstr).
Storer, D.A., 1984. A simple high volume ashing procedure for determining soil organic matter. Communications in Soil Science
and Plant Analysis 15, 759 772.
Temple, P.J., Linzon, S.N., Chai, B.I., 1977. Contamination of vegetation and soil by arsenic emission from secondary lead smelters. Environmental Pollution 12, 311 320.
Thornton, I., Abrahams, P.W., 1983. Soil ingestiona major pathway of heavy metals into livestock grazing contaminated land.
The Science of the Total Environment 28, 287 294.
Ure, A.M., Quevauviller, P., Muntau, H., Griepink, B., 1993. Speciation of heavy metals in soils and sediments. An account of
the improvement and harmonization of extraction techniques
undertaken under the auspices of the BCR of the Commission
of the European Communities. International Journal of Environmental Analytical Chemistry 51, 135 151.
Vink, B.W., 1996. Stability relations of antimony and arsenic compounds in the light of revised and extended Eh pH diagrams.
Chemical Geology 130, 21 30.

14

F. Baroni et al. / Journal of Geochemical Exploration 81 (2004) 114

Visoottiviseth, P., Francesconi, K., Sridokchan, W., 2002. The potential of Thai indigenous plant species for the phytoremediation
of arsenic contaminated land. Environmental Pollution 118,
453 461.
Warren, H.V., Delavault, R.E., Barasko, J., 1968. The arsenic content of Douglas Fir as a guide to some gold, silver and base
metal deposits. Canadian Mining Metallurgical Bulletin 7, 1 9.
Wedepohl, K.H., 1970. Handbook of Geochemistry. Springer,
Berlin.
Wenzel, W.W., Jockwer, F., 1999. Accumulation of heavy metals in
plants grown on mineralised soils of the Austrian Alps. Environmental Pollution 104, 145 155.

Yan-Chu, H., 1994. Arsenic distribution in soils. In: Nriagu, J.O.


(Ed.), Arsenic in the Environment (Part II). Wiley, New York,
pp. 17 49.
Ye, Z.H., Baker, A.J.M., Wong, M.H., Willis, A.J., 1997. Zinc, lead
and cadmium tolerance, uptake and accumulation by the common reed, Phragmites australis (Cav.) Trin. ex Steudel. Annals
of Botany 80, 363 370.
Zhu, Y.L., Zayed, A.M., Qian, J.-H., de Souza, M., Terry, N., 1999.
Phytoaccumulation of trace elements by wetland plants: II.
Water hyacinth. Journal of Environmental Quality 28, 339 344.

Anda mungkin juga menyukai