You are on page 1of 7

Chapter 23

Thallium

Paula Madejón

Abstract Thallium (Tl) is widely distributed in the natural environment although


at very low concentrations. It mainly occurs in the oxidation state Tl (I), whilst
Tl (III) increases under acid and oxidizing conditions. Geochemical behaviour of Tl
is analogous to that of potassium. Thallium does not occur in a free state in nature
although several minerals contain it as a major constituent. The most common
Tl-containing minerals are Lorandite and Crooksite. This element is mobilised by
the combustion of fuels and other industrial processes and tends to persist in soils,
depending on the soil type. It is considered a non-essential element and highly toxic
to living organism. It is relatively easily taken up by plants and enters the food chain
and it has been shown to accumulate in fish and other animals, with toxic effects.

Keywords Thallium • Cruciferae • Pollutant • Potassium • Rodenticide • Silicates


• Toxicity

23.1 Introduction

Thallium (Tl) is widely distributed in the natural environment, generally present at


very low concentrations. The Tl content of the earth’s crust ranges from 0.5 to
1.0 mg kg1 [43] and 0.013 mg kg1 in the oceanic crust [8]. Its content seems to
increase with increasing acidity (silicon content) of igneous rocks and with increasing
clay contents of sedimentary rocks [14, 15].
In nature, the primary oxidation state of Tl is I (thallous species), and it occurs in
soil via isomorphous substitution with K+ in feldspars and silicates. It can also occur
in the III oxidation state (thallic species) [25]. Geochemical behaviour of Tl is

P. Madejón (*)
Protection of the Soil-Water-Plant System, Instituto de Recursos Naturales y Agrobiologı́a de
Sevilla (IRNAS), CSIC, Avda. Reina Mercedes, 10, 41012 Seville, Spain
e-mail: pmadejon@irnase.csic.es

B.J. Alloway (ed.), Heavy Metals in Soils: Trace Metals and Metalloids in Soils 543
and their Bioavailability, Environmental Pollution 22, DOI 10.1007/978-94-007-4470-7_23,
# Springer Science+Business Media Dordrecht 2013
544 P. Madejón

analogous to that of K, the lower oxidation state may predominate in aqueous


systems [25]. The most dominant species in nature is Tl (I), whilst the proportion of
Tl (III) increases considerably under strongly acidic and oxidizing conditions [23].
Thallium (I) compounds resemble the compounds of alkali metals and are soluble
in water. Thallium (I) compounds are easily oxidized by Br, chloride, hydrogen
peroxide, or nitrous acid. Thallium (III) compounds are reduced to Tl (I), for example
by sulfurous acid. Thallium (III) compounds tend to be appreciably more stable than
Tl (I) compounds [17].
Concentrations of Tl in uncontaminated soils generally range from 0.01 to
3 mg kg1, but most have <1.00 mg kg1 [10]. Marine sediments of different
origins show Tl concentrations from 0.14 to 1.13 mg kg1 [43]. There has been little
research on Tl concentrations in world soils compared to other elements [27, 38].
Logan [28] reported total values of Tl (HNO3-extractable) in some British soils
in the range 0.03–0.99 mg kg1. Tremel et al. [46] reviewed several studies on Tl in
soils and reported that most arable soils in France contained 0.13–1.54 mg Tl kg1,
with a median value of 0.29 mg kg1. A slightly lower range of Tl concentrations
was found in 460 soil samples from Upper Austria (0.08–0.91 mg kg1) [11].
A range of 835 soil samples from China contained between 0.29 and 1.17 mg
Tl kg1 [49]. Non-polluted soil from the south of Spain contains approximately
0.70 mg Tl kg1 [7]. (See also Chap. 2, Table 2.1).
Although world annual production of Tl does not exceed 15,000 t, it is esti-
mated that about 2,000–5,000 t year1 are mobilised by combustion of fossil fuels,
ferrous and non ferrous metal smelting (Pb, Cu and Zn ores) and other industrial
processes such as cement production [9, 18]. Due to the high volatility of some Tl
compounds, there is considerable enrichment of Tl in the fine particles in dusts
and fly ashes that are not efficiently retained by electrostatic precipitators or other
emission control facilities. Thus, a large fraction of Tl is released into the atmo-
sphere [9].
In the past, Tl was extensively used for medical purposes and also as a rodenti-
cide. Its main current uses are in the electrical and electronics industries, in the
manufacture of semi conductors, scintillation counters and low temperature ther-
mometers, and in mixed crystals for infrared instruments and laser equipment. It is
also used with Se in the production of special glasses. Alloyed with Pb, Zn, Ag
and Sb, Tl enhances resistance to corrosion and is also used for catalysing
organic reactions. Radioactive isotopes of Tl are used in physics, in industry and
medicine [18, 36].

23.2 Lithogenic Occurrence and Geochemistry

Thallium does not occur in a free state in nature although several minerals contain it
as a major constituent [1]. Thallium is classed as a lithophile element (occurs with
or in silicates), forms sparingly soluble sulphides and is a minor constituent of
23 Thallium 545

pyrites and other sulphide minerals (e.g., ZnS sphalerite) and thus shows some
chalcophile characteristics [36].
It is found in sulphide ores of Pb, Zn and Cu and in coal [18] and the
deposits characteristically high in As are usually also high in Tl [1]. Relatively high
contents have been found in granite and shale, with intermediate values for limestone,
sandstone and coal; however, much higher values have been found in organic-rich
shales and coals of the Jurassic period with values of <1,000 mg kg1 [43]. Although
many Tl-containing minerals have been discovered and described, the following are
the most common: Lorandite (TlAsS2) and Crooksite (Cu, Tl, Ag)2Se (minerals with
thallium levels up to 60%); [18], Hutchisonite (Pb, Tl)2 (Cu, Ag)As5S10, and Urbaite
(Tl4Hg3 Sb2As8S20).
Thallium occurs naturally as two stable isotopes: 203Tl and 205Tl with relative
abundances of 29.5% and 70.5% [41]. The 205Tl/203Tl isotopic ratio of 2.378 varies
with geological processes and hence can not be used as a fingerprint for sources of
Tl in the environment.

23.3 Chemical Behaviour in Soil and Availability to Plants

Thallium is readily mobilized and transported together with alkaline metals


during weathering. However, once deposited, Tl tends to persist in soil, depending
on soil type. Greater retention is found in soils containing large amounts of clay,
organic matter, Fe and Mn oxides. Thallium is retained in the upper layers of soil,
less so in acid soils [18]. Yang et al. [53] found that natural and anthropogenic Tl is
distributed differently between soil components; anthropogenic Tl was mainly
retained in the labile fractions of soils (~80%) whereas natural Tl is predominantly
found in the residual fraction (~98%).
There is little information on the chemical form taken by Tl in soil. In contrast to
other pollutant elements (Zn, Pb, Cd, and Cu), extraction with aqua regia yielded
only 28–74% of total contents [30]. In soil samples from Bulgaria, aqua regia
and conc. HCl leached about half of the amount obtained with HNO3/HF. Only 6%
of total was found to be exchangeable with NH4Cl, pH ¼ 7, or with KCl [45].
Thallium is considered a non-essential element and highly toxic to living
organisms, [38]. Kaplan et al. [16] studied the toxicity of Tl in beans grown in
hydroponic culture and showed that Tl was accumulated in roots and not in the aerial
part. They also found that micromolar concentrations of Tl added to the nutrient
solution altered the concentration of certain nutrients in soybean tissues; drastically
reduced plant biomass with severely stunted, unbranched root systems and chlorosis
of mature leaves despite the presence of adequate levels of Fe. On the other hand
Allus et al. [2] studied rape and barley grown in culture solutions with a range of Tl
concentrations and found differences in the susceptibility of the two species to Tl.
Thallium content in barley was higher in roots whilst in rape, Tl contents were
higher in shoots.
The Tl content of plants seems to be a function of the concentration of the mobile
fraction in soils. It is relatively easily taken up by plants because it is generally
546 P. Madejón

present as thermodynamically stable Tl (I), an analogue of potassium; through


plants, it enters the food chain and it has been shown to accumulate in fish and
other animals, with toxic effects [3, 26]. Toxicity of Tl is probably due to its
interactions with K (both elements have similar ionic radii) especially by substitu-
tion into enzymatic systems, such as (Na+/K+)-ATPase and other monovalent
cation-activated enzymes, as well as to its high affinity with sulfhydryl groups of
proteins and other biomolecules [6, 24, 29]. Other systems similarly affected include
pyruvate kinase [39] and phosphatases [12].
The transfer from soil to plants generally depends on species-specific pro-
perties, as well as soil properties, such as soil texture and humus contents, cation
exchange capacity, pH and other properties. Madejón et al. [33] showed that the Tl
content of plants (Brassicae) growing in semi-arid conditions can be signific-
antly influenced by precipitation. In dry years, plant Tl accumulation may be
significantly reduced.
The reference content of Tl in plants has been calculated by Markert [34] at
0.05 mg kg1. Smith and Carson [43] gave Tl levels in various plants (in mg kg1):
vegetables 0.02–0.13 and clover 0.008–0.01. Madejón et al. [31] found Tl levels of
0.005 mg kg1 in sunflower seeds, 0.06 mg kg1 in roots (the pseudo total soil content
of Tl was 0.50 mg kg1) and 0.009 mg kg1 in seeds and 0.32 mg kg1 in roots of
sunflowers in soils with a level of 2 mg kg1 due to a mine spill accident.
Concentrations of Tl in fruits (seeds) of Holm oak (Quercus ilex) were 0.002 mg kg1
in non affected trees and 0.01 mg kg1 in seeds from trees affected by a mine spill [32].
It is well known that the uptake rate of heavy metals can vary between plant
species [35]. Some plants, especially in the Cruciferae and Graminae families,
exhibit a special ability to take up Tl. Tremel et al. [47] showed that Tl is accu-
mulated by S-rich plants from the family Brassicaceae. Several studies have
investigated Tl accumulation in different members of this family. Crop plants
such as green cabbage [50, 51], white cabbage (Brassica oleracea capitata) and
kale (Brassica oleracea acephala) [21] have been shown to accumulate Tl in edible
parts (shoots and foliage). Soriano and Fereres [44] reported relatively high Tl
accumulations in the aerial biomass of Brassica napus and B. carinata. Other
brassicas such as Iberis intermedia and Biscutella laegiviata have been found to
hyperaccumulate Tl in their shoot tissues [5, 22, 42].

23.4 Soil Contamination

There are several cases where soils with naturally elevated Tl contents of
>2 mg kg1 have been observed [4, 46–48]. These soils are generally developed
on K-rich magmatites (e.g., granites, syenites) where Tl+ isomorphically substitutes
for K+ or Rb+, especially in K-feldpars and micas [13, 19]. High Tl concentrations
of <73 mg kg1have been found in soils around old mines [40] due to contamina-
tion from Tl-rich sulphide ores of Hg, As and Au (40–124 mg kg1) [52] from
pyrite processing areas (5–15 mg kg1) [53] and from a zone of Pb-Zn exploitation
(9–28 mg kg1) [27].
23 Thallium 547

23.5 Risk Assessment

Thallium is a highly toxic element and is a US Environmental Protection Agency


(EPA) priority pollutant [37]. The ecotoxicological importance of Tl is derived from
its high acute toxicity to living organisms, comparable to that of Pb and Hg [19].
The major pathway of Tl exposure for animals and humans is the ingestion of
plants grown in Tl-contaminated soils [42]. In Germany, 1 mg kg1 Tl in soils has
been established as the tolerance level for agricultural use [20].
Xiao et al. [51] reported a case study drawing attention to the fact that
natural processes can mobilize Tl, which may enter the food chain as a “hidden
health killer” with severe health impacts on the local human population. The study
was carried out in Guizhou Province, China, in which “natural contamination” of
Tl in bedrocks/ores was within the range 6–35,000 mg kg1. This led to enrich-
ment of Tl in the aquatic system (0.005–1,100 mg L1 in groundwaters and
0.07–31 mg L1 in surface waters) and soil layers (1.5–124 mg kg1). Concen-
trations of 1–500 mg kg1Tl were determined in many food crops growing on
Tl-contaminated arable soils. The daily intake for inhabitants consuming these
crops was 1.9 mg of Tl, showing a latent health hazard with potential risk of
toxicity in humans within areas of “natural” contamination of Tl. In fact in a
retrospective study, Xiao et al. [52] found that majority of the volunteer subjects
from these communities had urinary Tl concentrations above 4.5–6 mg L1, imply-
ing early adverse health effects, and some had over 500 mg L1 urinary Tl,
considered to be on the threshold of clinical intoxication.

References

1. Adriano, D. C. (2001). Trace elements in terrestrial environments. Biochemistry, bioavailabil-


ity and risks of metals. New York: Springer.
2. Allus, M. A., Martin, M. H., & Nickless, G. (1987). Comparative toxicity of thallium to two
plant species. Chemosphere, 16, 929–932.
3. Al-Najar, H., Schulz, R., & Romheld, V. (2003). Plant availability of thallium in the rhizo-
sphere of hyperaccumulator plants: A key factor for assessment of phytoextraction. Plant and
Soil, 249, 97–105.
4. Al-Najar, H., Kaschl, A., Schulz, R., & R€omheld, V. (2005). Effect of thallium fractions in the
soil and pollution origins on Tl uptake by hyperaccumulator plants. International Journal of
Phytoremediation, 7, 55–67.
5. Anderson, C. W. N., Brooksa, R. R., Chiaruccib, A., LaCostea, C. J., Leblancc, M., Robinson,
B. H., Simcocke, R., & Stewarta, R. B. (1999). Phytomining for nickel, thallium and gold.
Journal of Geochemical Exploration, 67, 407–415.
6. Britten, J. S., & Blank, M. (1968). Thallium activation of the (Na+-K+)-activated ATPase of
rabbit kidney. Biochimica et Biophysica Acta, 159, 160–166.
7. Cabrera, F., Clemente, L., Dı́{az Barrientos, E., López, R., & Murillo, J. M. (1999). Heavy
metal pollution of soils affected by the Guadiamar toxic flood. Science of the Total Environ-
ment, 242, 117–129.
548 P. Madejón

8. Delvalls, T. A., Saenz, V., Arias, A. M., & Blasco, J. (1999). Thallium in the marine
environment: First ecotoxicological assessments in the Guadalquivir estuary and its potential
adverse effect on the Doñana European natural reserve after the Aznalcóllar mining spill.
Ciencias Marinas, 25, 161–175.
9. Ewers, U. (1988). Environmental exposure to thallium. Science of the Total Environment,
71, 285–292.
10. Fergusson, J. E. (1990). The heavy elements: Chemistry, environmental impact and health
effects. Oxford: Pergamon Press.
11. Hoffer, G. F., Aichberger, K., & Hochmair, U. S. (1990). Thalliumgehalte landwirtschaftlich
genutzter B€oden Ober€ osterreichs. Die Bodenkultur, 41, 187–193.
12. Inturrisi, C. E. (1969). Thallium-induced dephosphorylation of a phosphorylated intermediate
of the (sodium + thallium-activated) ATPase. Biochimica et Biophysica Acta – Enzymology,
178, 630–633.
13. Jović, V. (1999). Thallium. In C. P. Marshall & R. W. Fairbridge (Eds.), Encyclopedia of
geochemistry (pp. 622–623). Dordrecht: Kluwer Academic Publishers.
14. Kabata-Pendias, A., & Pendias, H. (2001). Trace elements in soils and plants (3rd ed.).
Boca Raton: CRC Press.
15. Kabata-Pendias, A., & Mukherjee, A. B. (2007). Trace elements from soil to human.
Berlin: Springer.
16. Kaplan, D. I., Adriano, D. L., & Sajwan, K. S. (1990). Thallium toxicity in bean. Journal of
Environmental Quality, 19, 359–365.
17. Kaplan, D. I., & Mattigod, S. V. (1998). Aqueous geochemistry of thallium. In J. O. Nriagu (Ed.),
Thallium in the environment (pp. 15–29). New York: Wiley.
18. Kazantzis, G. (2000). Thallium in the environment and health effects. Environmental Geo-
chemistry and Health, 22, 275–280.
19. Kemper, F. H., & Bertram, H. P. (1991). Thallium. In E. Merian (Ed.), Metals and their
compounds in the environment. Occurrence, analysis and biological relevance
(pp. 1227–1241). Weinheim: VCH.
20. Kloke, A. (1980). Richtwerte’80: Orientierungswerte f€ ur tolerierbare Gesamtgehalte einiger
Elemente in Kulturb€ oden. VDLUFA Sonderdruck (especial print).
21. Kurz, H. (1999). Selection of cultivars to reduce the concentration of cadmium and thallium in
food and fodder plants. Journal of Plant Nutrition and Soil Science, 162, 323–328.
22. Leblanc, M., Petit, D., Deram, A., Robinson, B. H., & Brooks, R. R. (1999). The phytomining
and environmental significance of hyperaccumulation of thallium by Iberis intermedia from
southern France. Economic Geology, 94, 109–113.
23. Lee, A. G. (1971). The chemistry of thallium. Amsterdam: Elsevier.
24. Lehn, H., & Schoer, J. (1987). Thallium-transfer from soils to plants: Correlation between
chemical form and plant uptake. Plant and Soil, 97, 253–265.
25. Lin, T. S., & Nriagu, J. O. (1999). Thallium speciation in great lakes. Environmental Science
and Technology, 33, 3394–3397.
26. Lin, T. S., Meier, P., & Nriagu, J. (2005). Acute toxicity of thallium to Daphnia magna and
Ceriodaphnia dubia. Bulletin of Environmental Contamination and Toxicology, 75, 350–355.
27. Lis, J., Pasieczna, A., Karbowska, B., Zembrzuski, W., & Lukaszewski, Z. (2003). Thallium in
soils and stream sediments of a Zn-Pb mining and smelting area. Environmental Science and
Technology, 37, 4569–4572.
28. Logan, P. G. (1985). Thallium uptake and transport in plants. PhD thesis, CNAA.
29. Logan, P. G., Lepp, N. W., & Phipps, D. A. (1984). Some aspects of thallium uptake by higher
plants. Trace Substances Environmental Health, 18, 570–575.
30. Lukaszewski, Z., & Zembrzuski, W. (1992). Determination of thallium in soils by flow-
injection-differential pulse anodic stripping voltammetry. Talanta, 39, 221–227.
31. Madejón, P., Murillo, J. M., Marañón, T., Cabrera, F., & Soriano, M. A. (2003). Trace element
and nutrient accumulation in sunflower plants two years after the Aznalcóllar mine spill.
Science of the Total Environment, 307, 239–257.
23 Thallium 549

32. Madejón, P., Marañón, T., & Murillo, J. M. (2006). Biomonitoring of trace elements in the
leaves and fruits of wild olive and holm oak trees. Science of the Total Environment,
355, 187–203.
33. Madejón, P., Murillo, J. M., Marañón, T., & Lepp, N. W. (2007). Factors affecting accumula-
tion of Thallium and other trace elements in two wild Brassicaceae spontaneously growing on
soils contaminated by tailing dam waste. Chemosphere, 67, 20–28.
34. Markert, B. (1992). Multi-element analysis in plant materials-analytical tools and biological
questions. In D. C. Adriano (Ed.), Biogeochemistry of trace elements (pp. 401–428).
Boca Raton: Lewis Publishers.
35. Marschner, H. (1995). Mineral nutrition of higher plants (2nd ed.). London: Academic.
36. Nriagu, J. O. (1998). History, production, and uses of thallium. In J. O. Nriagu (Ed.), Thallium
in the environment (pp. 1–14). New York: Wiley.
37. Peter, A. L. J., & Viraraghavan, T. (2005). Thallium: A review of public health and environ-
mental concerns. Environmental International, 31, 493–501.
38. Reimann, C., & de Caritat, P. (1998). Chemical elements in the environment. Factsheets for the
geochemist and environmental scientist. Berlin: Springer.
39. Reuben, J., & Kane, F. J. (1971). Thallium-205 nuclear magnetic resonance study of pyruvate
kinase and its substrates. Evidence for a substrate-induced conformational change. Journal of
Biological Chemistry, 20, 6227–6234.
40. Sager, M. (1998). Thallium in agricultural practice. In J. O. Nriagu (Ed.), Thallium in the
environment (pp. 59–87). New York: Wiley.
41. Sahl, K., Albuquerque, C. A. R., & Shaw, D. M. (1978). Thallium. In K. H. Wedepohl (Ed.),
Handbook of geochemistry. Berlin: Springer.
42. Scheckel, K. G., Lombi, E., Rock, S. A., & Mclaughlin, M. J. (2004). In vivo Synchrotron
study of thallium speciation and compartmentation in Iberis intermedia. Environmental
Science and Technology, 38, 5095–5100.
43. Smith, I. C., & Carson, B. L. (1977). Trace metals in the environment. I. Thallium. Michigan:
Ann Arbor Science Publishers.
44. Soriano, M. A., & Fereres, E. (2003). Use of crops for in situ phytoremediation of polluted
soils following a toxic flood from a mine spill. Plant and Soil, 256, 253–264.
45. Tsakovski, S., Ivanova, E., & Havezov, I. (1994). Flame AAS determination of thallium in
soils. Talanta, 41, 721–724.
46. Tremel, A., Masson, P., Sterckeman, T., Baize, D., & Mench, M. (1997). Thallium in French
ecosystems-I. Thallium content in arable soils. Environmental Pollution, 95, 293–302.
47. Tremel, A., Massona, P., Garraudb, H., Donardb, O. F. X., Baizec, D., & Mench, M. (1997).
Thallium in French agrosystems – II. Concentration of thallium in field-grown rape and some
other plant species. Environmental Pollution, 97, 161–168.
48. Vanek, A., Chrastný, V., Mihaljevič, M., Drahota, P., Grygar, T., & Komárek, M. (2009).
Lithogenic thallium behaviour in soils with different land use. Journal of Geochemical
Exploration, 102, 7–12.
49. Wenqi, Q., Yalei, C., & Jieshan, C. (1992). Indium and thallium background contents in soils
in China. International Journal of Environmental Studies, 40, 311–315.
50. Xiao, T., Guha, J., Boyle, D., Liu, C.-Q., & Chen, J. (2004). Environmental concerns related to
high thallium levels in soils and thallium uptake by plants in southwest Guizhou, China.
Science of the Total Environment, 318, 223–244.
51. Xiao, T., Guha, J., Boyle, D., Liu, C.-Q., Zheng, B., Wilson, G. C., Rouleau, A., & Chen, J. (2004).
Naturally occurring thallium: A hidden geoenvironmental health hazard? Environmental
International, 30, 501–507.
52. Xiao, T., Guha, J., Liu, C.-Q., Zheng, B., Wilson, G., Ning, Z., & He, L. (2007). Potential
health risk in areas of high natural concentrations of thallium and importance of urine
screening. Applied Geochemistry, 22, 919–929.
53. Yang, C., Chen, Y., Peng, P., Li, C., Chang, X., & Xie, C. (2005). Distribution of natural and
anthropogenic thallium in the soils in an industrial pyrite slag disposing area. Science of the
Total Environment, 341, 159–172.

You might also like