Sie sind auf Seite 1von 163

SOIL -SOLID SPECIES CONTROLLING HEAVY METAL

AVAILABILITY IN CONTAMINATED SOILS OF ZAMFARA


STATE, NORTHERN NIGERIA.

By
IBRAHIM MOHAMMED

DEPARTMENT OF SOIL SCIENCE,


FACULTY OF AGRICUTURE,
AHMADU BELLO UNIVERSITY, ZARIA

JULY, 2014.

SOIL -SOLID SPECIES CONTROLLING HEAVY METAL


AVAILABILITY IN CONTAMINATED SOILS OF ZAMFARA STATE,
NORTHERN NIGERIA.

By
IBRAHIM MOHAMMEDB. Agric. (ABU, 2009)
M Sc./Agric./3573/2011-12

A THESIS SUBMITTED TO THE SCHOOL OF POST GRADUATE STUDIES,


AHMADU BELLO UNIVERSITY, ZARIA

IN PARTIAL FULFILMENT OF THE REQUIREMENTS FOR THE AWARD


OF A
MASTER DEGREE IN SOIL SCIENCE

DEPARTMENT OF SOIL SCIENCE,


FACULTY OF AGRICUTURE,
AHMADU BELLO UNIVERSITY, ZARIA
NIGERIA

JULY, 2014.

DECLARATION
I declare that the work in this thesis entitled Soil -solid species controlling heavy metal
availability in contaminated soils of Zamfara State, northern Nigeria has been carried out
by me in the Department of Soil Science. The information derived from the literature has been
duly acknowledged in the text and a list of references provided. No part of this thesis was
previously presented for another degree or diploma at this or any other Institution.

MOHAMMED IBRAHIM
..............................................................

....................................

Name of Student

Signature

ii

......................................

Date

CERTIFICATION
This thesis entitled SOIL -SOLID SPECIES CONTROLLING HEAVY METAL
AVAILABILITY IN CONTAMINATED SOILS OF ZAMFARA STATE, NORTHERN
NIGERIA by IBRAHIM Mohammed meets the regulations governing the award of the degree
of Master of Science in Soil Science of Ahmadu Bello University, and is approved for its
contribution to knowledge and literary presentation.

.
Dr. Nafiu Abdu
Chairman, Supervisory Committee

.
Date

.
Dr. S. T. Abu.
Member, Supervisory committee

...
Date

Dr. E. Y. Oyinlola
Member, Supervisory committee

...
Date

.
Dr. Ado A. Yusuf
Ag. Head, Department of Soil Science

...
Date

Prof. A. A Joshua
Dean, Postgraduate School

...
Date

iii

ACKNOWLEDGEMENT
I am eternally indebted to my supervisor, Dr Nafiu Abdu, without whose assistance and genuine
interest this work wouldnt have been completed; I hope to find some way of expressing my
gratitude. My sincere gratitude also goes to other members of my Supervisory team, Drs S. T.
Abu and E. Y. Oyinlola. I appreciate the effort of the Head, Department of Soil Science and all
the members of staff in their quest to make the Department a pace-setter. I am grateful to the
group of lecturers whose intriguing lectures gave me a firm foundation for writing up this thesis.
The assistance of Messrs Ilu, Suleiman, Jatau, and Anyanwu, as well as Mrs. Cecilia of the
Department of Soil Science laboratories; and Mr Bashir of the multipurpose laboratory,
Chemistry Department, are highly acknowledged and appreciated. I am grateful to my
colleagues, friends, and other wonderful people, too numerous to mention, who have made my
exploits worthwhile.
I would like to thank Professor and Mrs. Lawal and family for their affection and ever-present
support; and of course mum, sisters and brother, I am forever grateful.
This study was supported financially by the Academy of Sciences for the Developing World
(TWAS) through grant No. 11-062 RG/CHE/AF/AC_I-UNESCO FR: 3240262664 to Dr. Nafiu
Abdu. The supplementary support received from the Institute for Agricultural Research, Ahmadu
Bello during field sampling is also acknowledged.

iv

DEDICATION

To:
my sister, HALIMA
for that simple question: why not soil chemistry?

ABSTRACT
Artisanal mining in Dareta village of Zamfara, northern Nigeria has resulted in the pollution of
vast area of land and water. However, in assessing environmental pollution as well as predicting
risk of contaminant transfer, the total concentration of the pollutant is not of utmost importance,
rather their geochemical distribution among the various soil fractions is paramount. This study
was undertaken to evaluate the geochemical fractions and the vertical and horizontal distribution
of lead (Pb), cadmium (Cd) and zinc (Zn) in the contaminated farmlands of Dareta. The total
heavy metal concentration in the soil was determined after mixed acid (HNO3 + HCl + HF)
digestion using AAS. The total metal concentration was separated into five operationally defined
fractions. The extent of metal retention by the soil and the effect of pH adjustment on metal
retention were evaluated through a model sorption experiment. The distribution of metal species
and complexes in the soil solution was calculated from equilibrium metal concentration and
other input parameters using the speciation model PHREEQC. The total concentration of Pb and
Cd recorded were extreme, however, the concentration of Zn was within acceptable limits. Metal
fractionation results allotted highest proportions of Pb, Cd, and Zn to the residual fraction. Down
the profile, concentration of Pb, Cd, and Zn showed two peaks. The Pollution intensity indices
indicated high soil contamination with Pb and Cd. Although the partitioning of the bulk of the
metal onto the residual fraction reduces the risk of bioavailability, the on-going mining activities
could result in the accumulation of mobile metal fractions in the soil which will eventually affect
soil fertility and food quality.

vi

TABLE OF CONTENTS
Title page.i
Declaration..ii
Certificationiii
Acknowledgements.....iv
Dedication ...v
Abstract..vi
Table of contentsvii
List of Tables......xi
List of Figures..xiii
CHAPTER ONE
1.0 INTRODUCTION...1
1.1 Statement of Problem.3
1.2 Justification of Study.3
1.3 Specific Objectives4
CHAPTER TWO
2.0 LITERATURE REVIEW.5
2.1 Metals and heavy metals....5
2.2 Metals in the environments....7
2.3 The Soil solid phase.12
2.4 Soil solution phase...14
2.5 Multiphase equilibria...15
2.6 Behavior of Heavy Metals in Soil ..15
vii

2.7 Influence of Soil Properties on behavior of Heavy Metals..16


2.7.1 Soil pH16
2.7.2 Organic Matter ......16
2.8 Effect of heavy metals in the environment17
2.8.1 Accumulation17
2.8.2 Solubility and Mobility.18
2.8.3 Bioavailability..19
2.8.4 Toxicity.21
2.9 Analysis of heavy metal concentrations..21
2.9.1. Sequential Extraction/Fractionation schemes..22
2.10 Speciation calculation: Computer Models.24
2.11 Complications from heavy metal pollution26
CHAPTER THREE
3.0 MATERIALS AND METHOD..27
3.1 Description of Study area.27
3.2 Soil sampling ..28
3.3 Analytical Procedures..30
3.3.1 Determination of Physicochemical properties..30
3.3.2. Determination of Total Phosphorus 30
3.3.3. Determination of Total Pb, Cd, and Zn31
3.3.4. Determination of Labile metal concentrations: CaCl2 extraction31
3.3.5. Sequential Extraction..31
3.3.6 Sorption and solubility study of added Pb, Cd, and Zn33

viii

3.3.7 Solubility of Pb, Cd, and Zn.35


3.3.8. Effect of changing redox condition on the solubility of Pb, Cd, and Zn36
3.4. Data Analysis..36
CHAPTER FOUR
4.0 RESULTS AND DISCUSSION.38
4.1 Soil physico-chemical parameters...38
4.2 Total concentration of Pb, Cd, and Zn44
4.2.1 Enrichment and accumulation of Pb, Cd and Zn..47
4.3 Horizontal Distribution of Pb, Cd, and Zn in Dareta...50
4.4 Vertical distribution of Pb, Cd, and Zn53
4.5 Extractable concentration of Pb, Cd, and Zn...59
4.6 Fractionation of Pb, Cd, and Zn.. 61
4.6.1. Lead fractions..61
4.6.2. Cadmium fractions...64
4.6.3 Zinc fractions66
4.6.4 Correlation between metal fractions.68
4.7 Mobility of Pb, Cd, and Zn fractions...70
4.8 Sorption of Pb, Cd, and Zn added to the soil...70
4.9 Metal solubility as a function of pH74
4.10 Heavy metals partitioning..76
4.11 Solid phase-solution equilibria of Pb, Cd, and Zn added to soil79
4.11.1. Lead solid-solution equilibria79
4.11.2. Zinc solid-solution equilibria.83

ix

4.11.3. Cadmium solid-solution equilibria86


4.12 Soil solution speciation..90
4.12.1 Inorganic complexes of Pb..90
4.12.2 Inorganic complexes of Cd.95
4.12.3 Inorganic complexes of Zn.99
4.13 Mechanism(s) controlling Pb, Cd, and Zn activities in the metal spiked soils of Dareta103
4.14. Changes in solubility of Pb, Cd, and Zn with incubation...105
4.15. Reduction of Pb, Cd and Zn in the contaminated soils of Dareta under native conditions
and different levels of glucose amendment.....110
CHAPTER FIVE
5.0 SUMMARY AND CONCLUSION...115
REFERENCES ...119
APPENDICES.135

LIST OF TABLES
Table 2.1 Content of Various Elements in Soils6
Table 2.2 Natural and anthropogenic sources and common forms in wastes of trace metals on the
priority pollutant list9
Table 2.3 Occurrences of trace elements in primary minerals...13
Table 4.1aPhysicochemical properties of the soil at Dareta, Zamfara state, northern Nigeria.39
Table 4.1b Profile Characteristics at Dareta North40
Table 4.1c Profile Characteristics at Dareta South .. 41
Table 4.1 d Profile Characteristics at Dareta East ... 42
Table 4.1e Profile Characteristics at Dareta West ....... 43
Table 4.2 Mean concentration and standard error of Pb, Cd, and Zn in Dareta...45
Table 4.3 International threshold values for heavy metals concentration in soils (mgkg -1)..46
Table 4.4 Measurements of metal pollution in soils and sediments..49
Table 4.5 Average and background concentration, enrichment factor, calculated I-geo index, and
grade of pollution intensity of Pb, Cd and Zn in analyzed samples from Dareta..49
Table 4.6 Mean and standard error of 0.01M CaCl2 extractable concentration of Pb, Cd, and
Zn...60
Table 4.7 Correlation coefficient matrix between the different Pb and Cd fractions69
Table 4.8 Correlation coefficient matrix between the different Zn fractions69
Table 4.9 Equilibrium reactions for Pb minerals and complexes likely to control Pb2+ activities in
a Pb polluted soil81
Table 4.10 Equilibrium reactions for Zn-minerals and complexes likely to control Zn2+ activities
in soils84

xi

Table 4.11 Equilibrium reactions for Cd minerals and complexes likely to control Cd2+ activities
in soils88
Table 4.12 Equilibrium reactions for the inorganic complexes of Pbin the soil solution.91
Table 4.13 The activities of dominant inorganic complexes of Pb in soil solution of Dareta,
Northern Nigeria expressed as percentages of total Pb (PbT) in soil solution...93
Table 4.14 Ion association constant for the formation of aqueous inorganic and organic
complexes of Pb, Cd, and Zn in solution..94
Table 4.15 Equilibrium reactions for the inorganic complexes of Cd in the soil solution96
Table 4.16 The activities of dominant inorganic complexes of Cd in soil solution of Dareta,
Northen Nigeria expressed as percentages of total Cd (CdT) in soil solution98
Table 4.17 Equilibrium reactions for the inorganic complexes of Zn in the soil solution..100
Table 4.18 The activities of dominant inorganic complexes of Zn in soil solution of Dareta,
Northern Nigeria expressed as percentages of total Zn (ZnT) in solution..102
Table 4.19 The slopes and coefficient of determination of -log Me2+ - pH relationship of soils
from the lead polluted farms of Dareta, Northern Nigeria104
Table 4.20 pH and Redox potential of soil samples after 7 days incubation under two glucose
rates112
Table 4.21 Mean concentrations (mg kg-1) of solubilized Pb, Cd, and Zn under different levels of
glucose amendment113

xii

LIST OF FIGURES
Figure 3.1 Map of Nigeria, showing the location of Dareta Village, where soils for the study
were obtained.27
Figure 3.2 Field Sampling Layout.29
Figure 4.1 Mean concentrations of Pb, Cd, and Zn across the sampling distance in Dareta
North51
Figure 4.2 Mean concentrations of Pb, Cd, and Zn across the sampling distance in Dareta
South51
Figure 4.3 Mean concentrations of Pb, Cd, and Zn across the sampling distance in Dareta
East 52
Figure 4.4 Mean concentrations of Pb, Cd, and Zn across the sampling distance in Dareta
West...52
Figure 4.5 Vertical distribution of Pb down the profile, in Dareta53
Figure 4.6 Vertical distribution of Cd down the profile, in Dareta56
Figure 4.7 Vertical distribution of Zn down the profile in Dareta.58
Figure 4.8 Distribution of Pb among the different geochemical fractions at the four study
directions at Dareta62
Figure 4.9 Distribution of Cd among the different geochemical fractions at the four study
directions at Dareta65
Figure 4.10 Distribution of Zn among the different geochemical fractions at the four study
directions at Dareta....67
Figure 4.11 Mobility indices for Pb, Cd, and Zn across the study directions in Dareta71
Figure 4.12 Sorption of Pb, Cd, and Zn as a function of pH on the soils of Dareta..72

xiii

Figure 4.13 Solubility of Pb, Cd, and Zn added to the soil of Dareta as a function of pH.75
Figure 4.14 Partition coefficients (Kd) for Pb, Cd, and Zn under varying pH 78
Figure 4.15 Logarithm (base 10) of soluble Pb from the amended soil of Dareta and from various
pure compounds.....82
Figure 4.16 Logarithm (base 10) of soluble Zn from the amended soil of Dareta and from various
pure compounds.....85
Figure 4.17 Logarithm (base 10) of soluble Cd from the amended soil of Dareta and from various
pure compounds.....89
Figure 4.18 The distribution of the inorganic complexes of Pb in the soil solution of Dareta,
Northern Nigeria92
Figure 4.19 The distribution of inorganic complexes of Cd in the soil solution of Dareta,
Northern Nigeria97
Figure 4.20 The distribution of inorganic complexes of Zn in the soil solution at Dareta,
Northern Nigeria .101
Figure 4.21 Changes in solubility of Pb in Dareta soil after incubation with distilled water..106
Figure 4.22 Changes in solubility of Cd in Dareta soil after incubation with distilled water.107
Figure 4.23 Changes in solubility of Zn in Dareta soil after incubation with distilled water..108
Figure 4.24 Eh-pH relationships after 7 days of incubation at 1% and 3% glucose levels.114

xiv

CHAPTER ONE
1.0 INTRODUCTION
Anthropogenic activities have heightened the incidences of contamination in the environment.
Contaminants of concern in soil and water monitoring range from organic to inorganic and they
include plant nutrients such as nitrate and phosphates; heavy metals such as cadmium,
chromium, and lead; oxyanions such as arsenite, arsenate, and selenite; organic chemicals;
inorganic acids; and radionucleotides (Sparks, 2003). The sources of these contaminants include
fertilizers, pesticides, acidic deposition, agricultural and industrial waste materials, and
mining/smelting activities (Ross, 1994; Sparks, 2003). While the organic contaminants are
biodegradable, inorganic contaminants persist in the soil at dangerously toxic levels (McLean
and Bledsoe, 1992), with attendant severe consequences to plants, animals, and humans
(Chukwuma et al., 2012). The major inorganic contaminants are in the form of metals. Metals
undergo an array of biogeochemical processes at reactive natural surfaces, including surfaces of
clay minerals, metal oxides and oxyhydroxides, humic substances, plant roots, and microbes.
These processes control the solubility, mobility, bioavailability, and consequent toxicity of
metals in the environment. Metals in the environment may enter the soil solution (the aqueous
liquid phase of the soil and its solutes) and be subject to a number of overlapping pathways
(Sparks, 2005), that eventually could result in phyto-toxicity.
The advent of the industrial Age has been associated with large scale emissions and
depositions of metals in the environment. Scenarios where metals have accumulated in terrestrial
and aquatic environments in high concentrations and cause harm to animals and humans have
been reported (Adriano et al., 2005). With the increasing world population the need for metals
will also increase and so will the potential for soil and water contamination, through metal

emission into the environment. The soil is usually the final destination of these metals; any
change (s) in this dynamic and complex system would surely change the fate of metals in the
environment, and this could have serious implications for human health and environmental
quality (Sparks, 2005).
Changes in the soil system could be a result of metal reactions with ligands in solution
and with surface functional groups on the soil solid phase with which the soil solution is in
contact (Krishnamurti and Naidu, 2003). Usually a dynamic equilibrium exists between the
metals in solution and the soil solid phase and this controls the availability of contaminants for
plant uptake (Bohn et al., 2003; Evangelou, 1998). Also, the sorption of metals on inorganic
minerals and organic humic substances is an important geochemical process controlling the fate,
transport, and bioavailability of metals in soil and water environments (Sparks, 2005). Although
an intrinsic relationship exist between the solid and solution phases of the soil nutrient balance,
the mobility and bioavailability of metals to plants and their leachability to contaminate ground
water does not depend only on the total concentration of the metal in solution but also on the
species or forms of the metals (Abdu et al., 2012). Hence, evaluating bioavailability should
include predicting the distribution and behavior of chemical species in solution because this
would provide information regarding risk of contaminant transfer and accumulation in the food
chain (Kabata-Pendias and Pendias, 2001; Chopin and Alloway, 2007a,b). Additionally, the total
heavy metal concentration in the soil solid fraction is not a good indicator of bioavailability
(Wang et al., 2004; Ortiz and Alcaniz, 2006) because metals can be associated with various soil
components (Organic matter, clays, Fe and Mn Oxides, lattice of silicate or carbonate minerals)
that differ in their ability to retain or release them (Ure et al., 1993, Alloway, 1995; Tack et al.,
1996).

1.1 Statement of problem


Artisanal mining practices in villages in the local government areas of Anka, Gummi and
Bukkuyum in Zamfara state has resulted into a case of severe Pb pollution of vast expanse of
lands and water. According to Katz (2013), gold ore is collected from mines near Sunke. The ore
is bagged and brought by motorcycle to villages like Dareta, where villagers grind the ore and
then wash the mix over a ridged board. The result is then sold to gold traders in Gusau, the
capital of Zamfara, where remaining dirt and impurities are separated. When the small scale
miners in these villages grind ore from local gold mines filled with lead (Pb), they spew lead dust
across the ground where children eventually play and animal graze. The US Blacksmith Institute
reported that residential soil in the United States contains approximately 400 parts Pb per million
while the soil in some of these villages has about 10,000 parts Pb per million (Blacksmith
Institute, 2010). The high levels of Pb in these residential soils have heightened the risk of a
contaminated food chain, with resultant risks of bioaccumulation and biomagnification.
Although Pb is the major metal of interest, other metals such as Zn and Cd have also been
indicated as contaminants from mining activities (Sparks, 2005) and have been reported in some
Pb contaminated farmlands of Zamfara State (Abdu and Yusuf, 2012).

1.2 Justification of study


Adequate inference on the mobility and bioavailability of metals to plants and their leachability
to contaminate ground water is dependent on the various associations of metals with various soil
components (Organic matter, clays, Fe and Mn Oxides, lattice of silicate or carbonate minerals),
as well as the species or forms that the metals assume in solution (Evangelou, 1998), and less on
the metal total concentration (Krishnamurti and Naidu, 2003; Sparks, 2005).

Therefore,

identification of the geochemical fractions of metals and other sparingly soluble compounds
3

controlling heavy metal solubility in soil, as well as the distribution of metal species in solution
are of great importance in predicting heavy metal bioavailability, maintaining and improving
food quality, chemical remediation of polluted environment and solving human health problems
at large (Krishnamurti and Naidu, 2003).

1.3 Aim and Objectives:


The major aim of this research work is to determine the soil solid fractions controlling the
availability of Pb, Cd, and Zn to crops growing in the farmlands of Dareta Village. The specific
objectives of this research work are:

To determine the geochemical species of Pb, Cd, and Zn in the farmlands of Dareta

To determine the vertical and horizontal distribution of Pb, Cd, and Zn in the farmlands
of Dareta

To predict the distribution of Pb, Cd, and Zn species in the soil solution of the farmlands
of Dareta.

To evaluate the effect of redox condition on the mobility of Pb, Cd and Zn from the soil
environment at Dareta.

CHAPTER TWO
2.0 LITERATURE REVIEW
2.1 Metals and heavy metals
Metals are defined as any element that has a silvery luster and is a good conductor of heat and
electricity (McLean and Bledsoe, 1992). Metals comprise about 75% of the known elements
and can form alloys with each other and with nonmetals (Morris, 1992) (Sparks, 2005).
Traditionally, metals have been classified into categories such as light, heavy, semimetal (i.e.
metalloids), toxic, and trace, depending on several chemical and physical criteria such as density,
weight, atomic number, and degree of toxicity (Roberts et al., 2005). Certain metals and
metalloids are essential for plant growth and for animal and human health. With respect to
plants, these are referred to as micronutrients and include B, Cu, Fe, Zn, Mn, and Mo. In
addition, As, Co, Cr, Ni, Se, Sn, and V are essential in animal nutrition. Micronutrients are also
referred to as trace elements since they are required in only small quantities, unlike major
nutrients such as N, P, and K. In excess, trace elements can be toxic to plants, microbes, animals,
and humans. Problems also arise when there is a deficiency in essential elements (Sparks, 2005).
Heavy metals are elements with densities greater than 5.0 g cm-3 and usually indicate
metals and metalloids associated with pollution and toxicity; however, the term also includes
elements that are required by organisms at low concentrations (Adriano, 2001). Heavy metals are
used interchangeably with the potentially toxic elements and/or trace elements in the soil. The
relative proportion of heavy metals in the lithosphere as reported by Lindsay (1979) is presented
in Table 2.1.

Table 2.1 Content of Various Elements in Soils


Metal

Selected average for soils


mg kg-1

Common range for soils


mg/kg

Al

7100

10,000-300,000

Fe

38, 000

7,000-550,000

Mn

600

20-3,000

Cu

30

2-100

Cr

100

1-1000

0.06

0.01-0.70

Zn

50

10-300

As

1.0-5.0

Se

0.3

0.1-2

Ni

40

5-500

Ag

0.05

0.01-5

Pb

10

2-200

Hg

0.03

0.01-3

Source: Lindsay (1979)

2.2 Metals in the environments


Heavy metals are naturally present in the soil because of the geology of the parent material from
which they are formed; therefore the presence of metals in the soil is not an indication of
contamination (McLean and Bledsoe, 1992). The metal could be precipitated in complexes, or
held in other forms which limit their availability (Sposito, 2008). The concentration of heavy
metals in soil is also closely associated with biological and geochemical cycles, and is influenced
by anthropogenic activities (Oliver, 1997). In the environment, metals undergo an array of
biogeochemical processes at reactive natural surfaces, including surfaces of clay minerals, metal
oxides and oxyhydroxides, humic substances, plant roots, and microbes. These processes control
the solubility, mobility, bioavailability, and toxicity of metals in the environment. These Metal
ions may enter the soil solution (the aqueous liquid phase of the soil and its solutes) and be
subject to numerous pathways, all of which can potentially overlap (Sparks, 2005). Although
some heavy metals occur naturally in the soil, anthropogenic input remains the main sources of
metal contaminant in the environment (Bartlett and James, 1993) and the most dangerous (Abdu,
2010). Also, the non-biodegradable nature of metals in soil environment has created interest in
the possible changes in their solubility (Oliver, 1997). Anthropogenic activities have accelerated
the rate of release of metals into the environment, thereby resulting in the contamination of land
and water bodies, with their attendant severe human consequences. According to Ross (1994),
the anthropogenic sources of metal contamination can be divided into five main groups:
1. Metalliferous mining and smelting (As, Cd, and Hg )
2. Industry (As, Cd, Cr, Co, Cu, Hg, Ni ,and Zn)
3. Atmospheric deposition (As, Cd, Cr Cu, Pb, Hg, U)
4. Agriculture (As, Cd, Cu, Pb, Se, U, Zn)
7

5. Waste disposal (As, Cd, Cr, Cu, Pb, Hg, Zn)


Anthropogenic activities such as mining, smelting, and agriculture have been widely implicated
in the increasing contamination of the environment with heavy metals (Navarro et al., 2008;
Brumelis et al., 1999; Vaalgamaa and Colney, 2008). A contaminant refers to a material which is
present in a part of the environment where it is not expected; when the material begins to cause a
problem, having a negative effect on health of humans or other organisms in the area, or causing
widespread changes in the local environment (Kebbeku and Mitra, 1998).Table 2.2 below list
trace metals according to their origin (Sparks, 2005):

Table 2.2 Natural and anthropogenic sources and common forms in wastes of trace metals
on the priority pollutant list
Element

Natural source or metallic

Anthropogenic sources

Common forms in

mineral

wastes

Ag

Free metal (Ag),


chlorargyrite (AgCl),
acanthite (Ag2S), copper,
lead, zinc ores.

Minning, photographic
industry

Ag metal, Ag-CN
complexes, Ag
halides,
Ag thiosulfates

As

Metal arsenides and arsenates,


sulfide ores (arsenopyrite),
arsenolite (As2O3), volcanic
gases, geothermal springs

Pyrometallurgical industry,
spoil heaps
and tailings, smelting, wood
preserving, fossil fuel
combustion, poultry manure,
pesticides, landfills

As oxides
(oxyanions),
organo-metallic
forms, H2AsO3CH3
(methylarsinic
acid),
(CH3)2-AsO2H
(dimethylarsinic
acid)

Be

Beryl (Be3Al2Si6O18),
phenakite (Be2SiO4)

Nuclear industry, electronic


industry

Be alloys, Be metal,
Be(OH)2

Cd

Zinc carbonate and sulfide


ores, copper carbonate
and sulfide

Mining and smelting, metal


finishing, plastic industry,
microlectronics,
battery manufacture, landfills
and refuse disposal, phosphate
fertilizer, sewage
sludge, metalscrapheaps

Cd ions, Cd
halides
and oxides, Cd-CN
complexes,
Cd(OH)2
sludge

Source: Sparks (2005)


9

2+

Table 2.2: Contd


Cr

Chromite (FeCr2O4),
eskolaite (Cr2O3)

Metal finishing, plastic


industry, wood treatment
refineries, pyrometallurgical
industry, landfills, scrapheaps

Cr metal, Cr oxides
3+
(oxyanions), Cr
complexes with
organic/inorganic
ligands

Cu

Native metal (Cu),


chalcocite (Cu2S),
chalcopyrite (CuFeS2),

Mining and smelting, metal


finishing,
microelectronics, wood
treatment, refuse disposal and
landfills, pyrometallurgical
industry, swine manure,
pesticides, scrapheaps,
mine drainage

Cu metal, Cu oxides,
Cu humic
complexes, alloys,
Cu ions

Hg

Native metal
cinnabar (HgS),
degassed from
Earth's crust and
oceans

Mining and smelting ,


electrolysis industry, plastic
industry, refuse
disposal/landfills,
paper/pulp industry,
fungicides

Organo Hgcomplexes Hg
halides and oxides,
2+
2+
0
Hg , (Hg2) , Hg

Ni

Ferromagnesian minerals,
ferrous sulfide ores,
pentlandite

Iron and steel industry,


mining and
smelting, metal
finishing,
microelectronics, battery
manufacture

Ni metal, Ni ions,
Ni amines, alloys

Source: Sparks (2005)

10

2+

Table 2.2: Contd


Pb

Galena (PbS)

Mining and smelting, iron


and steel industry,
refineries, paint industry,
automobile exhaust,
plumbing, battery
manufacture, sewage sludge,
refuse disposal/landfills,
pesticides, scrapheaps

Pb metal, Pb oxides
and carbonates,
Pb-metal-oxyanion
complexes

Sb

Stibnite (Sb2S3), geothermal


springs

Microelectronics,
pyrometallurgical
industry, smelting mine
drainage

Sb ions, Sb oxides
and halides

Se

Free element (Se), ferroselite


(FeSe2), uranium deposits,
black shales, chalcopyritepentlandite-pyrrhotite
deposits

Smelting, fossil fuel


combustion,
irrigation waters

Se oxides
(oxyanions), Seorganic complexes

Ti

Copper, lead, silver residues

Pyrometallurgical
industry,
microelectronics,
cement industry

Tl halides,
Tl-CN complexes

Zn

Sphalerite (ZnS),
willemite (Zn2SiO4),
smithsonite (ZnCO3)

Mining and smelting, metal


finishing,
textile, microelectronics,
refuse disposal and landfills,
pyrometallurgical industry,
sewage sludge, pesticides,
scrapheaps

Zn metal, Zn ions,
Zn oxides and
carbonates, alloys

Source: Sparks (2005)


11

3+

2+

2.3 The Soil solid phase


The soil is a dynamic system consisting of liquid, solid, and gaseous phases. The designation
phase according to Sposito (2008) means two things:
1. that it has uniform macroscopic properties (e.g. temperature and composition)
2. that it can be isolated from the soil profile, and investigated experimentally in the
laboratory.
The soil solid phase exerts the major effect on most soil properties and on the overall suitability
of soil as a plant growth medium (Bohn et al., 2001). The solid phase of the soil is what is
literarily implied by soil. It is composed of various components/fractions such as organic
matter, clays, Fe and Mn Oxides, lattice of silicate or carbonate minerals, which differ in their
ability to retain or release trace elements (Ure et al., 1993, Alloway, 1995; Tack et al., 1996).
The organic components, although usually present in smaller amounts than inorganic
components, affects soil properties significantly because of their high reactivity (Bohn et al.,
2001), due to the presence of numerous functional groups. Unlike the organic components, most
inorganic soil components of the soil solid phase are crystalline compounds of definite structures
called minerals (Bohn et al., 2001). These various geochemical solid fractions/components are
implied by the phrase, solid species, and they influence the mobility of metals in the soil
matrix. Trace elements seldom occur in soils as separate mineral phases, but instead are found as
constituents of these host minerals and humus (solid species); the chemical process underlying
these trace elements occurrences is called co-precipitation (Sposito, 2008). Trace elements occur
in primary minerals in different forms as presented in Table 2.3.

12

Table 2.3 Occurrences of trace elements in primary minerals


Element

Cr

Principal Modes of Occurrence in Primary Mineral

Chromite (FeCr2O4); isomorphic substitution for Fe or Al in other minerals of the


group

Cu

Sulfide inclusions in silicates; isomorphic substitution for Fe and Mg in olivines,


pyroxenes, amphiboles, and micas; and for Ca, K, or Na in feldspars.

Zn

Sulfide inclusions in silicates, isomorphic substitution for Mg and Fe in olivines,


pyroxenes and amphiboles; and for Fe or Mn in oxides

Cd

Sulfide inclusions and isomorphic substitution for Cu, Zn, Hg, and Pb in sulfides

Pb

Sulfide, phosphate, and carbonate inclusions; isomorphic substitution for K in


feldspars and micas; for Ca in feldspars, pyroxenes, and phosphates, and for Fe and
Mn in oxides.

Source: Sposito (2008)

13

2.4 Soil solution phase


The soil solution is the soil water plus its dissolved components which could be solid or gases.
With respect to dissolved solids, those that dissociates into ions (electrolytes) in the soil solution
are most important to the chemistry of soils (Sposito, 2008). Heavy metals in solution exist in
colloidal, particulate and dissolved phases (McBride, 1989; Adriano, 2001). The colloidal and
particulate metal may be found in (1) hydroxides, oxides, silicates and sulfides or (2) adsorbed to
clay, silicate or organic matter. The solubility of these metals is predominantly controlled by pH,
the type and concentration of ligands on which clay could interact, the oxidation state of the
metals and redox environment of the system (McBride, 1989). The dynamic equilibrium between
metals in soil solution and soil solid phase is determined by the properties of the soil and
composition of the soil solution. The equilibrium controls the availability of contaminants for
plant uptake. The sorption of metals on inorganic minerals and organic humic substances is one
of the most important geochemical processes controlling the fate, transport, and bioavailability of
metals in soil and water environments (Sparks, 2005).
Similarly, in soil solution, the behavior of metals is a function of the substrate sediment
composition, the suspended sediment composition, and the water chemistry (Essington, 2004).
The sediments of fine sand and silt generally have higher levels of adsorbed metals than quartz,
feldspars, and detrital carbonate-rich sediments (Kuo et al., 1986). The solution chemistry of the
soil system controls the rate of adsorption and desorption of metals from sediment; while metals
are removed from soil solution in adsorption reactions, they are returned to the soil solution
where recirculation and bio-assimilation may take place in desorption reactions (Danko, 2006).

14

2.5 Multiphase equilibria


The multiphase equilibria of metals within the soil system is important in understanding metal
behavior in soils (McLean and Bledsoe, 1992). The partitioning of gases between soil air and soil
solution is an important process contributing to the cycling of chemical elements in the soil
environment (Sposito, 2008). Metals in the soil solution are subject to mass transfer out of the
system by leaching to ground water, plant uptake, or volatilization, a potentially important
mechanism for Hg, Se, and As. At the same time metals participate in chemical reactions with
the soil solid phase. The concentration of metals in the soil solution, at any given time, is
governed by a number of interrelated processes, including inorganic and organic complexation,
oxidation-reduction reactions, precipitation/dissolution reactions, cation exchange, and
adsorption/desorption reactions. The ability to predict the concentration of a given metal in the
soil solution depends on the accuracy with which the multiphase equilibria can be determined or
calculated (McLean and Bledsoe, 1992).

2.6 Behavior of Heavy Metals in Soil


The behavior of heavy metals in soil is governed largely by their sorption and desorption
reactions with different soil constituents, especially clay components (Appel and Ma, 2002).
Chemical processes such as cation release from contaminated materials, and specific adsorption
onto surfaces of minerals and soil organic matter, and precipitation of secondary minerals
dictate the extent of these reactions (Manceau et al., 2000). The relative importance of these
processes depends on soil composition and pH. In general, cation exchange reactions and
complexation to organic matter are most important in acidic soils, while specific adsorption
and precipitation become more important at near- neutral to alkaline pH (Voegelin et al.,

15

2003). Oxidizing conditions generally increase the retention capacity of metals in soil, while
reducing conditions will generally reduce their retention capacity (McLean and Bledsoe, 1992).

2.7 Influence of Soil Properties on behavior of Heavy Metals


2.7.1 Soil pH
Changes in the soil pH influence the chemical reactions in the soil. At circum-neutral pH, Zn,
Ni and Pb have a strong relation with soil solids, and hence their downward movement will be
limited or very slow (Kabata-Pendias and Pendias, 1992). Colloid and metal mobility are
enhanced by decreases in solution pH and colloid size and increases in organic matter,
resulting in higher release of sorbed and soluble metal loads through a metal-organic complex
formation (Karathanasis et al., 2005). Soil Alkaline amendments reduce the concentration of
heavy metals in soil solution by raising the soil pH, thereby allowing the formation of insoluble
metal precipitates, complexes and secondary minerals (Mench et al., 1994). Also, heavy metal
cations form sparingly soluble phosphate, carbonates, sulfides and hydroxides, these sorption
reactions and many metal precipitation processes are highly pH dependent. The pH of the soil
system is often the most important chemical property governing sorption and precipitation of
heavy metals (Basta et al., 2005).

2.7.2 Organic Matter


The soil organic matter consists of a complex and varied mixture of organic substances. The
organic matter is an important factor in adsorption and accumulation of heavy metals in soil
(Alloway, 1995; Kabata-Pendias and Pendias, 1992). Chemical elements in the soil are more
adsorbed unto the finest soil particles (colloids) which are clay and humus; although they are

16

small in size, colloids play a major influence on soil properties. Heavy metal cations are easily
sorbed to the soil organic matter and other forms of humified natural organic matter. The type of
sorption by the organic matter affects the fate of the heavy metal in the environment.

2.8 Effect of heavy metals in the environment


2.8.1 Accumulation
The interest in heavy metals is due to their abundance in the environment (Atanassov et al.,
1999), which has increased considerably as a result of human activities. The fate of heavy
metals in polluted soils is a subject of many studies because of the direct potential toxicity to
biota and the indirect threat to human health via the contamination of groundwater and
accumulation in food crops (Martinez and Motto, 2000). Heavy metals are dangerous,
because they tend to bioaccumulate and biomagnify. Bioaccumulation is the concentration of a
chemical in a biological organism in higher amounts relative to the environmental
concentration (Kampa and Castanas, 2008) while biomagnification is the increase in the
concentration of a substance that occurs in a food chain as a consequence of:
1. Persistence
2. Food chain energetics, and
3. Low rate of internal degradation/excretion
Biomagnifications is the result of bioaccumulation. Bioaccumulation occurs when chemical
compounds are taken up faster than they are broken down; and plant cells have mechanisms for
bioaccumulation, selective absorption, and storage of a great variety of molecules. Heavy metal
pollution of soil enhances plant uptake, causing accumulation in plant tissues, eventual
phytotoxicity and change of plant community (Gimmler et al., 2002). Heavy metals
17

accumulating in soil enter the food chain, thus endangering herbivores, carnivores and
consequently, humans. Accumulation of heavy metals can also cause a considerable
detrimental effect on soil ecosystems, environment and human health due to their mobility and
solubility which determine their speciation (Kabata-Pendias and Pendias, 1992). Although
several studies have indicated that the crops grown on soils contaminated with heavy metals
have higher concentrations of heavy metals than those grown on uncontaminated soil (Nabulo,
2006; Abdu, 2010; Egwu, 2010), in environments with high nutrient levels, metal uptake can
be inhibited because of complex formation between nutrient and metal ions (Gothberg et al.,
2004). Therefore, a better understanding of heavy metal sources, their accumulation in the soil
and the effect of their presence in water and soil on plant systems seems to be particularly
important.

2.8.2 Solubility and Mobility


Heavy metal solubility and mobility in soils are of environmental significance due to their
potential toxicity to both humans and animals (Chirenje et al., 2003). Trace metal mobility is
closely related to metal solubility, which is further regulated by adsorption, precipitation and
ion exchange reactions in soils (Dong and Ma, 2004). Lead is reported to be the least mobile
among the other heavy metals, but Cd is known to be highly mobile under different soil
conditions (Kabata-Pendias and Pendias, 1992). The transfer of heavy metals from soils to
plants is dependent on three factors: (1) the total amount of potentially available elements
(quantity factor), (2) the activity as well as the ionic ratios of elements in the soil solution
(intensity factor), and (3) the rate of element transfer from solid to liquid phases and to plant
roots (reaction kinetics) (Brmmer et al., 1986). However, changes in soil solution chemistry,

18

such as pH, redox potential and ionic strength, may also significantly shift the retention
processes of trace metals by soils (Gerringa et al., 2001). These effects may be further
complicated by ligand competition from other cations (Norrstrom and Jacks, 1998). Reduction
in the redox potential may cause changes in metal oxidation state, formation of low-soluble
minerals and reduction of Fe, resulting in release of associated metals (Baumann et al., 2002;
Chuan et al., 1996). Metal solubility increases usually as the pH decreases, with the notable
exception of metals present in the form of oxyanions or amphoteric species. Since soil solution
properties might change with time, the solubility and speciation of metals might also be timedependent (Mo et al., 1999). As the soil pH increases, the solubility and availability of these
trace nutrients decrease (Mellbye and Hart, 2003). Immobilization of metals by the
mechanisms of adsorption and precipitation will prevent movement of the metals to the
ground water. Metal-soil interaction is such that when metals are introduced at the soil surface,
the downward transportation does not occur to any great extent unless the metal retention
capacity of the soil is overloaded or metal interaction with the associated waste matrix
enhances mobility (McLean and Bledsoe, 1992). Changes in soil environmental conditions over
time, such as the degradation of the organic matrix, changes in pH, redox potential, or soil
solution composition, due to various remediation schemes or to natural weathering processes
may also enhance the metal mobility (McLean and Bledsoe, 1992).

2.8.3 Bioavailability
Bioavailability refers to the ability of an element to be transferred from the soil to a living
organism (Kabata-Pendias and Pendias, 2001). An evaluation of the bioavailability of a metal is
used in predicting the risk of contaminant bioaccumulation and biomagnification (Chopin and

19

Alloway, 2007a,b). The mobility and bioavailability of metals to plants and their leachability to
contaminate ground water depend not only on the total concentration in solution but also on the
geochemical species or forms of the metals (Abdu et al., 2012). Evaluating bioavailability should
provide information regarding risk of contaminant transfer and accumulation into the food chain
(Kabata-Pendias and Pendias, 2001; Chopin and Alloway, 2007a,b), since total heavy metal
concentrations are not a good indicator of bioavailability (Wang et al., 2004; Ortiz and Alcaniz,
2006) because heavy metals can be associated with various soil components (Organic matter,
clays, Fe and Mn Oxides, lattice of silicate or carbonate minerals) that differ in their ability to
retain or release trace elements (Ure et al., 1993, Alloway, 1995; Tack et al., 1996).
Bioavailability depends on biological parameters and the physicochemical properties of
metals, and their species. Also, the bioavailability and mobility of metals in soil depend strongly
on the extent of their sorption with solid phases which is influenced by properties such as surface
area, surface charge, pH, ionic strength and concentration of complexing ligands (Petrovic et al.,
1999).
The pH and redox potential affect the bioavailability of metals in solution; at high pH,
the elements are present as anions, while at low pH the bioavailability of metal ions is
enhanced (Peterson et al., 1984). In natural systems, the bioavailability of trace metals is
primarily controlled by adsorption-desorption reactions at the particle-solution interface (Backes
et al., 1995). The soil pH will influence the dynamics of heavy metals. Generally, heavy
metals become increasingly mobile and available as the soil pH decreases (Tyler and Olsson,
2001).

20

2.8.4 Toxicity
The cumulative effect of low-levels of heavy metals encountered in agricultural soils posses a
potentially hazardous effect not only to crop plants but also to human health (Abdu et al.,
2011b; Das et al., 1997). Essential elements like copper (Cu), manganese (Mn), molybdenum
(Mo), and Zn, can also be toxic to plants at high concentrations in the soil. Some elements, not
known to be essential to plant growth, such as arsenic (As), barium (Ba), Cd, chromium (Cr),
Pb, nickel (Ni) and selenium (Se) are also toxic at high concentrations or under certain
environmental conditions in the soil (Slagle et al., 2004). Both pH and redox potential affect the
toxicity of heavy metals by limiting their availability (Peterson et al., 1984). At low pH,
metals generally exist as free cations; at alkaline pH, however, they tend to precipitate as
insoluble hydroxides, oxides, carbonates, or phosphates (Mamboya, 2007). Some metals, such
as Cr, As, Se and Hg can be transformed to other oxidation states in soil, thus influencing their
mobility and toxicity (McLean and Bledsoe, 1992). A lot of metals (Hg, Cd, Ni, Pb, Cu, Zn,
Cr, Co) are highly toxic both in elemental and soluble salt forms. The high concentration of
heavy metals in soils is toxic for soil organisms: bacteria, fungi and higher organisms
(Woolhouse, 1993). Short-term and long-term effects of pollution differ, depending on metal
and soil characteristics (Kadar, 1995; Nemeth and Kadar, 2005), the most serious consequence
of the higher metal concentrations is perhaps their negative effect on many of the decomposers
that live in the soil (Elvingson and Agren, 2004).

2.9 Analysis of heavy metal concentrations


Due to the association of trace metals with various solid fractions which determines their fate
within the soil environment (Ure et al., 1993, Alloway, 1995), quantifying the total concentration
of heavy metal in a soil matrix is not sufficient to infer on bioavailability or mobility (Wang et
21

al., 2004; Ortiz and Alcaniz, 2006), however, determining the total concentration of metals is
still important, as baselines for inferences on other procedures such as sequential extraction.
Sequential extraction is useful in assessing the potential hazard of toxic metals and metal
mobility in sediments (Atsuyuki et al., 2007). Chemical extraction is employed in assessing
operationally defined metal fractions, which can be related to chemical species, as well as to
potentially mobile, bioavailable, or ecotoxic phases of a metal. Operationally defined fractions
are defined by the methods used to separate them from other forms of the same element which
may be present (Kebbekus and Mitra, 1998). The mobile fraction is defined as the sum of the
amount dissolved in the liquid phase and an amount, which can be transferred into the liquid
phase (Hlavay et al., 2004). It has been generally accepted that the ecological effects of metals
(e.g., their bioavailability, ecotoxicology, and risk of groundwater contamination) are related to
such mobile fractions rather than the total concentration ( Brmmer et al., 1986).

2.9.1. Sequential Extraction/Fractionation schemes


Sequential extractions provide information about the probable origin, mode of occurrence,
biological and physicochemical availability, mobilization and transport of trace metals.
Sequential extractions have been achieved by fractionation of a sample using extractants with
progressively increasing extraction capacity. This eventually results in the removal of the
adsorbate phases and their associated metals. Several schemes have been developed to partition
species of the soil solid phase. The partitioning of metals obtained by such procedures is
operationally defined, which means that the extractant suffer from lack of selectivity (Filguerias
et al., 2002). Operational speciation entails the quantification and identification process of
different species, forms or phases of an element present in the sample, or the description of
22

amounts and types of them. The main goal of the studies on operational speciation is to convert
the metals bound in the solid phases into soluble forms with the extractant used at the each step.
Because sequential extraction is operational in manner, it is more appropriately called
fractionation (Filguerias et al., 2002).
Due to varying extraction conditions, similar procedures may extract a significantly different
amount of metals. Concentration, operational pH, solution/solid ratio, and duration of the
extraction may affect the selectivity of extractants (Hlavay et al., 2004). The resolution sought in
the chemical fractionation depends on the purpose of the study. However, the selectivity of the
procedure can be considerably improved by incorporation of the various nonselective single
extraction steps into a carefully designed sequential extraction scheme (Hlavay et al., 2004).
Fractionation procedures most commonly employed involve the chemical separation into
fractions such as
1. metal soluble in water and dilute acid medium
2. exchangeable
3. reducible or associated with Fe and Mn oxides
4. oxidizable or associated with organic matter and sulfur
5. residual or associated with silicates.
The concentration of metals obtained from each fraction are determined by the suitable analytical
technique, such as flame atomic absorption spectrometry (FAAS), electro thermal atomic
absorption spectrometry (ETAAS), inductively coupled plasma-atomic emission spectrometry
(ICP-AES), inductively coupled plasma-mass spectrometry (ICP-MS), anodic stripping
voltammetry, polarography, gas chromatography, 18 hydride generation-AAS, and neutron
activation analysis.

23

General applications of sequential extraction schemes can be summarized as follows:


(i) Characterization of pollution sources
(ii) Evaluation of metal mobility and bioavailability
(iii)Identification of binding sites of metals for assessing metal accumulation, pollution and
transport mechanisms.
Thus, sequential extraction results have proven useful in distinguishing between anthropogenic
and geochemical sources of most metal species in sediments. Sequential extraction schemes
provide information on the potential metal bioavailability and mobility of sediment-bound
metals, design of remediation processes, short and long term soil radionuclide and heavy metal
bioavailability.

2.10 Speciation calculation: Computer Models


The distribution of metal species and complexes in soil, natural, or polluted water bodies can be
measured by polarography, colorimetric determination, gas chromatography atomic absorption,
etc. These tools differentiate metals into their free ions and complexes. In their absence,
speciation can still be determined from total equilibrium metal concentration measured by
AAS/ICP. Results of total metal concentration, ligand concentration, and a pH are inputted into a
thermodynamic computer program which calculates speciation (McLean and Bledsoe, 1992).
Several equilibrium thermodynamic computer programs are available for modeling soil solution
and solid phase chemistry by providing information on the thermodynamic possibility of certain
reaction to occur. In addition to calculating the equilibrium speciation of chemical elements in
the soil solution and precipitation/dissolution reactions, models such as GEOCHEM (Sposito and
Mattigod, 1980), Visual MINTEQ (Gustafsson, 2011), SOILCHEM (Sposito and Coves, 1988),

24

and PHREEQC (Parkhurst and Appelo, 1999) provide information on cation exchange reactions
and metal ion adsorption. According to McLean and Bledsoe (1992), these models are used to:
1) calculate the distribution of free metal ions and metal-ligand complexes in a soil solution,
2) predict the fate of metals added to soil by providing a listing of which precipitation and
adsorption reactions are likely to be controlling the solution concentration of metals, and
3) provide a method for evaluating the effect that changing one or more soil solution
parameters, such as pH, redox, inorganic and organic ligand concentration, or metal
concentration, has on the adsorption/precipitation behavior of the metal of interest.

According to Sposito (2008), if the following are know:


1. Equilibrium metal concentration, plus a pH value
2. Conditional stability constants of all metal species and their complexes
3. Mass balance equations,
The distribution of metal species and complexes in solution can be calculated. The speciation
model assumes that mass balance expressions can be developed for each metal and ligand. These
expressions are converted into a set of coupled algebraic equations with free ion concentrations
as unknowns by substitution for the complex concentration. The algebraic equation with the free
ion concentration as unknowns can be solved numerically by standard algorithms based
estimated or guessed values. The resulting free ion concentrations are then used to calculate the
complex concentrations. These models are equilibrium models and as such do not consider the
kinetics of the reactions. Likewise, these models are limited by the accuracy of the
thermodynamic data base available (McLean and Bledsoe, 1992)

25

2.11 Complications from heavy metal pollution


Elevated concentrations of heavy metals adversely impacts on plant metabolism through
metabolic interferences related to enzyme synthesis, or DNA replication. Examples include
down regulation of essential protein synthesis, appurtenant to enzyme production, oxidative
damage of nucleotides, and mutagenesis (Hogan and Monosson, 2013). This usually manifests
as symptoms such as decrease in germination, increase in seedling mortality, inhibition of
growth rates and reduction in reproductive capability (Hogan and Monosson, 2013).
There are a wide variety of metabolic impacts of heavy metals upon animals (Merrill et al.,
2007). One of the fundamental issues in cell metabolic interference is the ability of some heavy
metals to mimic other essential nutrients in enzymatic processes. For example Pb often displaces
Zn or Ca in enzyme processes; these effects can lead to distortion of enzymatic activities (Hogan
and Monosson, 2013).
In humans Pb has long been known as a potent neurotoxin that can also damage other body
systems. Exposure to Pb can reduce IQ and cause behavioral problems. Acute exposure can
result in blindness, certain types of paralysis, seizures, encephalopathy, and death (Galadima et
al., 2012). Lead exposure in young children can have neurological consequences, preventing
sheaths from forming properly; it damages the central nervous systems, endocrine system,
kidneys and other internal body organs; exposure over long period of time results to death, also
the element and its compounds are toxic to fetuses and infants (Galadima et al., 2012).

26

CHAPTER THREE
3.0 MATERIALS AND METHOD
3.1 Description of Study area
The soils for this study were collected from the farmlands in Dareta village, Anka local
government area (LGA) of Zamfara State, northern Nigeria (Fig 3.1). Anka is located on
12o06o300N 5o5600E and has an area of 2,746 km2 and a population of 142,280 at the time of
the 2006 census (National Population Commission 2000). The climate of Zamfara is a Sudan
savannah with a mean annual rainfall of ~579 mm. The soils are formed from basement complex
rocks and are generally classified as ferruginous tropical lithosols according to Food and
Agricultural Organization classification (Olorunmi, (n.d.) Retrieved March 2, 2014, from
http://www.fao.org/ag/agp/AGPC/doc/Counprof/nigeria/nigeria.htm ).

Figure 3.1 Map of Nigeria, showing the location of Dareta Village, where soils for the study
were obtained
(Source: http://www.businessweek.com/investor/gold_boom/graphics/nigeria/).

27

3.2 Soil sampling


Surface soil samples were collected from the farmlands at depth of 020 cm. Farmlands on each
cardinal point of the village were sampled. Sampling proceeded along the four cardinal axes
(north, south, east, and west), starting from the first farmland after the last house, on a grid at
distances of 10, 30, 50 150, 300, 500, and 1,000 m. Composite samples were collected by
random sampling procedure in two replicates. Profile pits, one for each cardinal point of the
village, were also dug. Composite soil samples were also taken from each horizon of the soil
profile in duplicate up to a maximum depth of 3 m. The sampling proceeded from the bottom
horizon upward to maintain horizon boundary delineations and identity. The samples were
collected into polythene bags and labeled appropriately. The samples were air dried, crushed and
passed through a 2-mm mesh sieve before analysis.

28

Figure 3.2 Field Sampling Layout

29

3.3 Analytical Procedures


3.3.1 Determination of Physicochemical Properties
The physicochemical properties of the soils were determined in order to understand the general
properties of the soil. Particle size distribution was determined by the hydrometer method (Gee
and Or, 2002). Cation exchange capacity was determined using ammonium acetate (NH4OAc)
saturation (Rhoades, 1982). Soil pH was measured in 1:2.5 Soil/Water suspensions. Electrical
conductivity (EC) was determined using the Conductivity Measuring Bridge, Type M. C. S.
Model EBB/10. Organic carbon was determined by the acid-dichromate method of Walkley and
Black (Nelson and Sommers, 1982) and carbonate was determined by titration with 0.1N NaOH
(Agbenin, 1995).
3.3.2. Determination of Total Phosphorus (P)
Total P was determined by digesting 0.2g of finely ground soil sample with a mixture of 1ml of
60% HClO4 (v/v), 5ml of conc. HNO3 and 0.5 ml of conc. H2SO4 as described by Lim and
Jackson (1982). The soil in the digestion tube was gently swirled and digested at moderate heat
with gradual increment in temperature until white fumes appear. The tubes were allowed to cool
and diluted to 40ml with distilled water. The digest was transferred into a 50 ml volumetric flask
by filtering through Whatman No. 41 filter paper, and made up to the 50 ml mark with distilled
water. A blank digestion was made in the same way. Phosphorus in the digest was determined
using the method of Murphy and Riley (1962). A suitable aliquot of about 0 to 40 ml of the
working standard was taken into a 50-ml volumetric flask giving a standard of 0.04 mg P. 5
drops of 0.25 % P-nitrophenol indicator was added. 8ml of the developing reagent was added,

30

and the mixture was made to volume. The flask was shaken and left for 10mins. Absorbance was
read at 712 mm Bosch and Lomb Spectrometer.
3.3.3. Determination of Total Pb, Cd, and Zn
Total concentration of Pb, Cd and Zn in the soil was determined by Atomic Absorption
Spectrometer (AAS), after HNO3, HF and HCl digestion of a ground soil sample. 0.2g of airdried and finely ground soil was weighed into a glass beaker and digested with a 10:3:2 mixture
of HNO3 HF, and HCl, respectively, at 80C. The beaker was swirled gently and digested on a
digestion block until brown fumes appeared which subsequently turned white (indicating
complete digestion). The digested soil sample was set aside to cool and then filtered through
Whatman No. 42 filter paper and diluted to 50 ml with distilled water in a clean plastic vial prior
to analysis. Concentrations of Pb, Cd, and Zn in the filtrate were determined by AAS.
3.3.4. Determination of labile metal concentrations: CaCl2 extraction
The labile concentrations of Pb, Cd, and Zn were extracted by saturating 10 g soil with 20 mL of
0.01M CaCl2, and shaking the mixture for 16hrs. The soil suspension was filtered through
Whatman No. 42 filter paper. The concentrations of Pb, Cd, and Zn in the filtrate were read with
AAS.
3.3.5. Sequential Extraction
The method modified from Tessier et al. (1979) was used to extract operationally defined
fractions of Pb, Cd and Zn. The fractions analyzed include: exchangeable (EXCH), carbonate
(CARB), Fe and Mn Oxides (OX), organic matter (OM), and Residual (RES) fractions. The
Residual fraction (RES) was calculated by the difference between the concentration of total

31

metal and the sum of the first five fractions. The concentrations of Pb, Cd, and Zn were
determined using AAS.
3.3.5.1 Exchangeable (EXCH):
Two grams soil was weighed into a plastic bottle. The soil was extracted by adding 20mL of 1M
NH4OAc (pH 7.0). The mixture was agitated continuously for 30 minutes at room temperature
(20-25OC) and centrifuged. The clear supernatant was decanted into a clean plastic vial.
Concentration of Pb, Cd, and Zn in the supernatant were determined using AAS.
This represents the exchangeable fraction of Pb, Cd and Zn in the soil.

3.3.5.2 Bound to Carbonate (CARB):


The residue from the exchangeable fraction was extracted by adding 20 mL 1 M NaOAc
(adjusted to pH 5.0 with HOAc). The mixture was agitated continuously for 5hrs at room
temperature. The mixture was centrifuged and supernatant was transferred into a clean plastic
vial. Content of Pb, Cd, and Zn in the filtrate were determined using AAS.
This represents the fraction bound to carbonates.

3.3.5.3 Bound to Fe and Mn oxides (OX):


The residue from the carbonate fraction was extracted by adding 20 mL 0.04 M NH2OH.HCl in
acetic acid. The temperature of the mixture was raised to 96oC, by heating in a water bath for
6hrs, with occasional agitation. The mixture was centrifuged and the supernatant was transferred
into a clean plastic vial. Contents of Pb, Cd, and Zn in the supernatant were determined using
AAS.
This represents the fraction bound to Fe and Mn oxides.

32

3.3.5.4 Bound to Organic Matter (OM):


The residue from the Fe oxides fraction was extracted by adding 5 mL 0.1 N HNO3 and 10 mL 30%

H2O2. The temperature of the mixture was raised to 85oC by heating in a water bath for 5hrs,
with occasional agitation. The mixture was allowed to cool. After cooling, 15 mL of 3.2 M
NH4OAc in 20% HNO3 was added, and shaken continuously for 30mins at room temperature.
The mixture was centrifuged and clear supernatant was transferred into a clean plastic vial.
Content of Pb, Cd, and Zn in the supernatant were determined using AAS. This represents the
fraction bound to organic matter.

3.3.5.5 Residual Fraction (RES)


Residual fraction (RES) was calculated from the difference between the concentration of total
metal and the sum of the first five fractions.
RES = MT (EXCH + CARB + OX+ OM)
MT= total metal concentration

3.3.6 Sorption and solubility study of added Pb, Cd, and Zn


Sorption and solubility study was carried out to evaluate the extent of metal retention by the soil
as well as to evaluate the effect of adjustment of pH on metal retention by the soil, following the
procedure modified from Egwu and Agbenin (2012). The soils from 020cm depth were
equilibrated with stock solution of Pb(NO3)2, Zn(NO3)2.6H2O, and Cd(NO3)2 .4H2O solution in
0.01M KNO3 and left for 12 days. The stock solutions were prepared by dissolving 1.1g of Pb as
Pb(NO3)2, 0.6g of Cd as Cd(NO3)2.4H2O, and 1.8g of Zn as Zn(NO3)2.6H2O in 0.01M KNO3 to
give concentrations of 700 ppm Pb , 200 ppm Cd and 500 ppm Zn. One hundred grams (100g)
33

of the air-dried soil was weighed into plastic containers and replicated twice. The soils were
incubated with the individual metals from the stock solutions at a soil:solution ratio of 1:1 . The
pH of the soil suspension was adjusted to between 4 and 9 by adding appropriate amounts of
different concentrations of either NaOH or HNO3. The pH of the soil suspensions was
occasionally measured with a pH meter until desired pH values were obtained. The soil
suspensions were agitated and allowed to equilibrate for 12 days (d) after which aliquots of the
soil suspension were taken with the aid of syringes to determine pH; concentrations of Pb, Cd,
Zn, Ca, and Mg, were determined using atomic absorption spectrometry; while K and Na were
determined using flame emission photometer. Inorganic ligands such as SO42- , PO42- , HCO32- ,
and Cl- were also determined. Sulfate was determined by the turbidity method (Rhoades 1982),
phosphate by the method of Murphy and Riley (1962) while hydrogen carbonate (HCO 3) was
determined by titration with standard HCl (Rhoades 1982), chloride by direct potentiometric
titration with AgNO3 as described by Rhoades (1982). The metal sorbed by the soil was
calculated as the difference between metal added to the soil and metal remaining in the
equilibrium soil solution (Lo et al., 2009):

Where Madsorbed (%) is the adsorbed metal


Mtotal is the total metal content added to the soil
Msolubility is the equilibrium metal concentration remaining in solution

Soil solution speciation calculations were performed using PHREEQC Version 2 (Parkhurst and
Appelo, 1999). Input data for the model include concentration of cations and ligands, pH,
temperature, alkalinity, and density of water. Output data consist of free Pb2+, Cd2+ and Zn2+
34

concentrations and their complexes in the soil solution. From the metal sorption and free metal
ions concentration, the partitioning of the heavy metal between the solid and the solution phases
of the soil was measured and assessed in terms of the partitioning coefficient Kd (g L1)
Kd = [M]S / [M]L
Where [M]S is the amount of metal sorbed per unit weight of soil
[M]L is the equilibrium concentration of metal solubilized
3.3.7 Solubility of Pb, Cd, and Zn.
The maximum (100%) water holding capacity (WHC) of the soil was measured as the
gravimetric water content of the soil saturated and allowed to drain over 6 hrs (Fierer and
Shiemel, 2002). The soil samples were then brought to 70% of their WHC. Throughout the
course of the experiment, all the soils were adjusted to the same moisture/water content and left
for 1,7,14, and 21days at a temperature of 30 2 oC. The water level of the soil within this period
was maintained by periodic weighing of the sample after the initial saturation and ensuring
constants weight over the incubation period via drop-wise addition of predetermined amount of
distilled water to the centre of the sample to replenish lost moisture. The experiment was done in
three replicates including a control that was permanently dry. At the end of incubation period, an
equivalent of 2g oven dry sample weight of the moist sample was weighed and extracted with 50
ml of H2O for 2 h by shaking in a reciprocal shaker. The mixture was centrifuged and filtered
into plastic containers. The residue from the water extraction was extracted with 50 ml of 1 M
NH4OAc for 2 h to estimate the exchangeable amounts of heavy metals (Schlichting et al.,
1995). At the end of this procedure, the mixture was centrifuged and the supernatant filtered. The
concentrations of Pb, Cd, and Zn in the filtrate were determined by AAS

35

3.3.8. Effect of changing redox condition on the solubility of Pb, Cd, and Zn
Ten grams (10g) of soil was incubated under saturated condition with 1% and 3% glucose and
left to equilibrate for 7days. A control which received no glucose was also set-up. The pH of the
suspension was measured at the end of incubation period; likewise the soil redox condition was
measured using the combined pH/millivolt metre. The rate of metal dissolution was measured
from the changes in metal concentration in the soil solution. The suspension was filtered through
Whatman No 42 filter paper and the concentrations of Pb, Cd, and Zn in the filtrate were
measured using AAS.
3.4. Data Analysis
Analysis of variance (ANOVA) was used to determine significant differences in total metal
concentration, extractable metal concentrations, and the metal concentrations at the different
sampling intervals across the cardinal axes; correlation analysis was used to determine
relationships among various metals fractions. All statistical data analyses were performed using
Excel 2007 and SAS V9 (2002). Lead (Pb), Cd, and Zn mobility were assessed by the mobility
index (MI) (Liu et al., 2007).

MI=

Where Fi is the metal concentration in the EXCH fraction, Ti, the total concentration of the
metals, and n is the number of the soil samples.
Descriptive statistics technique was used to summarize the data of the various metal fractions.
Metal enrichment factor (EF) for Pb, Cd and Zn pollution was calculated based on the relation:
36

EF = (Pbsoil)/(Pbearth crust)

where Pbsoil is the Pb concentration in the soil and Pbearth crust is the average Pb concentration in
the earth crust, which is approximately 14 mg/kg (Sposito, 2008). The geoaccumulation index (Igeo) as proposed by Muller (1969) was calculated by computing the base 2 logarithm of the
measured total concentration of the metal over its background concentration using the
mathematical relation (Muller, 1969; Ntekim et al., 1993):
I-geo = log2 (Cn/1.5Bn)
where Cn is the concentration of metal n in the soil, Bn is the soil background
concentration of heavy metal n and 1.5 is a factor compensating the background data
(correction factor) due to lithogenic effects.

37

CHAPTER FOUR
4.0 RESULTS AND DISCUSSION
4.1 Soil physico-chemical parameters
The textural classes of the soils of the study area at depth of 0-20cm were majorly loam and
sandy clay loam, according to the USDA classification system (Table 4.1a). This indicates that
the soils of this region are suitable for agriculture, as most crops thrive best on loam textured
soils. Down the profile (Table 4.1b-e), the soil texture showed a textural change from sandy clay
loam through clay loam and loam, to clay. The CEC recorded for these soils ranged from 8.6 to
11cmolkg-1. The CEC values are strongly supported by the high clay content of the soils as well
as the pH which ranged from 6.14-7.45, indicating slight acidity to neutrality. Values for
electrical conductivity, EC which range from 0.002 to 0.05 dSm-1 did not indicate any
salt/salinity problem. Organic carbon of the soils was low and it ranged from 4.7 to 9.94 gkg-1,
this is similar to what was reported by Abdu et al. (2012) and Egwu (2010) for some soils in
northern Nigeria; although the values were considerably higher than the mean values for the
same region reported by Jones (1973). Total P values were low and ranged from 0.61 to 1.36
gkg-1, which is characteristic of a typical savanna soil of northern Nigeria

38

..

Table 4.1a. Physicochemical properties of the soil at Dareta, Zamfara state, northern Nigeria

Surface soil (0-20cm)


Sampling
Direction

pH

EC (dSm-1)

OC(gkg-1)

Carbonate
(gkg-1)

CEC
(cmol(+)kg-1)

Total P
(gkg-1)

Clay
(gkg-1)

Sand
(gkg-1)

Texture

Dareta North

6.14

0.05

9.92

3.53

10.6

1.12

259

428

Dareta South

6.47

0.04

7.34

3.89

8.72

1.36

255

460

Dareta East

6.23

0.01

8.32

4.09

9.68

0.82

251

532

SaCL

Dareta West

6.39

0.04

5.28

3.44

8.04

0.61

285

463

Texture Key: C-clay; CL-clay loam; SiC-silt clay; SiCL-silt clay loam; SaC-sand clay; SaCL-sand clay loam; SaL-sand loam

39

Table 4.1b Profile Characteristics at Dareta North


Depth

pH

EC (dSm-1)

OC(gkg-1)

Carbonate
(gkg-1)

CEC
(cmol(+)kg-1)

Total P
(gkg-1)

Clay
(gkg-1)

Sand
(gkg-1)

Texture

0-35cm

6.18

0.035

7.91

2.88

7.98

0.87

279

620

SaCL

35-75cm

6.3

0.01

7.34

2.7

8.31

0.75

279

660

SaCL

75-103cm

6.4

0.045

8.21

3.65

9.12

0.36

439

340

Cl

103-135cm

6.3

0.007

6.32

2.88

8.31

0.91

299

460

SaC

135-300cm

6.7

0.01

8.32

2.80

8.65

0.41

379

360

CL

Texture Key: C-clay; CL-clay loam; SiC-silt clay; SiCL-silt clay loam; SaC-sand clay; SaCL-sand clay loam; SaL-sand loam

40

Table 4.1c Profile Characteristics at Dareta South

Depth

pH

EC (dSm-1)

OC(gkg-1)

Carbonate
(gkg-1)

CEC
(cmol(+)kg-1)

Total P
(gkg-1)

Clay
(gkg-1)

Sand
(gkg-1)

Texture

0-24cm

6.4

0.04

8.45

3.90

8.96

0.76

339

240

CL

24-77cm

6.29

0.008

7.67

3.04

7.61

0.35

299

480

SaCL

77-103cm

7.45

0.035

8.93

4.19

8.21

0.52

439

340

CL

240

103-150cm
6.91
0.0025
9.34
3.52
9.87
0.41
299
Texture Key: C-clay; CL-clay loam; SiC-silt clay; SiCL-silt clay loam; SaC-sand clay; SaCL-sand clay loam; SaL-sand loam

41

Table 4.1d. Profile Characteristics at Dareta East

Depth

pH

EC (dSm-1)

OC(gkg-1)

0-23cm

6.6

0.03

23-59cm

6.79

59-103cm
103-200cm

CEC
(cmol(+)kg-1)
8.56

Total P
(gkg-1)
0.83

Clay
(gkg-1)
299

Sand
(gkg-1)
160

Texture

6.72

Carbonate
(gkg-1)
3.76

0.04

7.21

3.42

7.83

0.91

299

520

SaL

6.4

0.014

7.47

3.29

6.97

0.71

379

460

SaCL

6.77

0.01

6.31

3.64

6.31

0.54

319

460

SaCL

Texture Key: C-clay; CL-clay loam; SiC-silt clay; SiCL-silt clay loam; SaC-sand clay; SaCL-sand clay loam; SaL-sand loam

42

SiCL

Table 4.1e. Profile Characteristics at Dareta West

Depth

pH

EC (dSm-1)

OC(gkg-1)

Carbonate
(gkg-1)

CEC
(cmol(+)kg-1)

Total P
(gkg-1)

Clay
(gkg-1)

Sand
(gkg-1)

Texture

0-35cm

6.65

0.03

7.33

5.18

8.64

0.63

439

320

CL

35-77cm

6.54

0.045

8.42

4.91

9.11

0.77

219

560

SaC

77-103cm

7.2

0.009

5.21

4.58

8.31

0.91

239

460

103-200cm
6.83
0.056
6.12
3.65
7.83
0.54
359
240
Texture Key: C- clay; CL- clay loam; SiC- silt clay; SiCL- silt clay loam; SaC- sand clay; SaCL- sand clay loam; SaL- sand loam

43

SiC

4.2 Total concentration of Pb, Cd, and Zn


The total heavy metals concentration showed a very wide range, especially for Pb. The lowest
concentration of Pb (25.95mgkg-1) was recorded at Dareta north while the maximum
concentration of 2246.55mgkg -1 was recorded at Dareta west (Table 4.2). These observations
indicate extreme level of concentration considering the international threshold of 300 mgkg-1 for
Pb in arable soils (Table 4.3). This is a clear indication of Pb contamination and a possible health
risk. Of the four cardinal points along which sampling was proceeded, Dareta south recorded the
least Pb concentration in the soil matrix although the range of Pb recorded was far above the
threshold set by most regulatory bodies (Table 4.3).
Cadmium concentration in the soils was equally worrisome. The values ranged from
1.85 to 70.5

mgkg -1. These values were very much high when compared with standard

thresholds for arable soils and the mean concentration of Cd in the lithosphere (Lindsay, 1979;
Sparks, 2003). Although Dareta west recorded the highest concentration of Pb, the highest
concentration of Cd, 70.5 mgkg -1, was recorded along Dareta south while the lowest
concentration of 1.85 mgkg -1 was recorded along Dareta west. The concentration of Zn was
within the recommended threshold value of 300 mgkg-1 as set by the EU. The highest
concentration of 374 mg Zn kg -1 soil recorded at Dareta south, though slightly higher than the
threshold value of the EU, was still lower than the standard value of 1400 mgkg -1 set by the
regulatory bodies in Canada (Table 4.3). Although Zn contamination has been associated with
mining activities (Sparks, 2005), the results from this study did not find Zn to be a contaminant
in the farmland soils of Dareta. Above threshold value for the metal was only recorded at one
spot each in Dareta south and Dareta east where it exceeded the 300 mgkg -1 standard set by the
US and EU

44

Table 4.2 Mean concentration (mgkg-1) and standard error of Pb, Cd, and Zn in Dareta

Sampling
Axis

Pb

Cd

Zn

Dareta North

381 138.16c

27.76 10.19b

172.77 33.37b

Dareta South

424.2 113.15b

40.68 11.05a

193.98 47.60a

Dareta East

578.88 243.87a

39.48 10.13a

154.56 44.65c

Dareta West

555.69 295.98a

15.94 0.79c

145.5 7.20c

Means followed by the same letters are not statistically different at 5% probability level

45

Table 4.3 International threshold values for heavy metals concentration in soils (mgkg-1).

Heavy
Metal

EU

Regulatory system
USA
Canada

UK

Cd

3.0

3.0

19.5

1.4

Zn

300

200-300

1400

200

Cr

180

400

1500

6.4

Cu

140

80-200

170

63

Pb

300

300

150

70

Ni

75

50-110

210

50

Source: Abdu (2010).

46

4.2.1 Enrichment and accumulation of Pb, Cd and Zn


To better quantify or describe the level or risk of contamination from any environmental setting
many indices have been developed; most of them based on the total concentration of the element
of interest in the soil relative to the lithosphere. Enrichment factors (EF), accumulation indices
and various mobility indices have been widely used in assessing risk of contaminants in the
environment (Hazelton and Murphy, 2007; Banat et al., 2005; Sposito, 2008). The highest
concentration of Pb recorded in the soils of Dareta (2246.55 mgkg-1) was over 200 times greater
than the average worldwide concentration of 10 and 14-16 mg Pb mgkg-1 reported by Bowen
(1979) and Sposito (2008) respectively. The enrichment factor is the ratio of soil to crustal rock
concentration of an element; it gives a quantitative measure of the relative enrichment (or
depletion) of an element in soil as compared with parent material. Sposito (2008) gave metal
enrichment factor values of between 2 and 10 as indication of some enrichment of the element
under consideration while EF values of < 0.5 and >10 would indicate significant depletion and
strong enrichment respectively.
The calculated metal EF for Pb for the soils of Dareta ranged from 25.4 to 37.4 across the
four cardinal directions implying strong superficial enrichment of the element as a result of
anthropogenic activities. Similar observation was made for Cd which gave EFs between 79.7 and
203, while Zn only indicated some level of enrichment compared to the average rock (parent
material) concentration as its calculated values ranged from 1.8 to 2.4.
The calculated I-geo for the soils of this study ranged from 0.2 to 6.1 for Pb, 3.2 to 7.8 for Cd,
and -0.1 to 1.4 for Zn. According to the interpretations for pollution indices shown in Table 4.4

47

the highest grade, 6, reflects a 100-fold metal concentration relative to background values. (Abdu
et al., 2011c: Banat et al., 2005).

Although EF and I-geo allow for quick inferences on the status of a metal relative to mean soil
and lithosphere concentration, it is important to point out that these indices are based solely on
total concentration of a metal in the environment which is a poor indicator of pollution. Hence
they are not strong indicators for making inferences on contamination. A more reliable indicator
is the mobility index which considers the weakly adsorbed fractions alone, after a sequential
extraction is done. The results for mobility indices of the elements considered in this study are
discussed in section 4.7. Contrary to EF and I-geo factors which revealed highest grades for Pb
and Cd, the calculated mobility indices showed Zn to be the most mobile element in Dareta
(section 4.7), followed by Cd. Although lead recorded the highest total metal concentration, the
element was comparatively the least mobile. This shows that the mobility index is a more
accurate inference on the mobility of these metals and its results are strongly supported by the
fractionation results (section 4.6) that allotted highest proportion of all the elements studied to
the residual fraction. A summary of the mean Pb, Cd, and Zn concentration including their EF
and I-geo index is presented in Table 4.5.

48

Table 4.4 Measurements of metal pollution in soils and sediments


Index of geochemical
accumulation

I-geo class

Designation of soil quality

10-5

Extremely contaminated

4-5

Strongly/extremely contaminated

3-4

Strongly contaminated

2-3

Contaminated

1-2

Moderately contaminated

0-1

Uncontaminated/moderately
contaminated
Uncontaminated

Source: Abdu et al. (2011c).

Table 4.5 Average and background concentration, enrichment factor, calculated I-geo
index, and grade of pollution intensity of Pb, Cd and Zn in analyzed samples from Dareta.

Heavy
metal

Average
value
(mg/kg)

*Background
Concentration
(mgkg-1)

EF

I-geo

I-geo
grade

Pollution
intensity

Pb

491.3

15

32.8

4.4

Strongly/extremely
contaminated

Cd

16.6

0.2

83

5.8

Extremely
contaminated

Zn

164.8

80

2.1

1.4

Uncontaminated to
moderately
contaminated

*Lindsay (1979)

49

4.3 Horizontal Distribution of Pb, Cd, and Zn in Dareta.


Statistical comparison of heavy metal concentration across sampling intervals showed statistical
independence for most of the samples especially Pb and Zn (Figures 4.1 - 4.4). The few
statistical similarity observed were for Cd at Dareta East and Dareta west within the first 150 m
of sampling. There was no uniform trend in the distribution of the metal, but high concentrations
were observed within the first 150m distance (from starting point), although similar
concentration or even higher were observed much further (Figures 4.1 - 4.4). Highest
concentration of Pb was observed at 30 m away from the village at Dareta North and Dareta
south while similar observation were made at 50 m away from the village at Dareta East and
Dareta west. Highest concentration of Cd was recorded at 10 m away from the village in Dareta
North and Dareta East (Fig. 4.1, 4.4). However, maximum concentration for Zn at Dareta West
was observed at 1000 m away from the village (Fig. 4.4). The higher concentration of metals
recorded within the first 150 m of sampling suggest a higher level of enrichment of the soils
close to the ore processing points; but in general, the wide heterogeneity observed for the
samples is characteristic of environmental investigations, in which the parameters observed are
highly influenced by unquantifiable factors, such as soil-forming factors, which must have
influenced the distribution of the metals at the study location. This assertion was fully
investigated through a sequential extraction procedure which divided the soil solid into
operational fractions. Results of this procedure are discussed in section 4.6.

50

Figure 4.1 Mean concentrations of Pb, Cd, and Zn across the sampling distance in Dareta
North

Figure 4.2 Mean concentrations of Pb, Cd, and Zn across the sampling distance in Dareta

South
51

Figure 4.3 Mean concentrations of Pb, Cd, and Zn across the sampling distance in Dareta

East

Figure 4.4 Mean concentrations of Pb, Cd, and Zn across the sampling distance in Dareta

West
52

4.4 Vertical distribution of Pb, Cd, and Zn.


The concentration of Pb decreased from the surface (615 - 334.75 mgkg-1) to the subsurface
(572.1 - 167.55 mgkg-1) horizons except for Dareta west, where the concentrations within the
first two horizons were similar (335.5 and 348.3 mgkg -1) (Figure 4.5). At Dareta north, Dareta
South, and Dareta East, Pb concentration down the profile showed double depths of maximum
concentration; one maximum at the surface horizon (473.1, 342.75 and 615.9 mgkg -1) and the
other occurring at the third horizon (998.1, 316.68 and 573 mgkg -1). The concentration of Pb in
the bottom horizon in Dareta west (481.95 mgkg -1) and Dareta north (471.15 mgkg -1) likewise
showed an accumulation although the values recorded were less than those of the overlying
horizons (Figure 4.5). The trend of a higher concentration of heavy metals in the top soil than
subsoil can be attributed to superficial enrichment through human activities such as mining. It
could also reflect metals affinity for OM (Agbenin, 2002). Lead fixation by OM has been
reported to be more important than fixation by hydrous oxides (Li and Shuman, 1996); and the
surface horizons of most soils contain higher OM relative to the successive lower horizons.
The elevated concentration of Pb observed at the bottom horizons at Dareta west and Dareta
north, relative to the overlaying horizons could be attributed to enrichment through pedogenic
transformation of the underlying parent material as well as to leaching and mobility of metals
from overlying horizons, and probably as a result of capillary rise of Pb polluted ground water.
The source of pollution of the ground water is most likely from the washing of metal ores in
streams and household wells. Where the water saturating this horizon is polluted, it could result
in accumulation of metals at this zone relative to other zones; hence contamination from a
receding groundwater table could be suspected.

53

Figure 4.5 Vertical distribution of Pb down the profile, in Dareta

54

At the third horizon (77-103cm) where the concentration of Pb was also observed to increase
(Fig. 4.5), this could imply facilitated transport of colloidal particles (Mclean and Bledsoe, 1992)
which resulted in significant Pb movement and re-deposition within the profiles. These particles
include iron and manganese oxides, clay minerals and organic matter. Their surfaces have a high
capacity for metal sorption (Mclean and Bledsoe, 1992) and the majority of metals partition onto
particulate matter such as clay fraction (Eggleton and Thomas, 2004). Based on the observation
that the third horizon contained more clay (Table 4.1), it could be suggested that Pb was
translocated via facilitated transport with the colloidal clay particles during elluviation, hence redeposition of clay was accompanied by deposition of Pb. Furthermore, the second horizon
contained more sand (Table 4.1) compared to the underlying horizon. Abdu et al. (2012) showed
that heavy metal movement can be facilitated by sandy textures, likewise Gschwend and
Reynolds (1987) reported significant mobility of colloidal particles in a sandy medium. Similar
trend for Pb was observed by Agbenin (2002), although his observation was partly attributed to a
possible discontinuity in the profiles (Agbenin, 2002); similar inference cannot be made on these
soils because no sign of lithologic/profile discontinuity was observed.
The trend for Cd down the profile (Fig 4.6) was similar to that observed for Pb at the
bottom horizon (Fig. 4.5). Similar attribution to pedogenesis as made for Pb, can be made for this
elevated concentration. Except at Dareta north where there was a sharp decrease in Cd
concentration between the first (12.25 mgkg -1) and fourth (1.95 mgkg -1) horizons, Cd
concentration showed a relatively uniform distribution down the profile. The fairly uniform Cd
distribution from depth of 30 cm downwards suggests relative immobility of Cd in this soil.

55

Figure 4.6 Vertical distribution of Cd down the profile, in Dareta

56

The pH range of the soils of this region (6.14-7.45) also supports the immobility assertion; as it
has been suggested previously by Abdu et al. (2011a) for some garden soils in Kano that
exhibited a pH range of 6.9 to 7.4.
The distribution of Zn showed a major increase in Zn concentration only at the third
horizon (68.25 to 197.7 mgkg -1) (Fig. 4.4). This is similar to the observations for Pb and Cd. This
result could as well be ascribed to Zn partitioning unto clay fractions (Eggleton and Thomas,
2004) and subsequent translocation and illuviation (as indicated by the higher clay content at the
horizon). However, because the total Zn concentration recorded at the study location (66.45
330 mgkg-1) didnt indicate a possible pollution from Zn, this could be responsible for the lack
of any important observation on increased Zn concentrations (as observed for Pb and Cd) in the
last horizon studied.

57

Figure 4.7 Vertical distribution of Zn down the profile in Dareta

58

4.5 Extractable concentration of Pb, Cd, and Zn


The amounts of metals extractable by 0.01M CaCl2 represent quantities exchangeable with Ca
and the portion that form complexes with chloride ions (Egwu, 2010). These forms of metals are
labile and readily available for soil biota and uptake. Although no significant difference was
observed between extractable Pb concentrations recorded at Dareta north and Dareta South, these
values were significantly higher than the 2.4 mgkg-1 and 1.7 mgkg-1 recorded for Dareta East and
Dareta west respectively (Table 4.6). The percentage of the total Pb extractable by CaCl2 ranged
from 0.31 - 0.8. These low levels of extractable Pb in the soils indicate high insolubility of the
metal. Even though the total concentration was extremely high, only a small fraction of it was
labile and readily available. Therefore, the leaching of Pb and potential uptake of the metal by
plants is highly limited. Likewise, the potential for Pb accumulation in the food chain is
relatively low. The low levels of extractable Pb imply bulk partitioning of the metal to the nonexchangeable fractions of the soil and it also indicates a lithogenic origin of the contaminant.
Similar trends were observed for Cd and Zn, although lower concentrations of both metals,
compared to Pb, were extractable by 0.01M CaCl2 . Similar observations were made by Agbenin
(2002) who found that Pb was immobile in a Nigerian savanna soil under long-term cultivation.
This observation is further strengthened by the fact that no cases of Pb pollution were reported
from the study location until the heavy pedoturbation, through mining, began; thus the bulk of
the Pb content in the soil occur in the non-exchangeable fractions as shown in section 4.6.
Similar trend were observed for Cd and Zn, although lower concentrations extractable by 0.01M
CaCl2.

59

-1

Table 4.6 Mean (mgkg ) and standard error of 0.01M CaCl2 extractable concentration of Pb, Cd,
and Zn

Direction
Axis

Pb

Cd

Zn

Dareta North

3.33 1.66a

0.134 0.06b

0.504 0.15a

Dareta South

3.2 1.90a

0.5 0.12b

0.5 0.03a

Dareta East

2.4 1.37b

1.3 0.57a

0.5 0.07a

Dareta West

1.7 0.32c

1.8 0.67a

0.4 0.10a

Means followed by the same letter are statistically similar at p<0.05 (n=7)

60

4.6 Fractionation of Pb, Cd, and Zn


4.6.1. Lead (Pb) fractions
Fractionation results partitioned the bulk of the total concentration of Pb to the residual fractions
(Figure 4.8). The residual fraction represents metals associated with silicate clay minerals
(Hlavay et al., 2004). The relatively high percentage of the residual fractions in this soil indicates
a lithogenic origin of the metal contaminants (McLean and Bledsoe, 1992). In Dareta north, as
high as 89% of the total Pb concentration was bound to the residual fractions. The exchangeable
fraction of the soil held the least percentage (2 - 6%) of the total Pb concentration recorded. The
exchangeable metal ions measure those trace metals that are released most readily into the
environment; this fraction corresponds to the form of metals most available for plant uptake and
could be easily released following any change in the ionic strength of the soil solution medium
(Filguerias et al., 2002). Elements held onto the exchangeable fraction are easily exchanged by
other cations due to the weak electrostatic force of attraction in operation. The low proportion of
metals held on this fraction in these soils is an indication of a low risk of Pb being released into
the environment and also a low potential of food chain contamination. Fractionation results
reported by Rauret (1998) and Krishnamurti et al. (1995) also observed that the exchangeable
fraction usually represented a small fraction of the total metal content in unpolluted soils. The
proportion of the exchangeable fraction could also explain why no case(s) of Pb pollution was
reported before the artisanal mining began; because normal agricultural practices like tillage does
not perturb the soil to a significant level that would cause release of metals held unto the
residual fraction.

61

100

S u m o f le a d fra c tio n s (% )

80

60

40
EXCH
CARB
OXIDE
OM
RES

20

Dareta North Dareta South Dareta East

Dareta West

Figure 4.8 Distribution of Pb among the different geochemical fractions at the four study
directions at Dareta

62

The oxide bound fractions were the next most important fractions controlling Pb availability in
these soils. The percentage of Pb partitioned to this fraction ranged from 2.9 - 14.4. The oxide
bound fractions include Pb bound to FeO and MnO. Heavy metal desorption from this fraction is
majorly controlled by soil redox condition (Abdu et al., 2011; Filguerias et al., 2002), because
the oxide fractions are thermodynamically unstable under anoxic conditions (Filguerias et al.,
2002). The onset of reducing conditions would activate metal desorption from the Fe and Mn
oxides fraction. Therefore, this result suggests a positive relationship between the metal released
into the environment and changes in soil redox state.
A significantly high concentration of Pb fraction was held unto the OM fraction (24.9 %),
especially for soils sampled along the western coordinate. Organic matter has a high affinity for
heavy metals, forming stable complexes with them; especially the divalent ions, for which
organic matter exhibits a high degree of selectivity compared to monovalent ions (Filguerias et
al., 2002). Kennedy et al. (1997) observed that metallic pollutants associated with this phase
remains in the soil for longer periods but may be mobilized by decomposition processes (because
they are oxidizable). Hence, the degradation of OM under oxidizing conditions would facilitate
the release of Pb bound to this fraction (Filguerias et al., 2002).
The carbonate bound fractions ranged from 1.6 % to 18.2 % of the total Pb concentration
of the study area. Carbonate could be an important adsorbent for many metals when OM and FeMn oxides are less abundant (Stone and Droppo, 1996). The metal bound to carbonate is
sensitive to pH and is mobilizable when pH is lowered. Generally, heavy metals in the
exchangeable and carbonate fractions are readily and potentially bioavailable (Filguerias et al.,
2002).

63

4.6.2. Cadmium (Cd) fractions


The bulk of the total Cd was partitioned to the residual fraction (36-62 % of the total Cd),
followed by the exchangeable fraction which held 11-16% of total Cd (Fig. 4.9). The high levels
of exchangeable Cd compared to Pb represent a higher risk of Cd mobility in the soil. The
predominant occurrence of Cd as divalent cation makes electrostatic adsorption on exposed
adsorption sites an important bonding mechanism for the metal (de Haan and Zwerman, 1978).
Thus, Cd has a high potential to be released from the soil by simple ion exchange reactions
(Rogan et al., 2010), thereby representing potential phytotoxic and human hazards (Udom et al.,
2004; USEPA 2005) especially when it undergoes solubilization. However, the distributions of
Cd among the oxide, OM and carbonate fractions were similar to those of Pb. The carbonate
bound Cd was lower than that of Pb for the whole study area, however the oxide bound fraction
was next to the residual and exchangeable. The percentage of Cd partitioned to the oxide fraction
ranged from 2 % to 5 % for the whole study area.

64

S u m o f c a d m iu m fra c tio n s (% )

100

80

60

40
EXCH
CARB
OXIDE
OM
RES

20

Dareta North Dareta South

Dareta East

Dareta West

Figure 4.9 Distribution of Cd among the different geochemical fractions at the four study directions
at Dareta

65

4.6.3 Zinc (Zn) fractions


Organically bound fraction of Zn accounted for 17-25 % of total Zn (Figure 4.10). The strong
ability of Zn to form complexes with OM reduces Zn mobility and phytotoxicity in the
environment (Kashem et al., 2007), thereby making the risk of pollution from Zn in this region
minimal. The oxide fraction accounted for 31-55 % of the total Zn concentration. Under reducing
conditions, metals bound to this fraction become very important (Chao 1972). The association of
Zn with Fe-Mn oxide has also been reported in several other studies (Ma and Rao, 1997; Wilcke
et al.,. 1998; Kashem et al.,. 2007) some of which suggested that the association of Zn with the
oxide fraction reflects the high stability constant characteristic of Zn oxide. The major process
controlling metal release from this fraction is redox. The onset of reducing conditions enhances
desorption of metals held unto this fraction due to the reductive dissolution of Fe and Mn oxides
(Dong and Ma, 2004). Also, as the soil reaction becomes more acidic, the Zn metal cation is
easily displaced by more electronegative and higher valence cations such as H+ and Al3+,
respectively (McLean and Bledsoe, 1992). However, the relatively higher proportion of residual
fractions of Zn is similar to the observations for Pb and Cd, and it further suggest a lithogenic
origin of Zn in these soils.

66

100

S u m o f z in c fra c tio n s (% )

80

60

40
EXCH
CARB
OXIDE
OM
RES

20

Dareta North

Dareta South

Dareta East

Dareta West

Figure 4.10 Distribution of Zn among the different geochemical fractions at the four study
directions at Dareta

67

4.6.4 Correlation between metal fractions


Significant positive correlation was observed between the exchangeable Cd fractions and
carbonate bound Cd fractions (0.96*); similar levels of correlation were also observed between
the exchangeable Cd and the oxide (0.90*) and residual Cd fractions (0.88*) (Table 4.7; p
0.05). This suggests a strong buffering of the exchangeable Cd concentration by the oxide,
carbonate and residual fractions, thereby increasing the risk of Cd release into the environment.
However, for the Zn fractions only the residual fraction and the oxide bound fraction were
positively correlated (Table 4.8). This could imply that similar processes were most likely
controlling Zn release from both fractions and a possible transformation of Zn from the residual
to a more mobile oxide bound fraction.

68

Table 4.7 Correlation coefficient matrix between the different Pb and Cd fractions

EXCH

CARB

OXIDE

OM

RES

EXCH

-0.34

-0.25

-0.31

-0.67

CARB

0.96*

-0.63

0.145

-0.42

OXIDE

0.90*

0.85*

0.62

0.86*

OM

0.72

0.83

0.41

0.45

RES

0.88*

0.81

0.60

0.77

The lower left-hand section is correlation coefficient for Cd while the upper, right-hand part is correlation
coefficient for Pb; * , Significant at p0.05

Table 4.8 Correlation coefficient matrix between the different Zn fractions

EXCH

CARB

OXIDE

OM

EXCH

CARB

-0.34

OXIDE

-0.25

-0.63

OM

-0.31

0.15

0.62

RES

-0.68

-0.40

0.85*

0.45

*, Significant at p0.05

69

RES

4.7 Mobility of Pb, Cd, and Zn fractions


The mobility index (Liu et al., 2007; Banat et al., 2003) was used as a more accurate prediction
of the risk of contaminant transfer to the food chain or ground water. The calculated mobility
indices for the metals under study ranged from 0.46 - 1.76 for Pb; 3.21-3.66 for Cd; and 1.013.99 for Zn (Figure 4.11). These values showed that Zn was the most mobile in the soil. Lead
was the least mobile of the three studied metals. The observed order of mobility was Zn > Cd >
Pb. Although low concentration of Cd and Zn were observed to be associated with the available
pools, their high mobility indices may reflect potential risk of pollution through leaching and
plant uptake of metals.
4.8 Sorption of Pb, Cd, and Zn added to the soil
Sorption of metal cations is pH dependent and is characterized by a narrow pH range where
sorption increases to nearly 100%. This point is known as the adsorption edge (Sparks, 2003).
The pH range of the adsorption edge for a particular cation is related to its hydrolysis or acidbase characteristics. In addition to pH, sorption of metals is dependent on sorptive concentration,
surface coverage (loading), and the type of sorbent (Sparks, 2003).
As typical for all cationic metals, adsorption for the soils of Dareta increased with pH. Between
pH 4.8 and 5.8, the percentage of Pb sorbed reached 99% (Fig. 4.12). This pH range could be
considered as the adsorption edge for Pb in this soil. After this point, only relatively slight
increase in adsorption was observed. A peculiar feature of adsorption of metals in these soils was
the observed convergence of the adsorption lines of Cd and Zn from pH 7.3 onwards (Fig. 4.12).

70

Figure 4.11 Mobility indices for Pb, Cd, and Zn across the study directions in Dareta

71

Figure 4.12 Sorption of Pb, Cd, and Zn as a function of pH on the soils of Dareta

72

The adsorption pattern for Cd and Zn were similar to that of Pb however, the order of adsorption
was Pb > Zn > Cd across the entire pH range. The following order of relative affinity of soils for
metals: Pb > Cu > Zn > Cd > Ni had been reported (Biddappa et al., 1981); Zn > Ni > Cd (Tiller
et al., 1984). These are similar to what was observed for the soil of Dareta in the present study.
McBride and Blasiak (1979) showed increase retention of Zn with increasing pH, however
McLean and Bledsoe (1992) reported that maximum retention of cationic metals generally occur
at pH > 7. In the present study, maximum adsorption were observed just before pH 7. Many
adsorption sites in soils, especially the highly weathered tropical soils with which are the
predominant soil type in the study area are pH dependent, hence metal sorption is controlled
largely by changes in soil pH. The pH dependence of adsorption reactions of cationic metals is,
partly, due to the preferential adsorption of the hydrolyzed metal species in comparison to the
free metal ion (James and Healy, 1972; Davis and Leckie, 1978; Cavallaro and McBride, 1980);
and probably due to the effect of deprotonation of potential determining ions thereby creating
additional sorption sites onto which cations can adsorb. Lagerweff and Bower (1972) and Bittel
and Miller (1974) observed preferential adsorption of Pb2+ to Cd2+ at pH > 5. The observed
disparity in the pH at which maximum adsorption was attained by Pb, Cd, and Zn probably
reflect differences in the first hydrolysis constants (pK1) of the metals (Egwu, 2010) which is
about 7.71 for Pb2+, 8.96 for Zn2+ and 10.08 for Cd2+ (Essington, 2004).
The chemistry of metals, especially in terms of their adsorption and exchange behavior is altered
by hydrolysis. Since the hydrolysis species of metals are positively charged, it implies that these
species would adsorb readily to the negatively charged solid matrix. Where hydrolysis occurs at
a relatively lower pH, maximum adsorption would be attained quickly compared to a metal that
hydrolyses at a relatively higher pH. It therefore follows that high pH value for hydrolysis
73

implies low adsorption affinity whereas metals that hydrolyze at a low pH adsorbs strongly
(Sposito, 2008). However, the pH at which maximum adsorption was attained by Cd was lower
than its first hydrolysis constant (pK1) given as 10.08 (Essington 2004). This wide disparity
suggests an increasing organic matter (OM) complexation of Cd because of the increased surface
area and charge density, and also the solubility of OM at high pH (Egwu and Agbenin, 2012).
Almas et al. (2000), Suave et al. (2000), and Krishnamurti and Naidu (2003) extensively
demonstrated the role of OM in Cd complexation and this might explain the observed disparity.
Similarly, ionic potential and ionic radii may alter the sequence of adsorption of metal cations.
Metal cations with larger ionic radii will create a smaller electronic field and thus less likely to
remain solvated during complexation by surface functional groups. A larger ionic radius implies
a more labile electron configuration, a general tendency for metal cation to polarize in response
to electrical field of a charged surface functional group (Sposito, 2008). However, the observed
sequence of Pb > Zn > Cd contradicts the metal adsorption predictions based on ionic potential
and ionic radii. The manner of adsorption of Pb, Cd, and Zn demonstrated by the soils of the
study limits the risk of heavy metal leachability from long-term accumulation of heavy metals.
Therefore, in the non-arable soils around the goldmine, significant metal pollution control can be
achieved by simply maintaining the soil reaction at above neutral. One of the ways this could be
achieved is by liming.
4.9 Metal solubility as a function of pH
The relationship between soluble metal concentrations and pH after 2 weeks of equilibration
(aging) are presented in Figure 4.13. As expected, the solubility of Pb, Cd, and Zn increased with
decreasing pH (Fig. 4.13), almost linearly, up to pH of about 6.

74

Figure 4.13 Solubility of Pb, Cd, and Zn added to the soil of Dareta as a function of pH

75

However there was disparity on the pH at which solubility increased for Pb, Cd, and Zn. For Pb,
this was between pH 4.8 and 5.4 while for Zn it was between pH 5.4 and 5.8. Similar observation
was made for Cd between pH 4.8 and 6.6. The pH range at which solubility of metals increased
corresponds to the pH range of optimum crop growth. It therefore implies that over long period
of metal accumulation from aerial enrichment especially in the form of harmattan dust as well
as from irrigation of farmlands, the risk of metal accumulation in the food chain due to increased
plant uptake and also the pollution of ground water due to metal leaching are very high. The
differences among the pH values of maximum adsorption indicate different affinities for the
metals in soils and suggest a weaker reaction of Cd with soil constituents than Pb and Zn
(Martinez and Motto, 2000). Hence there is a higher risk of Cd accumulation in the food chain. A
number of researches have reported worrisome amounts of Cd in plants grown on arable fields
with no problem of metal contamination (Kirkham, 2006). This stresses the importance of
monitoring metal input into the soil. The nature of the solubility for Pb, Cd, and Zn observed in
this study corresponds with several other studies on metal interactions with soil solid (Martinez
and Motto, 2000; Cavallaro and McBride, 1980; Garcia-Miragaya, 1984; Agbenin and Olojo,
2004; Egwu, 2010). Given the nature of soils of this region (northern Nigeria), highly weathered
and dominated by low activity clays and variably charged surfaces, the decreasing metal
solubility with increasing pH is usually a consequence of the increasing charge density and the
hydrolysis of metal cations.
4.10 Heavy metals partitioning
Lead showed the highest metal partitioning coefficient (Kd) compared to other heavy metals
studied (Fig. 4.14). Metal partitioning coefficient (Kd) for Pb had a maximum value of
4995 g L1. Also, the metal partitioning of the three elements studied increased as pH increased
76

particularly at pH higher than 7 (Fig. 4.14). Metal adsorption versus metal solubility showed the
different distribution patterns for Pb, Cd, and Zn. Lead and Cd showed similar distribution
patterns with slightly wider adsorption versus lower solubility while that of Zn was narrower
(Fig. 4.14). Zinc showed an almost uniform distribution pattern across the pH range of the study.
Although calculated Kd values reflects adsorption, the values of soluble concentrations of the
metals studied suggest the occurrence of other sorption reactions such as precipitation. Some
researchers (Rickard and Nriagu, 1978) have suggested the possibility of precipitation of metals
when soluble concentration exceeds 4 mg L-1 at pH 4.

77

Figure 4.14 Partition coefficients (K d) for Pb, Cd, and Zn under varying pH

78

4.11 Solid phase-solution equilibria of Pb, Cd, and Zn added to soil


4.11.1. Lead solid-solution equilibria
Due to the toxicity of Pb to both man and animal, it is important to depict the solubility
relationship of Pb minerals and complexes in the soil environment. Santillan-Medrano and
Jurinak (1975) extensively used stability diagrams in predicting the formation of precipitates of
Pb and Cd in calcareous soils. The equilibrium reactions of selected minerals of Pb that could
probably form in Pb polluted soils are presented in Table 4.9. The activity isotherms presented in
this section were constructed by converting solubility relationships such as those in Table 4.9
into log terms and rearranging terms to form a straight line relationship. The solubility of Pb in
the soils of Dareta is indicated by the scatter data-points in Figure 4.15. The thick lines of
Pb(SO4) (anglesite), PbCO3 (cerrusite), and Pb5(PO4)3Cl (chloropyromorphite) represent the Pb
compounds considered to be regulating the activities of Pb in solution. The mixed
minerals/compounds of Pb that could form in soils decreases in solubility by the following order
(Lindsay, 1979):
Pb2CO3Cl2(s) > PbSO4.3PbO (s) > PbCO3.PbO (s) > PbCO3.2PbO (s) > PbSO4.PbO(s)
> Pb3(CO3)2(OH)2(s).
The solubility of these minerals makes them unlikely to control Pb2+ activities in unpolluted
soils.
The relative distribution of data points for Pb2+ activity with respect to pure minerals (Fig. 4.15)
does not suggest the formation of a pure mineral compound as controls of the solubility of Pb, as
it has been widely suggested by some researchers. There was no apparent equilibrium between
the free Pb2+ activities in solution and the pure mineral compounds considered. However, the
distribution of data points (Fig. 4.15) indicates that below pH 6 the soil solution was under79

saturated with respect to PbCO3. As pH increased towards 8, activity data points showed that the
free Pb2+ in the soil solution was super-saturated with respect to cerrusite (PbCO3). Pb5(PO4)3Cl
is highly stable at this point and could not be a likely candidate controlling the activity of Pb2+ in
solution. The increased solubility of Pb observed with decrease in pH could suggest the
possibility of Pb2+ activities being controlled by PbCO3. When soils of high pH become more
acidic in reaction, they may release fixed Pb especially when PbCO3 is involved (Bolt and
Bruggenwert, 1978) due to the high dependence of CO32- on pH. The control of above threshold
concentration of Pb by the formation of relatively insoluble compounds such as PbCO3,
Pb3(PO4)2, and PbSO4 has been described by Singer and Hansen (1969). Across all the pH of
observation, the solution activity of Pb was higher than for Cd and Zn, but that of Zn was slight.
This indicates that Pb may be more mobile in the environment over a long period of
accumulation (McLean and Bledsoe, 1992).

80

Table 4.9 Equilibrium reactions for Pb minerals and complexes likely to control Pb2+
activities in a Pb polluted soil.

Equilibrium Reaction

Log Ko

1. PbCO3 (Cerrusite) + 2H+ = Pb2+ + CO2 + H2O

4.65

2. Pb2CO3Cl2 + 2H+ = Pb2+ + 2Cl- + CO2 + H2O

-1.80

3. Pb(OH)2 + 2H+ = Pb2+ + 2H2O

-8.34

4. PbSO4.PbO + 8H+ =2Pb2+ + SO42- + H2O

-0.19

5. Pb3(PO4)2 + 4H+ = 3Pb2+ + 2H2PO4-

-5.26

6. Pb5(PO4)3OH(s) (Hydroxypyromorphite) + 7H+ = 5Pb2+ 3H2PO4- + H2O

-4.14

7. Pb5(PO4)3Cl + 6H+ (Chloropyromorphite) + 6H+ = 5Pb2+ + 3H2PO4- + Cl-

-25.05

8. FePO4.2H2O (Strengite) + 2H+ = Fe3+ + H2PO4- + H2O

-6.85

9. CaSO4.2H2O (Gypsum)= SO42- + Ca2+ + 2H2O

-4.64

10. CaCO3 (Calcite) + 2H+ = Ca2+ + CO2 + H2O

9.74

Source: Lindsay (1979)

81

Figure 4.15 Logarithm (base 10) of soluble Pb from the amended soil of Dareta and from
various pure compounds. Stability lines for minerals were calculated based on log K values
reported by Lindsay (1979).

82

4.11.2. Zinc solid-solution equilibria


Dissolution reaction of the minerals considered for the solubility equilibria of Zn under polluted
environment are presented in Table 4.10. The solubility of added Zn in the soils of Dareta are
indicated by the scatter data-points in Figure 4.16
According to Lindsay (1979) the solubility of Zn minerals likely to be encountered in soils, in
order of decreasing solubility, are as follows:
Zn(OH)2(amorph) > -Zn(OH)2 > -Zn(OH)2 > -Zn(OH)2 > -Zn(OH)2 > ZnCO3
(Smithsonite) > ZnO (Zincite) > Zn2SiO4 (Willemite) > soil-Zn > ZnFe2O4 (Franklinite).
The Zn(OH)2 minerals as well as zincite and willemite are too soluble to persist in soils hence
they cannot control the activities of Zn in unpolluted soils (Lindsay, 1979). Similar to the
observation for Pb, no apparent equilibrium between the activities of Zn2+ in solution and the
pure minerals considered could be gleaned from the results. However, at pH 8.3, activity data
point for Zn2+ was slightly above the solubility line of hopeite (Fig. 4.16), indicating supersaturation, while at lower pH values, activity data points for the soils fell below the solubility
line of hopeite. The relative positions of soil-Zn and franklinite (ZnFe2O4) renders both mineral
unlikely controls of Zn activities in solution.

83

Table 4.10 Equilibrium reactions for Zn-minerals and complexes likely to control Zn2+
activities in soils.

Equilibrium reaction

Log Ko

1. ZnO (zincite) + 2H+ = Zn2+ + H2O

11.79

2. ZnCO3 (smithsonite) + 2H+ = Zn2+ + CO2 + H2O

7.91

3. Soil-Zn + 2H+ = Zn2+

5.80

4. ZnFe2O4 (Franklinite) + 8H+ = Zn2+ + 2Fe3+ + 4H2O

9.85

5. Zn3(PO4)2.4H2O (hopeite) + 2H+ = 3Zn2+ + 2H2PO4- + 4H2O

3.80

6. Fe(OH)3 + 3H+ = Fe3+ + 3H2O

-2.70

7. FePO4.2H2O (Strengite) + 2H+ = Fe3+ + H2PO4- + H2O

-6.85

8. CaCO3 (Calcite) + 2H+ = Ca2+ + CO2 + H2O

9.74

Source: Lindsay (1979);

84

Figure 4.16 Logarithm (base 10) of soluble Zn from the amended soil of Dareta and from
various pure compounds. Stability lines for minerals were calculated based on log K values
reported by Lindsay (1979).

85

Based on the assumption that the solubility of H2PO4 in solution is controlled by strengite
(FePO4.H2O) the stability line for hopeite was established by the following relationship:
Log Zn2+ =11.10 -2pH, from information in Table 4.10. The stability line of ZnFe2O4 (willemite)
was constructed by the following relationship: Log Zn2+ = 4.45-2pH, after coupling reactions (4)
and (6) in Table 4.10 in order to obtain the dissolution reaction of Franklinite in the presence of
Fe(OH)3.
Willemite (Zn2SiO4) is of intermediate solubility and its solubility line falls below that of
hopeite, but the mineral is also too soluble to be considered as likely controlling Zn activities in
solution (Lindsay, 1979). The relative position of soil-Zn also makes the mineral an unlikely
control of Zn activities in solution, because it is expected that both minerals will precipitate out
of solution due to supersaturation. Although the dissolution of hopeite could be suggested it is
very likely that the Zn2+ maintains equilibrium with soil-Zn since hopeite is generally more
soluble than soil-Zn. The likely solubility depicted by soil-Zn can be accounted for by
franklinite, however the solubility of franklinite is further affected by the activity of Fe3+
(Lindsay, 1979). Hence the equilibrium levels of Zn2+ is dependent on the crystalline Fe(III)
oxides such as maghemite or goethite (Lindsay, 1979).
4.11.3. Cadmium solid-solution equilibria
No oxidation state of Cd is stable in the redox range of soils; hence the chemistry of Cd in soils
is limited to Cd(II) minerals and complexes (Lindsay, 1979). The solubility/equilibrium reactions
of the various Cd minerals considered are given in Table 4.11. In Figure 4.17, three Cd minerals,
including free soil-Cd, were considered as likely control of Cd activities in solution. The activity
of Cd2+ in the soil solution has been shown to be limited by CdCO3 (otavite) depending on the

86

CO2 (g) concentration. Hence an atmospheric CO2 pressure of 0.003 atm was assumed in
constructing the CdCO3 line (Fig. 4.17), since the CO2 in soils is generally higher than that of the
atmosphere. Three Cd minerals, including free soil-Cd, were considered as likely controls of Cd
activities in solution.
The relative distribution of the activity data points of Cd2+ with respect to the stability lines of
the pure minerals makes it tenuous to suggest the formation of any mineral as the controlling
mechanism of Cd2+ activity in the soil solution. Despite the various assumptions made while
constructing the solubility lines of the pure Cd minerals, no apparent equilibrium between Cd in
solution and the solid phase was discerned. Although McBride (1980), on the chemisorption of
Cd2+ on calcite surfaces, concluded that CdCO3 could precipitate at high pH values, and Santillan
and Medrano and Jurinak (1975) suggested that the formation of CdCO3 and Cd3(PO4)2 were
possible in Cd-contaminated soils, the distribution of the data points of the added Cd didnt
suggest the attainment of any sort of equilibrium between the Cd2+ activities in solution and the
pure minerals. Although high levels of Cd were added to the soil in order to simulate Cd
pollution from dust, sediment, and irrigation depositions over time, there was no clear evidence
to suggest the formation of either of CdCO3 and Cd3(PO4)2. Therefore, the observation from this
study suggests mechanism(s) other than formation of pure minerals as likely control of Cd
activities in the soil solution.

87

Table 4.11 Equilibrium reactions for Cd minerals and complexes likely to control Cd2+
activities in soils.
Equilibrium reaction

Log Ko

1. CdCO3 (s) (Otavite) + 2H+ = Cd2+ + CO2 + H2O

6.16

2. -Cd(OH)2 + 2H+ = Cd2+ + 2H2O

13.65

3. Cd3(PO4)2 + 4H+ = 3Cd2+ + 2H2PO4-

1.0

4. CdSO4.2Cd(OH)2 + 4H+ = 3Cd2+ + SO42- + 4H2O

22.65

5. FePO4.2H2O (Strengite) + 2H+ = Fe3+ + H2PO4- + H2O

-6.85

6. CaSO4.2H2O (Gypsum) = SO42- + Ca2+ + 2H2O

-4.64

7. CaCO3 (Calcite) + 2H+ = Ca2+ + CO2 + H2O

9.74

8. Ca3(PO4)2 (-Tricalcium Phosphate) + 4H+ = 3Ca2+ + 2H2PO4-

10.18

Source: Lindsay (1979)

88

Figure 4.17 Logarithm (base 10) of soluble Cd from the amended soil of Dareta and from
various pure compounds. Stability lines for minerals were calculated based on log K values
reported by Lindsay (1979).

89

4.12 Soil solution speciation


4.12.1 Inorganic complexes of Pb
The equilibrium reactions of Pb in solution are presented in Table 4.12. The distribution of
inorganic complexes of Pb is shown in Fig. 4.18. The soil solution is dominated by positively
charged and neutral complexes of SO42-, Cl-, and OH-. Across the pH range considered in this
study, free Pb (Pb2+) accounted for most the Pb in solution until around pH 7.6 after which
PbOH+ was the most dominant Pb specie in the soil solution. The activities of PbOH+ increased
with pH, similar to that of Pb(OH)2. Generally, the order of magnitude observed for the
distribution of inorganic complexes of Pb in the soil solution was: PbOH+ > PbSO4- > PbCl- >
Pb(OH)2. The general trend for these complexes was a downward curve after pH 6.6. In
comparison to PbOH+ and Pb(OH)2, PbSO4 - and PbCl- showed a similar trend although PbSO4
was more dominant. In total, the inorganic species of Pb accounted for just 1.3% (Table 4.13) of
the total Pb in solution while Pb2+ accounted for 70.24%. The remaining 28.46% represents Pborganic complexes. The percentage of these organic complexes reduced to 28.26% and 27.47%
as pH dropped to 6.6 and 7.1 respectively. At pH 7.1, PbOH+ accounted for 14.13% of the total
dissolved Pb in solution.
The high proportion of organic complexes corresponds with the high stability constants of
organic ligands relative to inorganic ligands, resulting in the predominance of organic complexes
under the soil conditions of this experiment.

90

Table 4.12 Equilibrium reactions for the inorganic complexes of Pb in the soil solution
Equilibrium reaction

Log Ko

Pb2+ + SO42- = PbSO4o

2.62

Pb2+ + Cl- = PbCl+

1.60

Pb2+ + H2O = PbOH+ + H+

-7.70

Pb2+ + 2H2O = Pb(OH)2 + 2H+

-17.75

Pb2+ + 3H2O = Pb(OH)3- + 3H+

-28.08

Source: Lindsay (1979)

91

Figure 4.18 The distribution of the inorganic complexes of Pb in the soil solution of Dareta,
Northern Nigeria

92

Table 4.13. The activities of dominant inorganic complexes of Pb in soil solution of Dareta,
Northern Nigeria expressed as percentages of total Pb (PbT) in soil solution

pH

Metal complex

Activity

5.4

PbT
Pb2+
PbSO4
PbOH+
PbCl+
Pb2OH3+
TOTAL

4.64E-05
3.26E-05
3.31E-07
1.60E-07
1.04E-07
1.17E-10

Pb T
Pb2+
PbOH+
PbSO4
PbCl+
Pb(OH)2
Pb2OH3+
TOTAL

7.44E-06
4.87E-06
3.96E-07
5.27E-08
1.46E-08
6.42E-10
4.32E-11

PbT
Pb2+
PbOH+
PbSO4
PbCl+
Pb(OH)2
Pb2OH3+
TOTAL

6.083E-06
3.501E-06
8.594E-07
3.55E-08
1.22E-08
4.21E-09
6.74E-11

6.6

7.1

93

70.24
0.71
0.34
0.25
0.00
71.54

65.5
5.32
0.71
0.2
0.01
0.00
71.74

57.55
14.13
0.58
0.2
0.07
0.00
72.53

Table 4.14 Ion association constant for the formation of aqueous inorganic and organic
complexes of Pb, Cd, and Zn in solution.
Zn

Pb

Cd
Kf

Ligand
Inorganic ligands
SO42-

2.34

2.75

2.46

Cl-

0.46

1.59

1.98

HPO42-

3.10

3.10

3.80

H2PO4-

1.60

1.50

3.20

OH-

8.96

7.70

10.10

Oxal3-

4.87

4.91

3.89

Citrate3-

6.10

5.70

5.00

EDTA4-

16.44

17.88

16.36

DTPA5-

18.29

18.67

19.00

Organic complexes

Source: Egwu, 2010.

94

4.12.2 Inorganic complexes of Cd


The equilibrium reactions of inorganic complexes of Cd are presented in Table 4.15.
The trend and differences in the distribution of the various inorganic complexes of Cd (Fig. 4.19)
are more striking compared to those of Pb. For all the inorganic species (including free Cd2+) the
trend appears relatively more uniform with a generally downward trend after pH 7.4. The order
of distribution of these complexes was: CdCl+ > CdSO4 > CdOH+ > CdCl2. The observed
dominance of CdCl+ is consistent with the high affinity of chloride for Cd as well as the higher
stability constant of CdCl complexes compared to other metal complexes of chloride such as
ZnCl and PbCl (Egwu and Agbenin, 2012). The complexation of Cd with hydroxide and chloride
ions may enhance the mobility of Cd in the environment (Hahne and Kroontje 1973).
The inorganic species of Cd accounted for just 0.89% of the total Cd in solution at pH 5.4 (Table
4.16), this increased to 0.90% and 0.99% as pH increased from 6.6 to 7.1 respectively.
Summation of calculated inorganic species plus free Cd concentration in solution added up to
71.12%, 70.38% and 69.01% at pH 5.4, 6.6, and 7.1 respectively. The organic Cd complexes
accounted for 28.88%, 29.62%, and 30.99% at the same pH. The formation of significant Cd
organic complexes in the solutions of several soils has also been reported in several other works
(Almas et al., 2000; Suave et al.. 2000; Krishnamurti and Naidu 2003) while the Cd organic
complexes in solution have been attributed to the formation of Cd-fulvate (Egwu and Agbenin
2012).

95

Table 4.15 Equilibrium reactions for the inorganic complexes of Cd in the soil solution
Equilibrium reaction

Log Ko

Cd2+ + Cl- = CdCl+

1.98

Cd2+ + 2Cl- = CdCl2o

2.60

Cd2+ + SO42- = CdSO4o

2.45

Cd2+ + H2O = CdOH+ + H+

-10.10

Cd2+ + 2H2O = Cd(OH)2o + 2H+

-20.30

Cd2+ + 3H2O = Cd(OH)3- + 3H+

-33.01

Cd2+ + 4H2O = Cd(OH)42- + 4H+

-47.29

Source: Lindsay (1979)

96

Figure 4.19 The distribution of inorganic complexes of Cd in the soil solution of Dareta,
Northern Nigeria

97

Table 4.16 The activities of dominant inorganic complexes of Cd in soil solution of Dareta, Northen
Nigeria expressed as percentages of total Cd (CdT) in soil solution
pH

5.4

6.6

7.1

Metal complex

Activity

CdT
Cd+2
CdCl+
CdSO4
CdOH+
TOTAL

1.24E-04
8.71E-05
6.61E-07
4.51E-07
1.82E-09

70.23
0.53
0.36
0.00

Cd T
Cd2+
CdCl+
CdSO4
CdOH+
CdCl2
TOTAL

3.51E-05
2.44E-05
1.74E-07
1.35E-07
8.45E-09
5.44E-11

69.48
0.50
0.38
0.02
0.00

CdT
Cd2+
CdCl+
CdSO4
CdOH+
CdCl2
TOTAL

3.30E-05
2.25E-05
1.87E-07
1.17E-07
2.35E-08
6.82E-11

68.02
0.57
0.35
0.07
0.00

98

4.12.3 Inorganic complexes of Zn


The equilibrium reactions of inorganic complexes of Zn are presented in Table 4.17. The
graphical relationship among the various inorganic complexes of Zn observed for the heavy
metal amended soils of Dareta is presented in Figure 4.20. The Zn complexes show a dual-trend
in distribution, in the sense that there was an upward trend for the hydrolysis species (ZnOH+
and Zn(OH)2) while the trend was downward for ZnCl- and ZnSO4- as the pH increased. The
order of distribution for these complexes was: ZnSO4 > ZnCl- > ZnOH- > Zn(OH)2. The
dominant species of a metal in solution is a strong indication of mobility and potential toxicity of
the metal. Inert complexes are least harmful while the free metal species are most toxic. The
ZnCl- implies high mobility of the metal especially in a net-negatively charged soil, while the
rate of adsorption will increase with increase in hydrolysis species. The inorganic complexes of
Zn together with the free Zn only accounted for 69.48% of total Zn dissolved in solution at pH
8.2 (Table 4.18). Inorganic complexes alone accounted for 1.4% of the total; hence the organic
complexes of Zn accounted for 30.52%. . It was reported by Egwu (2010) that dissolved organic
carbon (DOC) contributed more significantly to total Zn in several soils through the formation of
Zn-fulvate complexes. This possibility, together with the higher stability values of organic
ligands in comparison to inorganic ligands explains the high proportion of the Zn organic
complexes.

99

Table 4.17 Equilibrium reactions for the inorganic complexes of Zn in the soil solution
Equilibrium reaction

Log Ko

Zn2+ + H2O = ZnOH- + H+

-7.69

Zn2+ + 2H2O = Zn(OH)2o + 2H+

-16.80

Zn2+ + 3H2O = Zn(OH)3 + 3H+

-27.68

Zn2+ + 4H2O = Zn(OH)4 + 4H+

-38.29

Zn2+ Cl- = ZnCl+

0.43

Zn2+ + SO42- = ZnSO4

2.33

Source: Lindsay (1979)

100

Figure 4.20 The distribution of inorganic complexes of Zn in the soil solution at Dareta,
Northern Nigeria

101

Table 4.18 The activities of dominant inorganic complexes of Zn in soil solution of Dareta, Northern
Nigeria expressed as percentages of total Zn (ZnT) in solution
pH

Metal complex

Activity

5.4

ZnT
Zn2+
ZnSO4
ZnOH+
ZnCl+
Zn(SO4)22TOTAL

3.70E-04
2.61E-04
1.07E-06
7.18E-08
5.61E-08
1.51E-10

Zn T
Zn2+
ZnOH+
ZnSO4
ZnCl+
Zn(OH)2
Zn(SO4)22TOTAL

1.24E-04
8.69E-05
3.97E-07
3.86E-07
1.75E-08
1.90E-08
5.94E-11

ZnT
Zn2+
ZnOH+
ZnSO4
Zn(OH)2
ZnCl+
Zn(SO4)22Zn(OH)3TOTAL

1.14E-04
7.75E-05
1.07E-06
3.23E-07
1.55E-07
1.82E-08
4.68E-11
6.15E-12

6.6

7.1

102

70.4
0.29
0.02
0.02
0.00
70.73

69.93
0.32
0.31
0.01
0.02
0.00
70.59

68.08
0.94
0.28
0.14
0.02
0.02
0.00
69.48

4.13 Mechanism(s) controlling Pb, Cd, and Zn activities in the metal spiked soils of Dareta
A number of equilibrium studies have suggested the precipitation of pure minerals as controls of
metal activities in the soil solution, especially under conditions of high contamination. .
However, precipitation as a mechanism controlling mineral solubility in soils has been rejected
by other researchers (McBride and Blasiak, 1979; Cavallaro and McBride, 1980; Basta and
Tabatabai, 1992). This is because solubility diagrams of pure compounds show a uniform pH
effect on metal solubility, which is not the case under field condition. Also the solubility lines of
pure minerals usually have a slope of 2 (Martinez and Motto, 2000). The slopes of experimental
data are mostly far below this value, and for the soils of this study it was only Pb that gave slope
above unity (1.29) (Table 4.19). Slopes were 0.7 and 0.4 for Cd and Zn respectively. The
divergence of calculated slopes of these soils from 2 indicates that the slopes are lower than
those for pure compounds, and therefore only a slim likelihood of metal precipitation exist.
Despite their limitations, Santillan-Medrano and Jurinak (1975) as well as El-Falaky et al. (1991)
still argued for the use of solubility diagrams and the precipitation of minerals at high metal
concentrations and high pH values. However, Catlett et al. (2002) showed that extremely low
slopes could mean that the control for a metal is likely dominated by cation exchange reactions,
but Martinez and Motto (2000) argued that solubility studies cannot differentiate between ion
exchange, specific adsorption, and precipitation reaction. Although the soils in this study were
amended with very high concentrations of Pb, Cd, and Zn in order to simulate high level soil
contamination, the data generated from this experiment is not sufficient to establish the actual
mechanism controlling metal activities in the soil solution. The slopes obtained for the three
metals studies presumably rules out precipitation as mechanism controlling soil solution
activities of Pb, Cd, and Zn

103

Table 4.19 The slopes and coefficient of determination of -log Me2+ - pH relationship of
soils from the lead polluted farms of Dareta, Northern Nigeria

Me2+

Slope

r2

Pb2+
Cd2+
Zn2+

1.29
0.766
0.447

0.84
0.718
0.718

104

However this is not sufficient to entirely disregard the precipitation of mineral compounds
neither is it sufficient to differentiate between ion exchange, specific adsorption, and
precipitation. For a more definitive inference, detailed spectroscopic analysis is required to
confirm the formation of pure mineral compounds or otherwise. Also, there is need for additional
studies to compare, quantify the contribution of, and differentiate between cation exchange
reactions, specific adsorption, and precipitation with regards to the solubility of metals.
4.14. Changes in solubility of Pb, Cd, and Zn with incubation
After 21 days of incubation with distilled water, contrasting observations were recorded among
the various treatments. Although the incubated soils displayed a low concentration of extractable
Pb (Fig 4.21), Cd (Fig 4.22), and Zn (Fig 4.23) the trend varied with metal and extractant used.
Except for salt extractable Pb and Cd, the concentrations of all metals increased initially and then
decreased after approximately 7days for all measured metals. Concentrations of water extractable
Cd (Fig 4.22) had a similar trend with incubation time whereas those extracted by NH4OAc
increased gradually until 14 days of incubation, and then decreased again. However, the opposite
trend was observed for salt extractable Zn (Fig 4.23) where the concentration in extractant after
incubation periods of 14 and 21 days rose from 6.9 - 9.6mgkg-1.

105

Figure 4.21 Changes in solubility of Pb in Dareta soil after incubation with distilled water

106

Figure 4.22 Changes in solubility of Cd in Dareta soil after incubation with distilled water

107

Figure 4.23 Changes in solubility of Zn in Dareta soil after incubation with distilled water

108

The Concentrations of water extractable and salt extractable Cd were perfectly inversely related
over the incubation period (Fig 4.22). It was observed that the amount of Pb by NH4OAc
extraction increased gradually with increasing incubation time, reaching a maximum at about 7
days of incubation before declining sharply after 14 days (Fig. 4.21). A similar trend was
reported by Staunton and Wang (2003) for Cu, and they concluded that there is a tendency for
water soluble Cu content to increase and then decrease with time. Similarly, Dong and Ma
(2004) recorded decrease in metal mobility with incubation time for Pb and Cu. The exception in
this experiment was with water extractable Pb where its concentration increased gradually with
increasing incubation time. The initial increase in metal concentration with time could be due to
reductive dissolution of Fe and Mn. With reductive Fe dissolution, previously blocked
exchangeable sites are exposed (Dong and Ma, 2004). The reduction of Fe (III) to Fe (II) will
bring about the release of ferrous iron to the pore waters and also any metals that were adsorbed
to the ferric hydroxide surfaces. Jenne (1968) concluded that Fe and Mn oxides are the principal
soil surfaces that control the mobility of metals in soils and natural waters.
Dong and Ma (2004) further attributed the reduction in the soluble concentration of metals over
time to surface binding which probably ensues when soils are incubated for more time. Several
contradictory observations have been reported on changes in solubility and mobility of metals as
a result of incubation. This is in part due to the soil/water ratios used in various experiments
including metal concentrations determined in pore water, in filtrates separated from soil
suspension, and in leachate from a soil leached with pure water or an electrolyte (Dong and Ma,
2004). It was also suggested that CEC of a soil may change with incubation under water flooding
condition, and thus might greatly affect metal solubility and mobility (Dong and Ma, 2004).
Neither CEC nor Fe and Mn content at the different incubation times were measured in this
109

study, thus it cannot be conclusively stated that there was a change in CEC of the soil with
incubation nor the concentration of Fe and Mn increased with incubation, however the former is
most likely.
4.15. Reduction of Pb, Cd and Zn in the contaminated soils of Dareta under native
conditions and different levels of glucose amendments.
At the end of the 7-day incubation period, the measured redox potential (Eh) for the soils were
below 300, which is the borderline potential signaling the onset of reduction due to the depletion
of oxidants, such as O2 and NO3 (Ponnamperuma, 1972; Bohn et al., 2001). The lowest potential
for the samples was 98 mV with a corresponding pH value 5.37 (Table 4.20), whereas the
highest redox potential measured for the samples was 149 mV with a corresponding pH value of
5.19. Although a reduction reaction is a proton consuming process, only a slight correlation was
observed between pH and redox potential because observations were not uniform (Fig. 4.24);
however, all of the measured Eh values were within the range of Fe3+Fe2+ and O2H2O redox
couples (Bohn et al., 2001; McBride, 1994). The extent of metal solubility as induced by glucose
addition was measured from the changes in metal concentration in the soil solution. After the 7day incubation period, the concentrations of heavy metals in solution differed slightly from each
other and from the control. In most cases, the differences in metal concentrations were not
significant (Table 4.21). Although the addition of glucose was to provide energy required for
microbial catalysis/reduction of the soil by biotic mechanism, the absence of significant
differences in heavy metals solubilized between the two levels of amendments for most of the
samples suggests inability of the soil microbial population to recover after air-drying of the
samples and long-term storage. Karolien et al. (2001), Xiang et al. (2008) and Fierer and
Schiemel (2002) found that drying and rewetting impose significant stress on microbial
110

populations. Drying and rewetting cycles showed that the longer the time of drying, the more
difficult it was for the microbial population to re-establish; in most cases they do so only
sparingly. This observation could also imply decreased microbial activity due to the high levels
of the heavy metals in the samples. It is now widely accepted that the presence of abovethreshold concentration of metals in the environment imposes an adverse effect on microbial
populations. Under such conditions, only organisms that have developed special adaptive
mechanism to occlude or sequester these contaminants have been found to exist.

111

Table 4.20 pH and Redox potential of soil samples after 7 days incubation under two
glucose rates
Dareta North

Dareta South

Dareta East

Dareta West

pH
1%

5.37

4.74

5.04

5.19

3%

4.65

4.45

4.79

4.94

Redox Potential (mV)


1%

98

127

107

149

3%

136

148

126

119

112

Table 4.21 Mean concentrations (mg kg-1) of solubilized Pb, Cd, and Zn under different
levels of glucose amendment
Dareta North Dareta South

Dareta East

Dareta West

Pb
1%

0.09b

0.22a

0.20a

0.30a

3%

2.36a

0.19a

0.19a

0.45a

CONTROL

0.08b

0.18a

0.18a

0.32a

1%

0.03b

0.68a

0.13a

0.10a

3%

0.76a

0.61a

0.17a

0.09a

CONTROL

0.01b

0.58a

0.14a

0.12a

1%

0.24b

1.76a

0.29b

0.27a

3%

2.40a

0.98b

1.17a

0.27a

CONTROL

0.32b

1.11b

0.25b

0.24a

Cd

Zn

Means followed by the same letter are statistically the same at 5% probability

113

Figure 4.24 Eh-pH relationships after 7 days of incubation at 1% and 3% glucose levels

114

CHAPTER FIVE
5.0 SUMMARY AND CONCLUSIONS
This research work showed that the total heavy metal concentration as determined by mixed acid
digestion were extreme for Pb (2246.55 mg kg -1) and Cd (70 g kg-1) but the concentration of
Zn was within acceptable limit. Down the profile, the concentration of metals were observed to
decrease relative to the top horizon; however the metals all showed a double peak which could
be due to facilitated transport by colloids, as suggested by the texture change down the profile.
No uniform trend was observed in metal concentration with distance although higher
concentrations of Pb, Cd, and Zn were found within 150m of sampling. Out of the total
concentration of the metals only 0.5% Pb, 3.3% Cd, and 0.3% Zn were found exchangeable by
Ca as determined by 0.01M CaCl2 extraction. Fractionation results partitioned the bulk of the
metals onto the residual fraction. As high as 89% of Pb, 62 % of Cd and 37% of Zn were held
within the residual fraction. Based on mean percentages, the following orders of occurrence were
observed for Pb RES > OM > OX > CARB > EXCH, while Cd gave the following order; RES >
EXCH > OX > OM > CARB; and the order for Zn was: RES > OX > OM > EXCH > CARB.

Calculated enrichment factor (EF) and I-geo factor showed strong superficial enrichment for the
metals, especially Pb. Similar observation was made for Cd while Zn only indicated some level
of enrichment when compared to average crustal concentration.

Speciation calculations showed that the soil solution was dominated by free metal species of Pb,
Cd, and Zn. Complexes of SO42-, Cl- and OH- were the dominant inorganic complexes in
solution. The order of magnitude for distribution of inorganic complexes of Pb was: PbOH+ >
115

PbSO4- > PbCl> Pb(OH)2. The order of distribution of the inorganic complexes of Cd was:
CdCl+ > CdSO4> CdOH+ > CdCl2. The observed dominance of CdCl+ is consistent with the high
affinity of chloride for Cd. The order of distribution of the Zn complexes was: ZnSO4 > ZnCl- >
ZnOH- > Zn(OH)2.

In the end the following conclusions were drawn from this experiment:
1. Results from this research showed that the soils of the farmlands are highly contaminated
with Pb and Cd, but not Zn. However, the bulk of the metals were partitioned to the
residual fraction. This implies a low risk of metal transfer from the soil to the growing
crops, under normal agricultural practices.
2. The high level of enrichment of the metals as calculated from the enrichment factor and
I-geo factor only indicate the buffering of soil solution metal concentration through the
slow processes involved in soil development. The calculated mobility indices for the
three elements were also too low to warrant any major concern of pollution to growing
crops. Observations from this study reflected that, except for geological time, the total
metal concentrations recorded for the soils of the farmlands would not play a significant
role with respect to plant growth or in terms of most environmental processes
3. The predominance of free species of Pb, Cd, and Zn in the soil solution and the high
proportion of organic metal complexes observed under high level of pollution would
enhance plant uptake and the mobility of the metal contaminants in the environment.
4. Mechanism(s) other than the precipitation of pure mineral compounds would, most
likely, control the activities of Pb, Cd, and Zn in the soil solution over long term
accumulation.

116

5. The absence of an apparent equilibrium between the metals under study and any of the
pure solid phase(s) considered as likely controls of metal activity in solution suggests that
the progression of the chemical reactions may be limited by kinetic processes; this is
fairly indicated by the value of the Gibbs free energy (Gr) of the solid phases
considered. The spontaneity of a reaction is directly proportional to the magnitude of the
Gibbs free energy (Langmuir 1997).

Although the bulk partitioning of the total heavy metal concentration to the soil residual fraction
does not pose immediate threat to soil fertility, neither does it represent potential damage to food
quality, for soil health management of the farmlands around the goldmines it is recommended
that farmers around the contaminated goldmine exercise caution in the use of irrigation water and
animal dung so as to avoid soil poisoning through anthropogenic accumulation of heavy
metals in the farmlands, which may consequently decrease soil fertility and damage food quality.
From the observation of a bulk partitioning of Pb, Cd, and Zn onto the residual fraction, in order
to establish more definitive inference on the risk of metal transfer to the food chain, further
research is needed on the effect of the H+ produced in the rhizosphere on the solubilization of
metals bound to the residual fractions of the soil. This is of concern in soils such as those of this
study are with high metals concentrations held on the residual soil fraction.

Finally, it is important to state that even though numerous reports have been published on the
gold mining activities in Zamfara state and the number of human casualties suffered as well as
the extreme levels of Pb in the households where the ores are been prepared, farming activities in
Dareta Village of Anka LGA of Zamfara state is not at risk. The available fraction of Pb in the
117

farmland soils is below the threshold that may adversely impact on crop growth and contaminate
the food chain.

118

REFERENCES
Abdu, N. (2010). Availability, transfer and balances of heavy metals in urban agriculture of
WestAfrica. Ph D. Dissertation, Faculty of Organic Agricultural Sciences, University of
Kassel, Germany.

Abdu, N., Abdulkadir, A., Agbenin, J.O., and Buerkert, A. (2011a). Vertical distribution of
heavy metals in wastewater-irrigated vegetable garden soils of three West African cities.
Nutrient Cycling Agroecosystem. 89: 387397 doi: 10.1007/s10705-010-9403-3.

Abdu, N., Agbenin, J.O., and Buerkert, A. (2011b). Phytoavailability, Human Risk
Assessmentand Transfer Characteristics of Cadmium and Zinc Contamination from Urban
Gardens in Kano, Nigeria. Journal of Science Food and Agriculture. 15: 2722-2730.

Abdu, N., Agbenin, J.O., and Buerkert, A. (2011c). Geochemical assessment, distribution,
anddynamics of trace elements in urban agricultural soils under long-term wastewater
irrigation in Kano, northern Nigeria. Journal of Plant Nutrition and Soil Science. 174:
447-45.

Abdu, N., and Yusuf, A.A. (2012). Heavy metal contamination of soil and crop,and humanhealth
impacts in a contaminated village around gold mine in Zamfara state. Proceedings of the
46th Annual Conference of the Agricultural Society of Nigeria, Kano 2012. Pp 961-965

Abdu, N., Agbenin, J.O., and Buerkert, A. (2012). Fractionation and Mobility of Cadmium and
Zinc in Urban Vegetable Gardens of Kano, Northern Nigeria. Environmental Monitoring
Assessment. 184: 20572066. doi: 10.1007/s10661-011-2099-2

Adriano, D.C. (2001). Trace Metals in Terrestrial Environment:


Bioavailability, and Risk of Metals (2nd ed.). NY: Springer-Verlag..

Biogeochemistry,

Adriano, D.C., Bolan N.S., Vangronsveld J., and Wenzel W. W. (2005) Heavy metals.In: Hillel,
D. (ed).Encyclopedia of Soils in the Environment, pp 175-182.Amsterdam: Elsevier.

Agbenin, J.O. (1995). Laboratory manual for soil and plant analysis: Selected methods and data
analysis. 140pp. Zaria: Department of Soil Science, Ahmadu Bello University.
119

Agbenin, J.O. (2002). Lead in Nigerian savanna soils under long-term cultivation. Science ofThe
Total Environment. 286: 1-14.

Agbenin,J.O. and Olojo, A.A. (2004). Competitive adsorption of copper and zinc by a Bthorizon
of a savanna Alfisol as affected by pH and selective removal of hydrous oxide and organic
matter. Geoderma. 119: 85-95.

Almas, A.R., McBride, M.B. and Singh, B.R. (2000). Solubility and lability of cadmium and
zinc in two soils treated with organic matter. Soil Science. 165: 2509.

Alloway, B.J., (1995). Heavy Metals in Soils. pp. 11-17. London: Blackie and Academic
Professional.

Appel, C. and Ma, L. (2002). Concentration, pH, and surface charge effects on cadmium and
lead sorption in three tropical soils. Journal of Environmental Quality. 31: 581589.

Atanassov, I., Vassileva, V. andShegunova, P. (1999). Applications of data for


backgroundconcentrations of Pb, Zn, Cu and Cd in soils for calculating critical loads, pp.
137-140. In U: UBA. Effects-Based Approaches for Heavy Metals. Workshop Schwerin,
12-15, October. Germany,

Atsuyuki, O., Noboru I., Shigeru T. and Yoshiko T. (2007). Preliminary study for speciation
geochemical mapping using a sequential extraction method. Bulletin. Geological. Survey.
Japan. 58 (7/8):201-237

Backes, C.A., McLaren, R.G., Rate, A.W. and Swift, R.S. (1995). Kinetics of cadium and cobalt
desorption from iron and manganese oxides. Soil Science Society of America Journal. 59:
778-785.

Banat, K.M., Howari F.M., and Al-Hamad A.A. (2003) Heavy metals in urban soils of central
Jordan: should we worry about their environmental risks? Advances in Environmental
Research. 8:113120

120

Banat, K.M., Howarib, F.M. and Al-Hamad, A. A. (2005). Heavy metals in urban soils of
central Jordan: Should we worry about their environmental risks? Environmental
Research. 97: 258273

Bartlett, R.J. and James, B.R. (1993). Redox chemistry of soils. In: Sparks, D.L. (Ed.). Advances
in Agronomy. 50: 151-208.

Basta, N.T., Pantone, D.J., and Tabatabai, M.A. (1992). Path analysis of heavy metal adsorption
by soil. Agronomy Journal. 85: 1054-1057.

Basta, N.T., Ryan J.A., and Chaney R.L. (2005). Trace element chemistry in residual-treated
soil: key concepts and metal bioavailability. Environmental Quality. 34: 49-63

Baumann, T., Muller, S. and Niessner, R. (2002). Migration of dissolved heavy metal
compounds and PCP in the presence of colloids through a heterogeneous calcareous
gravel and a homogeneous quartz sand - pilot scale experiments. Water Resources. 36:
1213-1223.

Biddappa, C.C., Chino M. andKumazawa, K. (1981). Adsorption, desorption, potential and


selective distribution of heavy metals in selected soils of Japan. Journal of Environmental
Science and Health, Part B. 156: 511-528.

Bittel, J. R. and Miller, R. J. (1974). Lead, Cadmium, and Calcium Selectivity Coefficients on
Montmorillonite, Illite, and Kaolinite. Journal of Environmental Quality 3:250-253.

Blacksmith Institute (2010). 2010-2011 final report: UNICEF Programme Cooperation


Agreement. Environmental Remediation- Lead Poisoning in Zamfara.

Bohn, H.L., McNeal, B.L., and OConner, G.A. (2001). Soil Chemistry (3rd ed). NY: .John
Wiley and Sons.
Bolt, G.H. and Bruggenwert M.G.M. (Eds.). (1978). Soil Chemistry. Amsterdam: Elsevier
scientific publishing company

Bowen, H.J.M. (1979). Environmental Chemistry of the Elements. London : Academic Press.
121

Brmmer, G.W., Gerth, J., Herms, U. and Pflanzenernhr, Z. (1986).Heavy metal species
mobility, and availability in soils.Bodenk. 149, 382398.

Brumelis, G., Brown, D.H., Nikodemus, O. and Tjarve, D. (1999). The monitoring and risk
assessment of Zn deposition around a metal smelter in Latvia. Environmental Monitoring
and Assessment.58(2): 201-212

Cavallaro, N. and McBride, M.B. (1980). Activities of Cu+2 and Cd+2 in soil solutions as affected
by pH.Soil Science Society of America Journal 44: 729-732.

Catlett, K.M., Heil,D.M. and Lindsay, W.L. (2002). Soil chemical properties controlling Zn
activities in 18 Colorado soils. Soil Science Society of America Journal. 66: 1182-1189.

Chao, T.T., (1972). Selective dissolution of manganese oxides from soils and sediments with
acidified hydroxylamine hydrochlorid. Soil Science Society of America Proceedings. 36:
764768

Chirenje, T., Ma, L.Q., Clark, C. and Reeves, M. (2003). Cu, Cr and As distribution in soils
adjacent to pressure-treated decks, fences and poles. Journal of Environmental Pollution.
124: 407-417.

Chopin, E.I.B. and Alloway, B.J. (2007a). Distribution and Mobility of Trace Elements in Soils
and Vegetation around the Minning and Smelting Areas of Tharsis, Riotinto and Huelva,
Iberian Pyrite Belt, SW Spain. Water, Air, and Soil Pollution. 182: 245-261.

Chopin, E.I.B. and Alloway, B.J., (2007b). Trace Element Partitioning and Soil Particle
Characterization around the Minning and Smelting Areas of Tharsis, Riotinto and Huelva,
Iberian Pyrite Belt, SW Spain. Science of the Total Environment. 373: 488-500.

Chuan, M.C., Shu, G.Y. and Liu, J. C. (1996). Solubility of heavy metals in contaminated soils:
Effects of redox potential and pH. Water Air and Soil Pollution. 3:159-170.

122

Chukwuma, S.E., Richard, T.I., Ephraim, N.A., Obinna,A.O. and Ikechukwu, N.E.O. (2012).
Biotechnological Tools for Environmental Sustainability: Prospects and Challenges for
Environments in Nigeria- A Standard Review. Biotechnology Research International.
doi:10.1155/2012/450802

Citeau, L., Lamy, I., van Oort, F. and Elsass, F. (2003). Colloidal facilitated transfer of metals in
soils under different land use. Colloids and Surfaces. A: Physicochemical Engineering
Aspects 217: 11-19.

Danko, M.M. (2006). Speciation and bioavailability of selected metals heavy metals in
soilsunder long-term waste water irrigation water. M Sc. Thesis, Ahmadu Bello
University, Zaria.

Das, P., Samantaray, S. and Rout G.R. (1997). Studies on cadmium toxicity in plants: A review.
Environmental Pollution. 98: 29-36.

Davis, J.A. and Leckie, J.O. (1978). Surface ionization and complexation at theoxide/water
interface II. Surface properties of amorphous iron oxyhydroxide and adsorption of metal
ions. Journal of Colloid and Interface Science. 67(1): 90107

deHaan, F.A.M., Zwerman P.J. (1978). In: Bolt, GH, Bruggenwert, MGM (Eds). Soil Chemistry.
Amsterdam: Elsevier Scientific publishing company.

Dong, Y. and Ma, L.Q. (2004) Effects of incubation on solubility and mobility of trace metals in
two contaminated soils, Environmental Pollution130, 301-307

Egwu, G.N. (2010). Bioavailability and solubility equilibria of cadmium lead and zinc in urban
Garden soils. Ph D dissertation, Department of Soil Science, Ahmadu Bello University,
Zaria, Nigeria.

Egwu,G.N. and Agbenin J.O. (2012). Adsorption and solidsolution compositional relationships
of Cadmium in tropical savannah soils from Northern Nigeria. Toxicology and
Environmental Chemistry. 1: 1-11.

123

Eggleton J. and Thomas K.V. (2004). A review of factors affecting the release andbioavailability
of contaminants during sediment disturbance events. Environment International. 30: 973
980

El-Falaky, A.A., Aboulroos, S.A. and Lindsay, W.L. (1991). Measurement of cadmium activities
in slightly acidic to alkaline soils. Soil Science Society of America Journal. 55: 974-979.

Elvingson, P. and Agren C. (2004). Air and the environment. Goteborg: Swedish NGO
Secretariat on Acid Rain. 174p.

Essington, M.E. (2004). Soil and Water Chemistry. An Integrative Approach. Boca Raton: CRC
Press.

Evangelou, V.P. (1998). Environmental Soil and Water Chemistry. Principles and Application.
NY: John Wiley and Sons, Inc.

Fierer, N. and Schimel, J.P. (2002). Effects of dryingrewetting frequency on soil carbon and
nitrogen transformations. Soil Biology and Biochemistry. 34: 777787.

Filguerias, A.V., LavillaI. and Bendicho C. (2002). Chemical sequential extraction for metal
partitioning in environmental solid samples Journal of Environmental Monitoring. 4:
823857

Finck, A. (1982). Fertilizers and Fertilization. Verl. Chemie, Deerfield Beach, Florida. 1992: 2nd
Edit. Dnger und Dngung. VCH, Weinheim New York.

Galadima, A.I. and Garba, Z.N. (2012). Heavy metals pollution in Nigeria: causes and
consequences. Elixir Pollution. 45: 7917-7922

Garcia-Miragaya, J. (1984). Levels, chemical fractionation, and solubility of lead in roadside


soils of Araca, Venezuela. Soil Science 138: 147-152.

124

Gee, G.W. and Or, D. (2002). Particle-size analysis. In: Dane, J.H., Topp, G.C. (Eds.), Methods
of Soil Analysis, Part 4. SSSA Book Series No. 5. Madison: Soil Science Society of
America. WI, pp. 255293.

Gerringa, L.J.A., de Baar, H.J.W., Nolting, R.F. andPaucot, H. (2001). The influence of salinity
on the solubility of Zn and Cd sulphides in the Scheldt estuary. Journal of Sea Research.
46, 201-211.

Gimmler, H., Carandang, J., Boots, A., Reisberg, E. and Woitke, M. (2002). Heavy metal content
and distribution within a woody plant during and after seven years continuous growth on
municipal solid waste (MSW) bottom slag rich in heavy metals. Applied Botany. 76: 203217.

Gothberg, A., Greger, M., Holm, K. and Bengtsson, B. (2004). Influence of nutrients on uptake
and effects of Hg, Cd and Pb in Ipo-moeaaquatica. Journal of Environmental Quality. 33:
1247-1255.

Gschwend, P. M. and Reynolds, M. D. (1987). Monodisperse ferrous phosphate colloids in an


anoxic groundwater plume. Journal of Contaminant Hydrology. 1:309-327.

Gustafsson, J.P. (2005). Visual MINTEQ (Version 2.32); Department of Landand Water
Resources Engineering. The Royal Institute of Technology, Stockholm, Sweden.
Available from: <http://www.lwr.kth.se/english/OurSoftWare/Vminteq/>. Accessed on 28
February, 2013.

Hahne, H.C. and Kroontje, W. (1973). Significance of pH and chloride concentration onbehavior
of heavy metal pollutants: Mercury, Cadmium, Zinc and Lead. Jouranal of.
Environmental Quality. 2: 444-450.

Hezelton, P. and Murphy, B. (2007). Interpreting soil test results, what do all the numbers
mean? Victoria, Australia:CSIRO publishing.

Hlavay, J., Prohaska, T., Weisz, M., Wenzel, W.W., and Stingeder, G.J. (2004). Determination
of trace elements bound to soils and sediment fractions. IUPAC technical report; Pure
and Applied Chemistry. 76 (2): 415442.
125

Hogan, C.M. and Monosson, E. (2013). Heavy metal". In: Cutler J. (ed). Encyclopedia of Earth.
Washington, D.C: Environmental Information Coalition, National Council for Science and
the Environment.

James, R.O. and Healy, T.W.(1972). Adsorption of hydrolyzable metal ions at the oxide-water
interface: III. Thermodynamic model of adsorption. Journal of Colloid Interface Science.
40:65-81.

Jenne, E.A. (1968). Control of Mn, Fe, Co, Ni, Cu, and Zn concentrations in soils and water-the
dominant role of hydrous manganese and iron oxides. Advances in Chemistry. 7: 337-387.

Jones, M.J. (1973). The organic matter content of the savanna soils of West Africa. Journal of.
Soil Science. 24(1): 42-53.

Kabata-Pendias,A. and Pendias, H. (2001). Trace Elements in Soils and Plants. Boca Raton, FL:
CRC Press.

Kabata-Pendias, A., and Pendias, H. (1992). Trace Elements in Soils and Plants (2nd edn). Boca
Raton, FL: CRC Press,.

Kadar, I. (1995). Effect of heavy metal load on soil and crop. ActaAgronomicaHungarica. 43:
3D9.

Karathanasis, A.D., Johnson, D.M.C. and Matocha, C.J. (2005). Biosolid Colloid- Mediated
Transport of Copper, Zinc, and Lead in Waste-Amended Soils. Journal of Environmental
Quality. 34: 1153-1164

Karolien, D, Johan S., Heleen B., Serita D.F., Edward, T.E., Roel, M. and Keith, P. (2001).
Influence of dry-wet cycles on interrelationship between aggregate, particulate organic
matter, and microbial community dynamics. Soil Biology and Biochemistry. 33: 15991611

126

Kashem, M.A., Singh, B.R. and Kawai, S. (2007). Mobility and distribution of cadmium, nickel
and zinc in contaminated soil profiles from Bangladesh. Nutrient Cycling in
Agroecosystems 77: 187-198.

Kebbekus, B.B. and Mitra, S. (1998). Environmental Chemical Analysis. Glasgow:.Blackie


Academic and Professional,

Kennedy, H.V., Sanchez, A.L., Oughton, D.H. and Rowland, A.P. (1997). Use of Single and
Sequential Chemical Extractants to Assess Radionuclide and Heavy Metal Availability
From Soils for Root Uptake. Analyst. 122: 89R-100R

Kirkham, M.B. (2006). Cadmium in plants on polluted soils: Effects of soil factors,
hyperaccumulation, and amendments. Geoderma. 137, 19-32.

Krishnamurti, G.S.R. and Naidu. R. (2003). Soil-solution equilibria of cadmium in soils.


Geoderma113: 1730.

Krishnamurti, G.S.R., Huang, P.M., Van Rees, K.C.J., Kozak, L.M., and Rostad, H.P.W. (1995).
Speciation of particulate- bound cadmium in soils and its bioavailability. Analyst 120:
659 665.

Kuo, S. (1986). Concurrent sorption of phosphate and zinc, cadmium, or calcium by a hydrous
ferric oxide. Soil Science Society of America Journal. 50: 1412-1419.

Lagerweff, J. V. and Bower, D. L. (1972). Exchange adsorption or precipitation of lead in soil


treated with chlorides of aluminum, calcium, and sodium. Soil Science Society of America
Proceedings. 37: 11-13.

Langmuir, D. (1997). Aqueous Environmental Geochemistry. New Jersey (US): Pretince-Hall.


Li, Z. and Shuman, L.M. (1996). Heavy Metals movement in metal contaminated soil profiles.
Soil Science. 161: 656 666.

Lim, C.H., and Jackson, M.L. (1982). Expandable Phyllosilicates Reactions with Lithium on
Heating. Clays and Clay Minerals. 34: 346352.
127

Lindsay, W.L. (1979). Chemical Equilibria in Soils. NY: John Wiley and Sons.

Liu, J., Duan C-Q, Zhu, Y., Zhang, X., and Wang, C. (2007).Effect of chemical fertilizers on the
fractionation of Cu, Cr and Ni in contaminated soil. Environmental Geology.52(8): 16011606

Lo, H.M., Lin, K.C., Liu, M.H., Pai, T.Z., Lin, C.Y., Liu, W.F., Fange, G.C., Luf, C., Chiang,
C.F., Wanga, S.C. et al. (2009). Solubility of heavy metals added to MSW. Journal of
Hazardous Materials. 161: 294299.

Ma, Q.L. and Rao, G.N. (1997). Chemical fractionation of cadmium, copper, nickel and zinc in
contaminated soils. Journal of Environmental. Quality 26: 259-264.

Mamboya, F.A. (2007). Heavy metal contamination and toxicity: studies of macroalgae from the
Tanzanian coast.Ph.D Dissertation. Stockholm University. ISBN 91-7155-374-6.

Manceau, A., Lanson, B., Schlegel, M.L., Harge, J.C., Musso, M., Eybert-Berard, L. and
Lamble, G.M. (2000). Quantitative Zn speciation in smelter-contaminated soils by
EXAFS spectroscopy. American Journal of Science. 300:289343.

Martinez, C. E. and Motto, H. L. (2000). Solubility of lead, zinc and copper added to mineral
soils. Environmental Pollution. 107: 153-158.

McBride, M. B. (1994). Environmental Chemistry of Soils. NY: Oxford University Press.


McBride, M. B. (1980). Chemisorption of Cd2+ on calcite surfaces. Soil ScienceSociety of
America Journal. 44: 26-28

McBride, M.B., and Blasiak, J.J. (1979). Zinc and copper solubility as a function of pH in an
acid soil. Soil Science Society of America Journal. 43: 866-870.

McBride, M.B. (1989). Reactions Controlling Heavy Metal Solubility in Soils. Advances in Soil
Science. 10: 1-56.
128

McLean, J.E. and Bledsoe, B. E. (1992). Behavior of materials in soils. Ground Water
Issue.EPA/540/S-92/018October 1992. Accesses February, 2012.

Mench, M.J., Didier, V.L., Loffler, M., Gomez, A. and Masson, P. (1994). A Mimicked in-situ
remediation study of metal-contaminated soils with emphasis on cadmium and lead.
Journal of Environmental Quality. 23: 58-63.

Mellbye, M. and Hart, J. (2003). Micronutrients for willamette valley grass seed production.
Crop and soil news/notes 17 (2). Corvallis, OR: Oregon state university extension service.

Merrill, J.C., Morton, J.J.P. and Soileau, S.D. (2007). Metals. In:.Hayes, A.W.. Principles and
Methods of Toxicology (5th ed.). CRC Press.

Mo, L.S., Moo, C.C. and Ho, Y.S. (1999). Speciation of Cd, Cu and Zn, in sewage sludge treated
soils incubated under aerobic and unaerobic conditions. Agricultural Chemistry and.
Biotechnology. 42: 85-91.

Morris, C. (1992) (ed). Academic Press Dictionary of Science and Technology. San Diego,
CA:Academic Press.

Murphy, J. and Riley, J.P. (1962). A Modified Single Solution for the Determination of
Phosphate in Natural Waters. Analalytica Chemistry Acta. 27: 31-36.

Muller, G. (1969). Index of geo-accumulation in sediments of the Rhine River.Geojournal2:


108118.

Nabulo, G. (2006). Assessment of heavy metal contamination of food crops and vegetables
grown in and around Kampala city, Uganda. Ph.D. Dissertation, Makerere University.

Navarro, M.C., Prez-Sirvent, C., Martnez-Snchez, M.J., Vidal, J., Tovar, P.J.and Bech, J.
(2008). Abandoned mine sites as a source of contamination by heavy metals: A case study
in a semi-arid zone. Journal of Geochemical Exploration. 96(2-3): 183-193.

129

Nelson, D.W. and Sommers, L.M. (1982). Total Carbon, Organic Carbon and Organic Matter. In
Sparks, D.L. (Ed.).Methods of soil Analysis Part 2. Chemical Methods. pp. 9611010.
Madison: American Society of Agronomy.

Nemeth, T. and Kadar, I. (2005). Leaching of Microelement Contaminants: a long-term Field


Study. ZeitschriftfrNaturforschung. 60c: 260-264.

Norrstrom, A.C. and Jacks, G. (1998): Concentration and fractionation of heavy metals in
roadside soils receiving de-icing salts. Science of the Total Environment. 218: 161-174.

NPC. (2006). National Population Census. National Population Commission.

Ntekim, E.E., Ekwere, S.J. and Ukpong, E.E. (1993). Heavy metal distribution in sediments from
Calabar River, Southeastern Nigeria. Environmental Geology. 21: 237241

Oliver, M.A. (1997). Soil and Human Health: A Review. European Journal of Soil Science. 48:
573-592.

Ortiz, O., Alcaniz, J.M. (2006). Bioaccumulation of Heavy Metals in Dactylisglomerata L.


Growing in a Calcareous Soil Amended with Sewage Sludge. Bioresearch Technology.
97: 545-552.

Parkhurst, V. and Appelo, C.A.J. (1999). Users Guide to PHREEQC (Version 2). A Computer
Program for Speciation, Batch-Reaction, One-Dimensional Transport, and Inverse
Geochemical Calculations. Water-Resources Investigations Report 99-4259. U.S.
Department of the Interior U.S. Geological Survey
Peterson, H.G., Healey, F.P. and Wagemann, R. (1984). Metal toxicity to algae: a highly pH
dependent phenomenon. Canadian Journal of Fishery and Aquatic Science. 41: 974-979.

Petrovic, M., Kastelan-Macan, M. and Horvat A.J.M. (1999). Interactive sorption of metal ions
and humic acids onto mineral particles. Water, Air and Soil Pollution. 111: 41-56.

130

Ponnamperuma, F.N. (1972). The chemistry of submerged soils. Advances in Agronomy 24: 29 96.
Rauret, G. (1998). Extraction procedures for the determination of heavy metals in contaminated
soil and sediment.Talanta. 46: 449-455

Rickard, D.T. and Nriagu, J.E. (1978). Aqueous Environmental Chemistry of Lead. In The
Biogeochemistry of Lead in the Environment. Part A. Ecological Cycles. J. O. Nriagu
(ed.), pp. 291-284. North Holland, NY: Elsevier.

Rogan, N., Dolenec, T., Serafimovski, T., Tasev, G. and Dolenec, M. (2010). Distribution and
mobility of heavy metals in paddy soils of the Koani field in Macedonia. Environmental
Earth Science. 61: 899907.

Rhoades, J.D. (1982). Cation Exchange Capacity. In: A.L. Page, R. H. Miller and Keeney, D. R.
(Eds.). Methods of Soil Analysis. Part 2. Chemical and Microbiological Properties. pp.
167-179. Madison WI: Soil Science Society of America.

Roberts, D.R., Nachtegaal, M. and Sparks D.L. (2005). Speciation of metals in soils. In:
Tabatabai M.A., Sparks D.L. (eds) Chemistry of Soil Processes. Madison, WI: Soil
Science Society of America.

Ross, S. (1994). Toxic metals in Soil-Plant Systems. Chichester, UK: John Wiley and Sons.

SAS (2002). Statistical Analysis System (SAS) Version 9.0. Users Guide Inst. Cary, N. C.

Santillan-Medrano, J. and Jurinak, J.J. (1975). The chemistry of lead and cadmium in soil: solid
phase formation. Soil Science Society of America Proceedings.39: 851-856.

Schlichting, E., Blume, H.P. and Stahr, K. (1995). BodenkundlichesPraktikum (2nd ed.). Berlin:
Blackwell.

131

Singer, M.J. and Hanson, L. (1969). Lead accumulations in soils near highways in the
TwinCities metropolitan area. Soil Science Society of America Proceedings. 33: 152-153

Sparks D.L. (2003). Environmental Soil Chemistry (2nd ed.). San Diego: Academic Press.

Sparks, D.L. (2005). Toxic metals in the environment: the role of surfaces. Elements 1: 193-197.

Sposito, G. (2008).The Chemistry of Soils (2nd ed.). NY: Oxford University Press.

Sposito, G. and Coves, J. (1988). SOILCHEM: A Computer Program for the Calculation of
ChemicalEquilibria in Soil Solutions and Other Natural Water Systems. University of
California, Riverside: Kearney Foundation of Soil Science.

Sposito, G. and Mattigod, S.V. (1980). GEOCHEM: A Computer Program for the Calculation of
Chemical Equilibria in Soil Solutions and Other Natural Water Systems. University of
California, Riverside: Kearney Foundation of Soil Science.

Stone, M. and Droppo, I.G. (1996). Distribution of lead, copper and zinc in size-fractionated
river bed sediment in two agricultural catchments of southern Ontario, Canada.
Environmental Pollution 93: 353-362.

Staunton, S. and Wang, G. (2003). Evolution of redox properties and trace metal solubility in soil
during incubation. Geophysical Research Abstracts. 7: 72-83.

Suave, S., Hendershot, W. and Allen, H.E. (2000). Solidsolution partitioning of metals in
contaminated soils: Dependence on pH, total metal burden, and organic matter.
Environmental Science and Technology. 34: 112531.

Tessier, A. Campbell, P.G.C. and Bissom, M. (1979) Sequential extraction procedure for the
speciation of particulate trace metals. Analytical Chemistry. 51:844850.

132

Tiller, K.G., Gerth, J. and Brummer, G. (1984). The relative affinities of Cd, Ni, and Zn for
different soil clay fractions and goethite. Geoderma. 34:17-35.

Tack, F.M., Callewaert, O.W.J.J. and Verloo, M.G. (1996). Metal Solubility as a Function of pH
in a Contaminated, Dredged Sediment Affected by Oxidation. Environmental Pollution.91
(2): 199-208.

Tyler, G. and Olsson, T. (2001). Concentrations of 60 elements in the soil solution as related to
the soil acidity. European Journal of Soil Science. 52: 151-165.

Udom, B.E., Mbagwu, J.S.C., Adesodun, J.K., and Agbim, N.N.(2004). Distribution of zinc,
copper, cadmium and lead in a tropical Ultisol after long-term disposal of sewage sludge.
Environmental International. 30: 467 470.

USEPA (2005). Guideline for carcinogen risk assessment: Risk assessment forum. Washington,
DC: U.S. Environmental Protection Agency.

Ure, A.M., Quevauviller, P., Muntau, H., and Griepink, B. (1993). Speciation of Heavy Metal in
Soils and Sediments. An account of the Improvement and Harmonixation of Extraction
Techniques Undertaken under the Auspices of the BCR of the Commission of the
European Communities. International Journal of Environmental and Analytical
Chemistry. 51: 135-151.

Vaalgamaa, S. andConley, D.J. (2008). Detecting environmental change in estuaries: Nutrient


and heavy metal distributions in sediment cores in estuaries from the Gulf of Finland,
Baltic Sea. Estuarine, Coastal and Shelf Science. 76(1): 45-56.

Voegelin, A., Barmettler, K., and Kretzschmar, R. (2003): Heavy metal release from
contaminated soils: comparison of column leaching and batch extraction results. Journal
of Environmental Quality. 32(3):865-75.

Wang, X., Shan, X., Zhang, S. and Wen, B. (2004). A Model for Evaluation of the
Phytoavailability of Trace Elements to Vegetables under Field Conditions.
Chemosphere.55: 811-822.
133

Wilcke, W., Kretzschmar, S., Bunt, M., Saboro, G. and Zech, W. (1998). Aluminum and heavy
metal partitioning in A horizons of soils in Costa Rican coffee plantations. Soil Science
163: 463-471.

Woolhouse, H.W. (1993). Toxicity and tolerance in the response of plants to metals. pp. 245300. In: Lange O.L., P.S. Nobel, C.B. Osmond, and H. Ziegler (eds.) Encyclopedia of
Plant Physiology. Volume 12C: Physiological Plant Ecology. Berlin: Springer Verlag.

Xiang, S-R, Doyle, A., Holden, P.A. and Schimel, J.P. (2008). Drying and rewetting effects on C
and N mineralization and microbial activity in surface and subsurface California grassland
soils. Soil Biology and Biochemistry. 40: 2812289.

134

APPENDIX ONE

SAMPLE OUTPUT OF SPECIATION CALCULATION FROM PHREEQC MODEL


TITLE DARETA (Pb SURFACE) pH 4.8
SOLUTION 1 SURFACE SOLUTION
temp
25
pH
4.8
pe
4
redox
pe
units
ppm
density
1
Pb
24.02
Ca
24.5
Na
9.6
S(6)
3.21
Mg
45.6
K
46.8
P
145.2
water
1 # kg
END
----TITLE
----DARETA (Pb SURFACE) pH 4.8
------------------------------------------Beginning of initial solution calculations.
------------------------------------------Initial solution 1.

SURFACE SOLUTION

-----------------------------Solution composition----------------------------Elements
Ca
K
Mg
Na
P
Pb
S(6)

Molality

Moles

6.115e-004
1.197e-003
1.876e-003
4.177e-004
4.689e-003
1.160e-004
3.343e-005

6.115e-004
1.197e-003
1.876e-003
4.177e-004
4.689e-003
1.160e-004
3.343e-005

----------------------------Description of solution--------------------------pH
pe
Specific Conductance (uS/cm, 25 oC)
Density (g/cm3)

135

=
4.800
=
4.000
= 472
=
0.99724

Activity of water
Ionic strength
Mass of water (kg)
Total alkalinity (eq/kg)
Total carbon (mol/kg)
Total CO2 (mol/kg)
Temperature (deg C)
Electrical balance (eq)
Percent error, 100*(Cat-|An|)/(Cat+|An|)
Iterations
Total H
Total O

=
1.000
= 7.967e-003
= 1.000e+000
= 2.317e-005
= 0.000e+000
= 0.000e+000
= 25.000
= 2.043e-003
= 18.39
= 10
= 1.110218e+002
= 5.552511e+001

----------------------------Distribution of species---------------------------

Species
H+
OHH2O
Ca
Ca+2
CaH2PO4+
CaHPO4
CaSO4
CaPO4CaHSO4+
CaOH+
H(0)
H2
K
K+
KSO4KHPO4KOH
Mg
Mg+2
MgH2PO4+
MgHPO4
MgSO4
MgPO4MgOH+
Na
Na+
NaSO4NaHPO4NaOH
O(0)
O2
P
H2PO4MgH2PO4+
CaH2PO4+
HPO4-2
MgHPO4

Molality

Activity

Log
Molality

Log
Activity

Log
Gamma

1.722e-005
6.944e-010
5.551e+001
6.115e-004
5.622e-004
4.444e-005
3.392e-006
1.406e-006
5.569e-010
1.436e-010
4.499e-012
3.550e-021
1.775e-021
1.197e-003
1.197e-003
1.515e-007
3.691e-008
2.378e-013
1.876e-003
1.687e-003
1.707e-004
1.383e-005
4.983e-006
2.266e-009
2.968e-010
4.177e-004
4.177e-004
3.780e-008
1.291e-008
1.584e-013
0.000e+000
0.000e+000
4.689e-003
4.434e-003
1.707e-004
4.444e-005
2.303e-005
1.383e-005

1.585e-005
6.315e-010
9.999e-001

-4.764
-9.158
1.744

-4.800
-9.200
-0.000

-0.036
-0.041
0.000

3.914e-004
4.048e-005
3.398e-006
1.409e-006
5.073e-010
1.308e-010
4.098e-012

-3.250
-4.352
-5.470
-5.852
-9.254
-9.843
-11.347

-3.407
-4.393
-5.469
-5.851
-9.295
-9.883
-11.387

-0.157
-0.041
0.001
0.001
-0.041
-0.041
-0.041

1.778e-021

-20.751

-20.750

0.001

1.089e-003
1.380e-007
3.362e-008
2.382e-013

-2.922
-6.819
-7.433
-12.624

-2.963
-6.860
-7.473
-12.623

-0.041
-0.041
-0.041
0.001

1.180e-003
1.555e-004
1.386e-005
4.992e-006
2.064e-009
2.704e-010

-2.773
-3.768
-4.859
-5.302
-8.645
-9.528

-2.928
-3.808
-4.858
-5.302
-8.685
-9.568

-0.155
-0.041
0.001
0.001
-0.041
-0.041

3.808e-004
3.443e-008
1.176e-008
1.587e-013

-3.379
-7.422
-7.889
-12.800

-3.419
-7.463
-7.930
-12.799

-0.040
-0.041
-0.041
0.001

0.000e+000

-50.881

-50.880

0.001

4.042e-003
1.555e-004
4.048e-005
1.583e-005
1.386e-005

-2.353
-3.768
-4.352
-4.638
-4.859

-2.393
-3.808
-4.393
-4.800
-4.858

-0.040
-0.041
-0.041
-0.163
0.001

136

CaHPO4
KHPO4NaHPO4MgPO4CaPO4PO4-3
Pb
Pb+2
PbSO4
PbOH+
Pb2OH+3
Pb(SO4)2-2
Pb(OH)2
Pb(OH)3Pb(OH)4-2
S(6)
SO4-2
MgSO4
CaSO4
PbSO4
KSO4NaSO4HSO4CaHSO4+
Pb(SO4)2-2

3.392e-006
3.691e-008
1.291e-008
2.266e-009
5.569e-010
1.047e-012
1.160e-004
1.151e-004
8.023e-007
1.070e-007
4.002e-010
1.105e-010
2.387e-012
1.902e-018
3.636e-025
3.343e-005
2.601e-005
4.983e-006
1.406e-006
8.023e-007
1.515e-007
3.780e-008
3.052e-008
1.436e-010
1.105e-010

3.398e-006
3.362e-008
1.176e-008
2.064e-009
5.073e-010
4.504e-013

-5.470
-7.433
-7.889
-8.645
-9.254
-11.980

-5.469
-7.473
-7.930
-8.685
-9.295
-12.346

0.001
-0.041
-0.041
-0.041
-0.041
-0.366

7.922e-005
8.038e-007
9.745e-008
1.728e-010
7.610e-011
2.392e-012
1.732e-018
2.504e-025

-3.939
-6.096
-6.971
-9.398
-9.957
-11.622
-17.721
-24.439

-4.101
-6.095
-7.011
-9.762
-10.119
-11.621
-17.761
-24.601

-0.162
0.001
-0.041
-0.365
-0.162
0.001
-0.041
-0.162

1.804e-005
4.992e-006
1.409e-006
8.038e-007
1.380e-007
3.443e-008
2.780e-008
1.308e-010
7.610e-011

-4.585
-5.302
-5.852
-6.096
-6.819
-7.422
-7.515
-9.843
-9.957

-4.744
-5.302
-5.851
-6.095
-6.860
-7.463
-7.556
-9.883
-10.119

-0.159
0.001
0.001
0.001
-0.041
-0.041
-0.041
-0.041
-0.162

------------------------------Saturation indices-----------------------------Phase
Anglesite
Anhydrite
Gypsum
H2(g)
H2O(g)
Hydroxyapatite
O2(g)
Pb(OH)2

SI log IAP
-1.05
-3.79
-3.57
-17.60
-1.51
-8.82
-47.99
-2.65

-8.84
-8.15
-8.15
-20.75
-0.00
-12.24
-50.88
5.50

log KT
-7.79
-4.36
-4.58
-3.15
1.51
-3.42
-2.89
8.15

-----------------End of simulation.
----------------------------------------------------Reading input data for simulation 2.
---------------------------------------------End of run.
-----------

137

PbSO4
CaSO4
CaSO4:2H2O
H2
H2O
Ca5(PO4)3OH
O2
Pb(OH)2

APPENDIX TWO
Relative distribution of Pb (A) Cd (B) and Zn (C) among the various soil solid fractions
(A)
Pb fractions (mg kg-1)
Sampling Direction
Dareta North

Dareta south

Dareta east

EXCH

CARB

OXIDE OM

RES

Distance (m)
10

25.94

12.97

38.91

64.86

51.88

50

9.37

2.17

2.22

2.79

317.95

100

1.73

6.92

8.65

3.46

5.19

150

1.54

3.42

3.93

5.94

811.67

300

1.44

4.23

1.99

2.58

513.26

500

6.21

4.91

9.85

15.34

221.2

1000

3.22

1.45

10.2

32.1

463.7

10

2.16

83.75

6.02

4.065

621.55

50

47.04

23.52

58.8

11.76

35.28

100

0.47

3.22

4.69

3.286

178.535

150

3.21

2.29

2.16

4.695

640.895

300

51.2

25.6

102.4

76.8

128

500

7.65

2.54

23.5

13.1

231.32

1000

2.45

44.56

1.23

56.31

321.53

10

83.54

41.77

167.08

125.31

208

50

0.69

3.57

2.02

2.82

174

100

1.11

2.41

3.62

3.9

169.71

138

Dareta west

150

1.77

2.41

24.93

3.525

1464.365

300

27.1

54.2

108.4

81.3

135.5

500

8.32

2.41

134.5

162.6

623.13

1000

25.3

4.5

45.67

27.45

125.4

10

5.57

11.14

22.28

16.71

27.85

50

4.68

2.34

7.02

11.7

9.36

100

149.77

449.31

299.54

599.08

748.85

150

7.87

2.03

5.15

5.055

107.095

300

13.48

26.96

53.92

40.44

67.4

500

73.34

36.67

146.68

183.35

110.46

1000

42.98

85.96

171.92

128.28

214.9

139

(B)
Cd Fractions (mg kg-1)
Sampling Direction
Dareta North

Dareta south

Dareta east

EXCH

CARB

OXIDE

OM

RES

Distance (m)
10

13.74

9.16

4.58

18.32

22.9

50

3.495

1.74

6.33

0.3

56.835

100

4.48

3.36

1.12

2.24

5.6

150

1.32

0.39

0.3

0.21

14.58

300

0.21

0.075

0.375

0.24

17.1

500

3.11

0.12

0.42

1.72

16.89

1000

4.2

0.03

1.21

5.62

18.3

10

14.445

5.37

7.365

0.825

9.495

50

2.24

5.6

4.48

3.36

1.12

100

7.44

2.625

5.865

2.985

51.585

150

0.405

0.21

0.165

0.51

69.21

300

16.32

8.16

12.24

20.4

4.08

500

11.23

1.12

6.32

5.87

9.74

1000

12.22

0.98

10.21

5.23

9.85

10

3.12

2.43

4.21

5.68

1.11

50

1.23

0.93

1.29

0.33

12.87

100

0.3

0.24

0.405

0.33

37.725

150

23.7

0.645

0.735

0.39

36.63

300

4.16

12.48

16.64

8.32

20.8

140

Dareta west

500

5.14

1.21

3.85

3.79

18.28

1000

10.21

0.87

3.45

1.98

28.2

10

3.57

4.76

1.19

5.95

2.38

50

3.42

2.28

4.56

5.7

1.14

100

3.45

4.6

2.3

1.15

5.75

150

0.825

0.3

0.225

0.21

19.89

300

2.44

3.66

6.1

1.22

4.88

500

2.24

1.12

3.36

4.48

5.6

1000

0.57

0.38

0.76

0.19

0.95

141

(C)
Zn fractions (mg kg-1)
Sampling Direction
Dareta North

Dareta south

Dareta east

EXCH

CARB

OXIDE

OM

RES

Distance (m)
10

39.34

19.67

59.01

98.35

79.13

50

36.645

14.28

25.815

15.675

58.485

100

13

19.5

6.5

26

32.5

150

13.155

11.565

21.435

20.145

74.4

300

12.53

26.775

61.275

71.85

5.82

500

11.4

9.2

45.1

58.5

73.1

1000

19.8

2.54

3.87

62.7

58.6

10

5.37

23.205

132.36

18.795

194.82

50

6.38

12.76

25.52

19.14

31.9

100

3.06

14.43

61.605

31.47

48.885

150

4.425

16.11

9.825

98.715

22.125

300

25.2

12.26

50.4

37.8

63

500

23.7

8.23

30.34

87.3

63.4

1000

11.23

28.74

12.45

67.29

54.12

10

5.91

11.82

23.64

29.55

17.13

50

29.295

15.825

16.245

19.635

21.45

100

10.59

28.26

32.775

26.805

13.47

150

75.855

11.985

89.49

14.91

137.76

300

18.64

9.32

27.96

46.6

37.28

142

Dareta west

500

17.37

18.14

39.13

37.14

68.5

1000

6.85

1.22

8.21

25.3

88.4

10

10.3

30.9

20.6

41.2

51.5

50

11.6

34.8

23.2

46.4

58

100

16.16

8.08

32.32

24.24

40.4

150

13.98

17.505

26.175

17.85

53.04

300

9.06

18.12

36.24

45.3

27.18

500

4.43

8.86

13.29

17.72

22.15

1000

16.3

32.6

65.2

48.9

81.5

143

APPENDIX THREE

(A). Relative distribution of free species and inorganic complexes of Pb in contaminated soil solution of Dareta
Pb2+
Equilibrium
concentration

pH

PbSO4

PbCl+

PbOH+

Pb2OH+3 Pb(SO4)22- PbCl2

Pb(OH)2 PbCl3

Pb(OH)3- PbCl42-

Pb(OH)2-

4.8

24.02

-4.091

-6.055

-6.665

-7.001

-9.741

-10.05 -10.639

-11.611

-14.913

-17.751

-19.407

-24.591

5.4

9.62

-4.487

-6.48

-6.983

-6.797

-9.933

-10.503 -10.879

-10.807

-15.075

-16.347

-19.491

-22.587

5.8

1.74

-5.232

-7.259

-7.756

-7.143

-11.025

-11.315 -11.679

-10.753

-15.903

-15.893

-20.346

-21.733

6.6

1.54

-5.312

-7.279

-7.837

-6.403

-10.365

-11.275 -11.762

-9.913

-15.986

-13.513

-20.278

-18.533

7.1

1.26

-5.456

-7.449

-7.914

-6.066

-10.172

-11.473 -11.772

-8.376

-15.93

-12.216

-20.308

-16.756

7.3

0.05

-6.9

-8.846

-9.359

-7.31

-12.859

-12.822 -13.218

-9.42

-17.377

-13.06

-21.757

-17.4

7.6

0.022

-7.349

-9.285

-9.84

-7.459

-13.25

-13.25 -13.731

-9.269

-17.922

-12.609

-22.333

-16.649

8.2

0.001

-9.024

-10.953

-11.472

-8.534

-14.911

-14.911 -15.406

-9.744

-19.597

-12.484

-24.007

-15.929

144

(B). Relative distribution of free species and inorganic complexes of Cd in contaminated soil solution of Dareta
Cd2+
Equilibrium
concentration

pH

CdCl+

CdSO4

CdOH+

CdCl2

Cd(SO4)22-

CdCl3-

Cd(OH)2

Cd(OH)3-

Cd(OH)42-

4.8

18.9

-3.929

-6.127

-6.182

-9.209

-9.685

-9.854

-14.063

-14.679

-22.83

-32.08

5.4

13.93

-4.06

-6.18

-6.346

-8.74

-9.659

-10.052

-13.959

-13.61

-21.16

-29.81

5.8

10.25

-4.194

-6.341

-6.516

-8.474

-9.847

-10.257

-14.174

-12.944

-20.094

-28.194

6.6

3.94

-4.613

-6.759

-6.871

-8.073

-10.265

-10.549

-14.59

-11.723

-18.053

-25.483

7.1

3.71

-4.649

-6.728

-6.934

-7.629

-10.167

-10.639

-14.426

-10.799

-16.649

-23.599

7.3

1.58

-5.024

-7.104

-7.261

-7.804

-10.544

-10.918

-14.803

-10.775

-16.425

-23.175

7.6

0.93

-5.26

-7.327

-7.486

-7.74

-10.843

-11.132

-15.134

-10.41

-15.761

-22.211

8.2

0.01

-7.231

-9.342

-9.45

-9.111

-12.813

-13.089

-17.104

-11.181

-15.931

-21.781

145

(C). Relative distribution of free species and inorganic complexes of Zn in contaminated soil solution of Dareta
Zn2+
Equilibrium
concentration

pH

ZnSO4

ZnCl+

ZnOH+

Zn(SO4)22- Zn(OH)2

ZnCl2

ZnCl3-

Zn(OH)3- ZnCl42-

Zn(OH)42-

4.8

33.9 -3.441

-5.8

-7.186

-7.601

-9.62

-10.741 -11.341 -15.466

-17.441 -19.942

-25.441

5.4

24.2 -3.584

-5.972

-7.251

-7.144

-9.821

-9.684 -11.329 -15.376

-15.784 -19.774

-23.184

5.8

9.6 -3.983

-6.397

-7.677

-7.143

-10.271

-9.283 -11.781 -15.856

-14.983

-20.28

-21.983

6.6

8.12 -4.061

-6.414

-7.756

-6.401

-10.226

-7.721 -11.862 -15.937

-12.601 -20.362

-18.781

7.1

7.44 -4.111

-6.49

-7.739

-5.971

-10.33

-6.811 -11.778 -15.786

-11.211 -20.145

-16.911

7.3

4.4 -4.347

-6.676

-7.976

-6.007

-10.466

-6.647 -12.016 -16.025

-10.847 -20.385

-16.347

7.6

2.21 -4.661

-6.979

-8.323

-6.021

-10.756

-6.361 -12.394 -16.435

-10.262 -21.299

-15.462

8.2

0.95 -5.136

-7.445

-8.797

-5.896

-11.214

-5.636 -12.867 -16.908

146

-8.936

-20.28

-13.536

APPENDIX FOUR
Range, mean concentration, and standard error of Pb, Cd, and Zn in Dareta

Sampling
Direction

Pb
(mgkg-1)

Cd
(mgkg-1)

Zn
(mgkg-1)

Dareta North (n=7)

25.95-523.5
(381 138.16c)

16.8-68.7
(27.76 10.19b)

97-295
(172.77 33.37b)

Dareta South (n=7)

190.2-717.15
(424.2 113.15b)

16.8-70.5
(40.68 11.05a)

95-374
(193.98 47.60a)

Dareta East

(n=7)

180.75-1497
(578.88 243.87a)

16.65-62.4
(39.48 10.13a)

88.65 330
(154.56 44.65c)

Dareta West (n=7)

35.1-2246.55
(555.69 295.98a)

1.85 21.45
(15.94 0.79c)

66.45-244.5
(145.5 7.20c)

Means followed by the same letters are not statistically different at 5% probability level

147

Das könnte Ihnen auch gefallen