14
Heavy metal distribution and controlling factors within coastal plain sediments, Bells Creek catchment, southeast Queensland, Australia Tania Liaghati * , Micaela Preda, Malcolm Cox Received 24 November 2002; accepted 27 February 2003 Abstract Bells Creek catchment in southeast Queensland (Australia) is a non-industrialised coastal plain limited to small settlements and agricultural land. A study was initiated to examine elevated metal concentrations and to assess horizontal and vertical distribution of those elements. Ninety- nine samples were analysed for Cr, V, Ni, Cu, Zn, Pb, As, Fe, Mn and Al. Total organic carbon, sulfur content and mineralogy of samples along with land-use practices across the catchment were used to identify processes which influence metal distribution. A comparison between metal concentration within the study area and mean heavy metal content of standard sandstone showed that except for Mn, all other metals showed elevated levels throughout the catchment. When metal concentrations were compared to parent bedrock, however, it was concluded that elevated levels are likely to be natural. A normalisation procedure was applied to the data set and this analysis validated that elevated trace metal concentrations in most samples are not due to artificial contamination. While surficial estuarine sediments were only enriched in V, soils were dominantly enriched in Cr, Zn and V. Overall, geochemistry and mineralogy of the samples show the effect of both natural and anthropogenic inputs to the catchment, however, natural processes are more dominant than anthropogenic inputs in concentrating metals. D 2003 Elsevier Ltd. All rights reserved. Keywords: Trace metals; Fluvial/estuarine sediments; Geochemistry; Mineralogy; Bells creek catchment; Normalisation; Enrichment factor 1. Introduction Estuarine and marine sediments are sinks for various metals transported from the land. Metals may be mobilised as a result of natural processes (e.g. weathering and erosion of geological formations) as well as by anthropogenic activity. In the mobilisation process, trace elements may be adsorbed by clays, can complex with organic compounds or may co-precipitate with oxides and hydroxides. As many metals occur naturally in weathered materials and drainage systems due to their presence in local rocks, the relative influence of natural and anthropogenic sources on the geo- chemistry of coastal sediments is not always clear. There- fore, for a better assessment of metal distributions within such environment, it is important to distinguish between metallic elements released by natural processes and those introduced by human-related activities. The amounts of trace elements in natural systems can be of environmental significance because where elevated they may contaminate surface and shallow groundwater. In addition, marine organisms and vegetation in coastal environments can uptake metals, increasing the potential for the entry of some metals into the food chain. Furthermore, while sediment data is useful for describing trace metal occurrence and assessing their distribution in coastal plains, in spatial or vertical trend studies such as the present investigation, errors associated with sampling and analysis can make it difficult to detect differences between sites (e.g. Kelly et al., 1994). To establish the potential environmental impact of con- taminated sediments, vertical sediment profiles obtained by coring are important because they can preserve the historical sequence of pollution and, at the same time, enable a reaso- nable estimation of the background levels and the variations in input of pollutants over an extended period of time. Vertical sediment sections provide a record of level of contamination over time, if pollutants are persistent and the sediment stratum has not been greatly disturbed by human activities such as dredging (e.g. Forstner et al., 1984; Fung, 1993). Preda and Cox (2000) conducted a baseline study of trace metal distribution in the Pumicestone coastal area of south- east Queensland. That study showed that regardless of the sample origin and composition (bedrock, estuarine sediment, soil or mangrove tissue), metal occurrence is systematic 0160-4120/$ - see front matter D 2003 Elsevier Ltd. All rights reserved. doi:10.1016/S0160-4120(03)00060-6 * Corresponding author. Tel.: +61-7-3864-4185; fax: +61-7-3864- 1535. E-mail address: [email protected] (T. Liaghati). www.elsevier.com/locate/envint Environment International 29 (2003) 935 – 948

Heavy metal distribution and controlling factors …dzumenvis.nic.in/Microbes and Metals Interaction/pdf...Heavy metal distribution and controlling factors within coastal plain sediments,

  • Upload
    others

  • View
    5

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Heavy metal distribution and controlling factors …dzumenvis.nic.in/Microbes and Metals Interaction/pdf...Heavy metal distribution and controlling factors within coastal plain sediments,

www.elsevier.com/locate/envint

Environment International 29 (2003) 935–948

Heavy metal distribution and controlling factors within coastal plain

sediments, Bells Creek catchment, southeast Queensland, Australia

Tania Liaghati*, Micaela Preda, Malcolm Cox

Received 24 November 2002; accepted 27 February 2003

Abstract

Bells Creek catchment in southeast Queensland (Australia) is a non-industrialised coastal plain limited to small settlements and agricultural

land. A study was initiated to examine elevated metal concentrations and to assess horizontal and vertical distribution of those elements. Ninety-

nine samples were analysed for Cr, V, Ni, Cu, Zn, Pb, As, Fe, Mn and Al. Total organic carbon, sulfur content and mineralogy of samples along

with land-use practices across the catchment were used to identify processes which influence metal distribution. A comparison between metal

concentration within the study area and mean heavy metal content of standard sandstone showed that except for Mn, all other metals showed

elevated levels throughout the catchment.Whenmetal concentrations were compared to parent bedrock, however, it was concluded that elevated

levels are likely to be natural. A normalisation procedure was applied to the data set and this analysis validated that elevated trace metal

concentrations in most samples are not due to artificial contamination. While surficial estuarine sediments were only enriched in V, soils were

dominantly enriched in Cr, Zn and V. Overall, geochemistry and mineralogy of the samples show the effect of both natural and anthropogenic

inputs to the catchment, however, natural processes are more dominant than anthropogenic inputs in concentrating metals.

D 2003 Elsevier Ltd. All rights reserved.

Keywords: Trace metals; Fluvial/estuarine sediments; Geochemistry; Mineralogy; Bells creek catchment; Normalisation; Enrichment factor

1. Introduction

Estuarine and marine sediments are sinks for various

metals transported from the land. Metals may be mobilised

as a result of natural processes (e.g. weathering and erosion

of geological formations) as well as by anthropogenic

activity. In the mobilisation process, trace elements may

be adsorbed by clays, can complex with organic compounds

or may co-precipitate with oxides and hydroxides. As many

metals occur naturally in weathered materials and drainage

systems due to their presence in local rocks, the relative

influence of natural and anthropogenic sources on the geo-

chemistry of coastal sediments is not always clear. There-

fore, for a better assessment of metal distributions within

such environment, it is important to distinguish between

metallic elements released by natural processes and those

introduced by human-related activities.

The amounts of trace elements in natural systems can be of

environmental significance because where elevated they may

0160-4120/$ - see front matter D 2003 Elsevier Ltd. All rights reserved.

doi:10.1016/S0160-4120(03)00060-6

* Corresponding author. Tel.: +61-7-3864-4185; fax: +61-7-3864-

1535.

E-mail address: [email protected] (T. Liaghati).

contaminate surface and shallow groundwater. In addition,

marine organisms and vegetation in coastal environments can

uptake metals, increasing the potential for the entry of some

metals into the food chain. Furthermore, while sediment data

is useful for describing trace metal occurrence and assessing

their distribution in coastal plains, in spatial or vertical trend

studies such as the present investigation, errors associated

with sampling and analysis can make it difficult to detect

differences between sites (e.g. Kelly et al., 1994).

To establish the potential environmental impact of con-

taminated sediments, vertical sediment profiles obtained by

coring are important because they can preserve the historical

sequence of pollution and, at the same time, enable a reaso-

nable estimation of the background levels and the variations

in input of pollutants over an extended period of time. Vertical

sediment sections provide a record of level of contamination

over time, if pollutants are persistent and the sediment stratum

has not been greatly disturbed by human activities such as

dredging (e.g. Forstner et al., 1984; Fung, 1993).

Preda and Cox (2000) conducted a baseline study of trace

metal distribution in the Pumicestone coastal area of south-

east Queensland. That study showed that regardless of the

sample origin and composition (bedrock, estuarine sediment,

soil or mangrove tissue), metal occurrence is systematic

Page 2: Heavy metal distribution and controlling factors …dzumenvis.nic.in/Microbes and Metals Interaction/pdf...Heavy metal distribution and controlling factors within coastal plain sediments,

T. Liaghati et al. / Environment International 29 (2003) 935–948936

which suggests an interaction between various natural sys-

tems. With respect to controls over adsorption and mobilisa-

tion of trace metals, the above study looked only at the

influence of grain size, and iron and manganese oxides. This

current study examines the occurrence and distribution of

metals within a sub-catchment of the Pumicestone region in

detail and investigates other factors which could control

metal chemical behaviour; factors considered are sediment

source (fluvial/estuarine), organic matter as organic carbon,

mineralogy (with special regard to clay speciation) and local

land-use practices (for example, pine plantations versus

native vegetation). Thus, the present study complements

previous investigations (Cox et al., 2000; Preda and Cox,

2000) that have focused on the geological settings and

occurrence of minor and trace metals within a broader area.

A range of studies have been conducted on trace metal

distribution in estuarine and coastal plain sediments

throughout the world. The majority of these studies, how-

ever, have concentrated on polluted areas with the objective

of describing trace metal concentration levels and/or iden-

tify sources of the pollutants (e.g. Ellaway et al., 1982;

Daskalakis and O’Connor, 1995; Schneider and Davey,

1995; Angelidis and Aloupi, 1997; Data and Subramanian,

1998; Power et al., 1999; Angelidis and Aloupi, 2000;

Owen and Sandhu, 2000; Chen et al., 2001; Ruiz, 2001;

De Carlo and Anthony, 2002). Determining background

metal levels in unpolluted areas, however, has not been

studied widely. Identifying naturally elevated metal concen-

tration is important because some metals can appear to be

enriched, although when compared to their source (e.g.

bedrock material), it is concluded that such elevation may

still be natural. Some studies have examined background

Fig. 1. Location of the study area in

levels (e.g. Windom et al., 1989; McMurtry et al., 1995),

and others have concentrated on identifying the major

factors which influence the geochemical character of sedi-

ments, their mobility and transport processes (Forstner et al.,

1982, 1984; Arakel and Hongjun, 1992). In addition,

broader studies on the natural physico-chemical processes

which affect metal levels within heterogeneous coastal

settings are scarce. Here we report an investigation into

the spatial and vertical distribution of trace elements within

a subtropical catchment; the study aimed to assess factors

such as land-use practices and sediment/soil character and

their role over trace metal concentrations within a non-

industrialised coastal plain. From a geochemical point of

view, the ability to establish the average composition of

metals in parent rock material is significant, as it can be used

as a basis for comparison and recognition of anomalous

concentrations of metals in soils and sediments. These metal

concentrations in soils and sediments were therefore com-

pared to parent material to identify enrichment in a setting

with potentially low levels of pollution and a variety of

material textures and land-use practices. Due to high varia-

tion in sediment characteristics, however, the large geo-

chemical data set produced could not be interpreted using

absolute concentration values in isolation. Thus, to enable a

more effective interpretation, a normalisation procedure was

applied to the data set.

2. Features of the study area

The study area is the Bells Creek catchment, which is

located 80 km north of Brisbane (Fig. 1) and is com-

relation to Pumicestone region.

Page 3: Heavy metal distribution and controlling factors …dzumenvis.nic.in/Microbes and Metals Interaction/pdf...Heavy metal distribution and controlling factors within coastal plain sediments,

T. Liaghati et al. / Environment International 29 (2003) 935–948 937

prised of three sub-catchments: Bells, Lamerough and

Halls (Fig. 2). This catchment is within the northern part

of the broader Pumicestone Catchment and drains into

Pumicestone Passage.

Pumicestone Passage forms a unique tidal water body in

southeast Queensland. It is semi-enclosed and occurs

between the narrow coastal plain of the mainland and Bribie

Island, a large sand barrier island. The mainland coastal

plain is underlain by sandstone bedrock (Preda et al., 2000).

The large input of suspended material carried by the 10

creek systems that flow into the passage has the potential to

contain nutrients and metals, which can affect local water

quality.

2.1. Geological setting

The Bells Creek coastal plain developed in the Late

Quaternary as a result of fluvial and marine processes. As

the sea level rose during the last marine transgression, the

river channel and floodplain drowned with formation of

estuarine settings. Overall, the highest sea level was

Fig. 2. Sampling sites with respect to surficial material

reached around 6500 years ago (Williams et al., 1998);

in the Pumicestone region, sea level fell to its present

position around 3000 years BP (Flood, 1981; Lester,

2000). The estuarine sedimentary units that formed due

to changing sea levels are important local features of this

coastal plain.

Landsborough Sandstone (Triassic–Jurassic) is the dom-

inant bedrock throughout the Pumicestone region. The

formation consists of quartzo-feldspathic sandstone with

significant amounts of lithic fragments of volcanic origin

and minor shale, pebble conglomerate and coal, deposited

in a Mesozoic fluvial environment (e.g. Murphy et al.,

1987). The mineralogy of this bedrock is significant, as it

is likely to have strongly influenced the natural background

levels of the metals investigated here (e.g. Preda and Cox,

2001).

Based on a report produced by Geological Survey of

Queensland (1999) and a soil survey conducted for the

study area (NSR Environmental Consultants, 1999b), four

distinct soil types were identified throughout the area: (1)

alluvial, (2) sand of marine origin, (3) organic clay and (4)

(after NSR Environmental Consultants, 1999a,b).

Page 4: Heavy metal distribution and controlling factors …dzumenvis.nic.in/Microbes and Metals Interaction/pdf...Heavy metal distribution and controlling factors within coastal plain sediments,

T. Liaghati et al. / Environment International 29 (2003) 935–948938

weathered bedrock (soils developed from the weathering

sandstone bedrock). The ‘‘organic clay’’ in the region varies

from clayey sand to medium heavy clay and typically occurs

below 80 cm depth (Fig. 2).

2.2. Climate and land use

The climate of the Bells Creek catchment like most of

southeast Queensland is subtropical humid with hot wet

summers and mild dry winters. Average monthly rainfall in

1998–1999 was 242 mm with the highest rainfall occur-

ring in late summer and autumn. Mean daily temperatures

range from 15 to 30 jC in summer and 9 to 20 jC in

winter.

Within the region there is a wide range of land-use

practices including pasture, forestry (pine plantation and

associated milling and timber processing), as well as urban

and rural residential; coastal native vegetation largely con-

sists of mangroves, Melaleuca and heath. Land-use patterns

within the catchment are changing continually, largely due

to population growth. As a result, from 1974 to 1991, the

amount of native vegetation decreased dramatically as other

land uses such as pine plantation and urban development

increased (NSR, 1999a) (Fig. 3).

Fig. 3. Land-use practices throughout

3. Sampling strategy and analytical methods

3.1. Sampling design

The following materials were analysed in this study to

determine the distribution of trace and major metals within

sediments of the catchment and to examine the effect of

various land-use and geological settings.

1. Bedrock samples (Landsborough Sandstone): to establish

the natural occurrence of trace metals (22 samples from 7

cores).

2. Soil samples of fluvial/estuarine origin referred to as

‘soils’: located above the water table and currently away

from direct tidal influence (53 samples from 15 cores).

3. Recent surficial fluvial/estuarine sediments referred to as

‘sediments’: 21 samples collected along the three major

creeks discharging into Pumicestone Passage.

3.2. Sample collection

Grab sediment samples (category 3) were collected from

creek banks; samples were analysed for extractable cations,

organic C, total S and mineral composition. Bedrock and

catchment and soil sample sites.

Page 5: Heavy metal distribution and controlling factors …dzumenvis.nic.in/Microbes and Metals Interaction/pdf...Heavy metal distribution and controlling factors within coastal plain sediments,

Table 1

Recovery of total metal concentration (mg/kg)

Element Certifieda Foundb Recovery

(%)

Precision

(%RSD)

V 133F 5 62F 1.7 46.6 3

Cr 90.7F 4.6 42F 1.2 46.3 3

Mo 5.43F 0.28 3.4F 0.4 62 12

Co 11.5F 0.3 8.2F 0.3 71 4

Ni 39.5F 2.3 25F 0.7 63 2

Cu 310F 12 289F 6 93 2

Zn 364F 23 346F 8 95 2

Pb 183F 8 166F 7 91 4

As 26.2F 1.5 21F 0.5 80 2

Fe 40,900F 6000 27,628F 828 68 3

Mn 440F 19 208F 6 47.3 3

Al 66,200F 3200 13,385F 534 20 4

a Certified Standard Reference Material, PACS-2 (harbour sediment,

Canada).b

T. Liaghati et al. / Environment International 29 (2003) 935–948 939

soil samples (categories 1 and 2) were powdered archive

material, provided by Queensland Acid Sulfate Soils Inves-

tigation Team (QASSIT). These samples were collected

from auger holes as part of a large program, the purpose

of which was to map pyritic sediments in southeast Queens-

land in 1997.

3.3. Physical and chemical analyses

Sediment pH was measured in the laboratory by inserting

a probe into unconsolidated samples which had been stored

under refrigeration. Extractable cations were analysed on

total sample, dried to 80 jC, by digesting the sediment in

aqua regia [1:3 nitric acid (HNO3)/hydrochloric acid (HCl)]

following a procedure described by Loring and Rantala

(1992). This acid mixture is significant in the above proce-

dure because it is not able to attack the silicate lattice, and

therefore, only cations absorbed by clays, organic matter and

oxides are extracted into solution and can be analysed. The

trace elements analysed included: V, Cr, Mo, Co, Ni, Cu, Zn,

Pb and As; minor metals analysed included: Fe, Mn and Al.

Extractable Al represents clay minerals, therefore, is consid-

ered to be a measure of grain size (Chao, 1984; Koljonen and

Malisa, 1991; Raisanen, 1996; Osterhom and Astrom, 2002).

Quantification was achieved using a Varian Liberty 200

inductively coupled plasma optional emission spectrometer

(ICP-OES). The instrument was calibrated using synthetic

standards and a reagent blank (Loring and Rantala, 1992).

Total S was measured using LECO and total organic carbon

(TOC) concentrations were obtained by digesting 0.5 g of

sample with dichromate according to the method of Walkey

and Black (in Loring and Rantala, 1992).

3.4. Quality control procedures

The precision and recovery level of the analytical proce-

dure for acid digestion (aqua regia) and metal determination

by ICP-OES were tested using marine sediment reference

standard PACS-2 (National Research Council Canada). This

is a certified material (CRM) for total digestion, as no CRM

was available for the extractable method using aqua regia.

Recoveries of most metals from the reference material were

>60%, except for V, Cr, Al and Mn (Table 1); vanadium and

Cr are refractory metals, which strongly adsorb to siliceous

materials. Moreover, Al is part of the aluminosilicate matrix

and the aqua regia digestion method is not robust enough to

extract these metals. A low recovery percentage for Cr and Al

had also been recorded by previous workers using similar

digestion method (Tam and Yao, 1998). Higher recovery can

be obtained if perchloric or hydrofluoric acids are used;

however, such digestion will bias results for environmental

assessment. Mean and standard deviations of the replicate

samples were calculated. A comparison between the standard

deviation of certified samples and replicate samples in this

study revealed that repeatability generally was < 5 (%RSD)

even when recoveries were low (Table 1).

3.5. Mineralogical analyses

Mineral composition of bedrock and unconsolidated

material was determined using X-ray diffraction (Philips

PW 1050 diffractometer equipped with a cobalt anticathode)

on non-oriented samples and on oriented specimens for clay

identification. Quantification of mineral phases was assisted

by SIROQUANT (quantification program which expresses

the composition of crystalline material within a sample as

percentage of dry weight).

3.6. Statistical methods

3.6.1. Descriptive data analysis and correlation coefficient

Descriptive data analyses (mean, standard deviation,

maximum and minimum concentrations) were carried out.

In addition, correlation factors were calculated to determine

relationships among different metals. Understanding such

relationships may help to clarify the path by which individ-

ual metals are carried and deposited within the estuary and

helps to determine the processes involved. Metal oxides

such as iron oxyhydroxides and manganese oxides, and

organic carbon commonly act as scavengers for heavy

metals. Therefore, correlations between any metal oxides

and individual heavy metals may help to understand the

processes which result in particular metal associations. As

geochemical data such as those produced in this study are

not normally distributed, Spearman’s rank correlations were

used as they are more appropriate than a simple linear

correlation (e.g. Pearson). Descriptive analysis was pro-

duced using Excel v.2000 and Spearman’s correlation

matrix was generated using SPSS v.10.

3.6.2. Normalisation procedure

It is well established that trace metals may be introduced

to coastal environments by both natural processes (e.g.

weathering and erosion) and human activities within the

catchment or adjacent to the coast (e.g. McConchie et al.,

Mean and standard deviation of three replicated extractions are shown.

Page 6: Heavy metal distribution and controlling factors …dzumenvis.nic.in/Microbes and Metals Interaction/pdf...Heavy metal distribution and controlling factors within coastal plain sediments,

Table 2

Concentration ranges of trace elements in different material types

Metals Mean heavy metal

content of standard

sandstone (mg/kg)

Concentration (mg/kg and % for S and TOC)

(a) (b)

(1) Bedrock,

n= 25

(2) Recent

estuarine

sediment,

n= 21

(3) Soils,

n= 53

V 30 nr 1.8–40.8 2.2–33 0.4–47.7

Cr 35 120 1.7–16.2 1.0–26.0 1.0–47.0

Mo 0.2 nr 1–14.4 1–4.9 1–15.3

Co 0.3 nr 1.2–3.3 1.2–4.7 1.2–6.6

Ni 9 nr 1.5–4.7 1.5–6.4 1.5–15.0

Cu 30 15 0.5–14.5 0.5–5.8 0.5–9.6

Zn 30 16 20.6–53.9 18.2–68.0 16.8–62.9

Pb 10 14 3–18.6 3–7.3 3–17.3

As 1 1 3–8.4 3–10.8 3–13.7

Fe nr 19,000 7600–27,000 9400–41,000 1800–35,000

Mn 460 392 0.4–20.8 0.4–229 0.4–45.0

Al nr 32,000 537–19,342 955–9821 210–28,112

S – – < 1–0.1 < 1–0.6 < 1–1.5

TOC – – < 1–8.6 0.3–5.5 < 1–10.3

(a) Adapted from Krauskopf (1967), Rose et al. (1979) and Alloway (1995).

(b) Leckie and Parks (1978).

nr = not reported.

T. Liaghati et al. / Environment International 29 (2003) 935–948940

1988; Chakrapani and Subramanian, 1993; Niencheski et

al., 1994; Balls et al., 1997; Power et al., 1999; Rubio et al.,

2000; Preda and Cox, 2001; Ruiz, 2001). Estuarine and

coastal sediments, which act as sinks for these metals, are

regarded as a mixture of inorganic and organic material. For

some metals, however (e.g. Cd and Hg), organic material

may be a metal carrier, but due to its low abundance ( < 5%

by weight) in most sediments, it is not usually considered as

major contributor to total metal levels (Loring, 1991).

Concentration of trace elements in natural estuarine and

coastal marine sediments are largely determined by inor-

ganic material resulting from physical and chemical weath-

ering of landmasses. Inorganic detritus is composed mainly

of a limited number of silicate minerals such as quartz,

feldspar, micas and clay minerals and smaller amounts of

metal oxides and sulfide phases. Of these minerals, the

clays are finer and tend to adsorb more metals than courser

material (Windom et al., 1989). Therefore, natural varia-

tions in metal concentrations have often been related to the

‘‘grain size effect’’ and analyses have been carried out on a

specific size fraction to correct for natural variability (e.g.

Forstner and Salomons, 1980). This approach, however,

requires a separation step and concentrations in a certain

size fraction may not reflect the concentration in the total

sediment.

To compensate for this natural variability, metal concen-

trations were normalized. This procedure can be done by

calculating the ratio of natural concentrations to that of a

normalizing factor whose concentration is not affected by

anthropogenic processes (Daskalakis and O’Connor, 1995).

There is no consensus on the appropriate sediment constit-

uent to be used for normalisation; however, two broad

categories have been well established: granulometric and

geochemical. Granulometric techniques rely on normalisa-

tion against total weight percent fines ( < 62.5 Am) or the

total clay size particles ( < 4 Am) present in the sediment

(Loring, 1991). Geochemical methods are based on a

comparison between metal concentration in sediment and

the concentration of other ‘‘reference’’ elements (Trimble

and Hoenstine, 1997). Reference elements that have been

used previously include aluminium, chromium, iron,

organic carbon and lithium (Windom et al., 1989; Loring,

1990, 1991; McMurtry et al., 1995; Daskalakis and O’Con-

nor, 1995; Balls et al., 1997; Trimble and Hoenstine, 1997;

Tam and Yao, 1998; Fang and Hong, 1999).

In all the above examples, normalisation of element con-

centration assumes a linear relationship between either

geochemical or sedimentological characteristics and the

element of interest. Regardless of the type of normalizing

method used, the concentration of normalizing metal is used

to establish the relationship between natural trace metal

concentrations in sediments from different areas. Overall,

geochemical normalisation is superior to granulometric

methods, as it compensates for both mineralogical and the

natural granular variability of trace metal concentrations in

sediments (Loring, 1991).

One of the drawbacks of the geochemical approach is

that it generates a ratio instead of a total concentration. This

problem can be overcome by standardising the contents to a

reference material and by defining an enrichment factor

(EF). For instance, EF (for Zn relative to Al)=(Zn/Al

sample)/(Zn/Al reference material). The validity of such

an enrichment factor will differ with values used for the

reference material. Most workers have used metal concen-

trations in the Earth’s crust as reference for interpreting the

results. This approach has several limitations, however, and

was not used in this study. Concentrations for crustal

abundances are not appropriate because they neither repre-

sent regional background level nor the analytical uncertain-

ties associated with their measurements (Loring, 1991).

Further, the metal concentration of the Earth’s crust reported

in the literature (e.g. Taylor, 1964) is based on total

digestion and the trace metal concentrations reported in this

study are only extractable cations. As a result, normalizing

with the metal concentration of the Earth’s crust will not

reflect the real situation (e.g. Fang and Hong, 1999). The

present study, therefore, has used a geochemical approach to

test correlations between concentrations of trace elements

and three candidate-normalizing factors: iron, aluminium

and total organic carbon.

4. Results and interpretation

4.1. Mineral and chemical character of the samples

While concentrations of trace metals determined within

each of the three categories of material analysed are broadly

similar, distinct differences do exist (Table 2).

Page 7: Heavy metal distribution and controlling factors …dzumenvis.nic.in/Microbes and Metals Interaction/pdf...Heavy metal distribution and controlling factors within coastal plain sediments,

nt International 29 (2003) 935–948 941

4.1.1. Soils developed on bedrock

Primary minerals such as quartz, feldspars (dominantly

plagioclase) and secondary minerals (e.g. hematite and

different types of clays) are major components of the

Triassic–Jurassic sandstone, which underlies the Late Qua-

ternary sediments. Due to variable degrees of weathering,

this bedrock is spatially heterogeneous and contains 48.5–

82% quartz. In fresher less weathered samples, feldspars

occur up to 14%. The most abundant type of clay through-

out bedrock samples is kaolinite followed by smectite and

traces of illite. Hematite was the only iron oxide detectable

in bedrock samples. Samples with a high proportion of

secondary minerals (e.g. kaolinite and smectite) contained

higher concentration of metals such as V, Cr and Zn as well

as Fe, showing the greater capacity of clay-rich weathered

sediments for metal adsorption. The dominant trace metals

were Zn and V, followed by Pb and Cr. Metals such as As,

Co, Cu, Ni and Mo occur at lower concentrations commonly

around the detection limit for the analytical method used.

4.1.2. Soils of fluvial/estuarine origin

The most abundant primary minerals were quartz (20–

100%) followed by feldspars (0–10%). The clay component

was made up of kaolinite (0–61%), smectite (0–7.3%),

illite–smectite mixed layers (0–33.7%) and occasional illite

(e.g. 3.6% in sample 661 at a depth of 80 cm). Minerals such

as pyrite, hematite and jarosite (oxidation products of pyrite)

occur only in some samples located in the estuarine section.

Of note was the positive correlation between these minerals

and the iron content of the sample in which they were found.

Further clay analysis revealed that in almost all samples, the

peak for kaolinite was asymmetrical. After glycolation not

only did the peak become more asymmetrical but also the d-

spacing increased (e.g. in sample 563–130, d-spacing

changed from 7.23 to 7.31 A). As smectite is the only type

of clay mineral in the study area that shifts after glycolation,

the asymmetrical kaolinite peak was believed to be due to the

presence of a kaolinite–smectite (K–S) mixed layer. Sam-

ples 563–130 and 655–380 were good examples because

they contained significant amounts of kaolinite (51.5% and

61%, respectively); therefore, the shift for the peak after

glycolation was easily detectable. The occurrence of K–S

mixed layer is significant as this phase is likely to have a

higher capacity to adsorb metals compared to kaolinite alone.

The reason for the occurrence of this mixed layer clay will

require a more detailed mineralogical investigation.

Due to existence of a bedrock with little lithological

variation (Landsborough Sandstone), the clay speciation is

homogeneous throughout the catchment regardless of sedi-

ment type and location. However, clay minerals may have

depositional significance within the catchment. For exam-

ple, due to limited physical and chemical reworking locally,

the alluvial material was expected to contain more smectite.

The results show that smectite was concentrated in the

upper, fresher sections of the catchment; downstream,

smectite disappears probably due to either weathering to

T. Liaghati et al. / Environme

kaolinite or physical removal by tidal currents. The latter

process has been documented in estuarine conditions where

the deposition of smectite is not favoured, as this mineral is

preferentially retained in suspension (because of its small

size and platy form) and tends to be transported directly to

the near-shore environment (e.g. Chamley, 1989).

Within the fluvial/estuarine soils, the dominant trace

metals were Zn and V, followed by Cr and Pb. This pattern

was preserved in the A- and C-horizons with exception of

some samples in which Cr concentration was higher than that

of V. In the B-horizon, however, this pattern was not clear,

possibly due to the limited number of samples taken from this

horizon. Fig. 4 shows the distribution of Vand Cr throughout

different soil horizons which demonstrates that regardless of

the metal, A is the most leached horizon, while B was the

horizon with the most metal concentration. Metal concen-

trations are shown to be strongly related to the mineralogy of

the materials. Figs. 5–7 summarise lateral (e.g. H1 and BN5

from upper and lower sections of the catchment) and vertical

(e.g. core 651) correlations between geochemistry and min-

eralogy of the samples. The influence of mineral composition

over trace element distribution (trends shown in graphs) is

very clear for all samples analysed.

4.1.3. Recent surficial fluvial/estuarine sediments

The mineralogy of recent fluvial/estuarine sediments was

very similar to soils of fluvial/estuarine origin and the parent

material. Distribution patterns for the primary minerals such

as quartz are very similar to that of soils (up to 89%);

feldspars, however, were more abundant compared to soils

(up to 23%). The clay assemblage was very similar to soil

samples: kaolinite (6.5–30%), smectite (0–5.5%), illite–

smectite mixed layers (0–3%) and occasional illite (e.g. in

H1 and BS4, 2.7% and 7.2%, respectively). The dominant

trace metals were Zn and V, followed by Cr and Pb, while Fe

occurred as pyrite, hematite and goethite in surficial sedi-

ments. The strong correlation between trace metal concen-

trations and sample mineralogy was preserved similar to that

seen in soil samples.

In summary, the comparison of metal concentration

between bedrock, recent surficial fluvial/estuarine, and soils

with fluvial/estuarine origin showed that the feldspathic

signature of the sandstone was preserved in all unconsoli-

dated sediments. In terms of metal occurrence, Zn was the

dominant trace metal followed by V and Cr. In unconsoli-

dated sediments this overall pattern was also preserved, with

the exception of Cr, which in some areas was more abundant

than V. Elevation of Cr levels may be due to the overall

immobility of the metal. Considering the mineralogical and

geochemical data, it can be concluded that the Landsborough

Sandstone is the primary source of unconsolidated sediments.

4.2. Relationship between trace and minor metals

There are inorganic (e.g. clay minerals, Fe and Mn

oxides) and organic (represented by organic carbon) scav-

Page 8: Heavy metal distribution and controlling factors …dzumenvis.nic.in/Microbes and Metals Interaction/pdf...Heavy metal distribution and controlling factors within coastal plain sediments,

Fig. 4. Metal distribution patterns throughout different horizons. The graphs show that A-horizon is always the most leached horizon. While C-horizon shows

some metals depletion, due to the high heterogeneity of this horizon, adding a trend line could be misleading. The same pattern was observed for other trace

elements measured in this study. Outliers in both figures belong to B-horizon where most of the A-horizon elements accumulate.

T. Liaghati et al. / Environment International 29 (2003) 935–948942

engers influencing metal mobility and distribution within

coastal sediments and soils. Spearman’s rank coefficient

matrices enabled the identification of relationships between

metals and potential controlling factors in metal mobility. In

order to understand variability in metal correlations with the

controlling scavengers in different populations (sediments

and soils), correlation matrices were calculated for each of

the above categories separately.

Based on the correlation matrix obtained for sediments

(Table 3), inorganic scavengers (clays followed by Fe andMn

oxides and S) are the dominant factors controlling trace metal

distribution in the catchment. Organic carbon (OC) showed

weak or no association with trace metals. This confirms the

observation that OC is not a strong metal scavenger as are

oxides or clays in nonmarine environments. Metals such as

Cu and Ni (r = 0.6 and 0.5, respectively) were exceptions due

to their tendency to be kept in solution by chelation with

organic material (e.g. fulvic acids) (Thornton, 1981). As for

trace metals such as Mo and Co, concentrations were low

(around detection limit) for most samples and their correla-

tions were considered unreliable.

A weak correlation between Fe and S (r = 0.4) together

with the results of mineralogical analysis suggested that iron

is not predominantly present as pyrite but is in the form of

oxides such as goethite (FeO.OH) and/or hematite (Fe2O3). A

strong correlation between Fe and Al (r = 0.7) confirmed that

iron was primarily associated with the silt–clay fraction.

Finally, a moderate to strong correlation between Al and trace

elements demonstrated that these elements are associated

with clays.

In order to interpret vertical metal distribution within each

soil profile, cores were treated individually and subdivided

into A-, B- and C-horizons (based on descriptions and logs

provided by QASSIT). Cores were then regrouped according

Page 9: Heavy metal distribution and controlling factors …dzumenvis.nic.in/Microbes and Metals Interaction/pdf...Heavy metal distribution and controlling factors within coastal plain sediments,

Fig. 5. Lateral variation of metal distribution and its correlation with

mineralogy. Sample H1 contains larger amounts of clay minerals and

significantly more V and Cr compared to BN5. Sites shown in Fig. 2.

Fig. 6. Vertical variation of metal distribution with regard to mineralogy.

The above samples belong to the same core and were taken from different

depths: 0 cm (A-horizon), 130 cm (B-horizon) and 330 cm (C-horizon),

respectively. Sites in Fig. 3.

Fig. 7. Comparison of clay-rich with sand-rich material shows that sample

661 (80 cm) has a higher metal content than 662 at a similar depth.

T. Liaghati et al. / Environment International 29 (2003) 935–948 943

to their horizons and correlation matrices calculated for each

horizon separately (Table 4). The correlationmatrix for the A-

horizon (generally 0–100 cm deep) showed that all metal

scavengers (clays followed by Fe, Mn, S and OC) were

present. While OC is not necessarily a good terrestrial metal

scavenger, the A-horizon samples showed the highest organic

carbon and, therefore, the presence of OC as a scavenger in

this horizon was not surprising. In samples from the B-

horizon (about 50–280 cm deep), the role of S and C as

scavengers is reduced probably due to their lower concen-

Table 3

Spearman’s rank correlation matrix for surficial sediments (n= 21)

V Cr Mo Co Ni Cu Zn Pb As Mn Fe AL S OC

V 1.0

Cr 0.9 1.0

Mo 0.4 0.5 1.0

Co 0.8 0.7 0.3 1.0

Ni 0.7 0.7 0.5 0.6 1.0

Cu 0.6 0.7 0.4 0.6 0.9 1.0

Zn 0.5 0.4 0.5 0.6 0.7 0.7 1.0

Pb 0.8 0.8 0.3 0.7 0.6 0.6 0.4 1.0

As 0.4 0.3 0.1 0.2 0.5 0.3 0.2 0.3 1.0

Mn 0.6 0.5 0.5 0.5 0.6 0.6 0.5 0.3 0.2 1.0

Fe 0.9 0.8 0.4 0.6 0.6 0.5 0.5 0.5 0.5 0.6 1.0

AL 0.8 0.9 0.4 0.7 0.7 0.6 0.4 0.8 0.3 0.5 0.7 1.0

S 0.5 0.6 0.4 0.5 0.8 0.7 0.4 0.5 0.5 0.3 0.4 0.5 1.0

OC 0.2 0.2 0.0 0.4 0.5 0.6 0.3 0.4 0.3 0.0 � 0.1 0.2 0.7 1.0

Bold text shows strong correlations.

0.60–1.00 = strong correlation; 0.50–0.59 =moderate; 0.40–0.49 =weak;

0.00–0.39 = little or no association.

Page 10: Heavy metal distribution and controlling factors …dzumenvis.nic.in/Microbes and Metals Interaction/pdf...Heavy metal distribution and controlling factors within coastal plain sediments,

Table 4

Correlation matrices for A-, B- and C-horizons in soils

V Cr Mo Co Ni Cu Zn Pb As Fe Mn Al S OC

A-horizon

V 1.0

Cr 0.6 1.0

Mo 0.7 0.6 1.0

Co 0.4 0.1 0.4 1.0

Ni 0.4 0.8 0.6 0.2 1.0

Cu 0.4 0.5 0.3 0.3 0.4 1.0

Zn 0.6 0.6 0.6 0.2 0.3 0.4 1.0

Pb 0.8 0.4 0.8 0.5 0.2 0.3 0.7 1.0

As 0.4 0.1 0.2 0.7 0.2 0.1 0.0 0.2 1.0

Fe 0.9 0.7 0.7 0.5 0.6 0.5 0.7 0.6 0.5 1.0

Mn 0.8 0.6 0.6 0.6 0.6 0.6 0.6 0.7 0.5 0.9 1.0

Al 0.9 0.5 0.8 0.5 0.3 0.3 0.7 0.9 0.3 0.8 0.8 1.0

S 0.8 0.6 0.6 0.6 0.4 0.5 0.6 0.7 0.5 0.8 0.9 0.8 1.0

OC 0.6 0.5 0.6 0.3 0.5 0.0 0.5 0.7 0.2 0.6 0.6 0.7 0.7 1.0

B-horizon

V 1.0

Cr 0.3 1.0

Mo 0.7 0.4 1.0

Co 0.5 0.2 0.5 1.0

Ni 0.0 0.9 0.2 0.2 1.0

Cu 0.4 � 0.1 0.1 0.4 � 0.2 1.0

Zn 0.2 0.0 0.3 0.7 0.0 0.5 1.0

Pb 0.7 0.3 0.9 0.7 0.2 0.2 0.4 1.0

As 0.7 0.3 0.6 0.7 0.1 0.6 0.4 0.7 1.0

Fe 0.8 0.4 0.6 0.6 0.1 0.2 0.3 0.7 0.8 1.0

Mn 0.8 0.5 0.4 0.6 0.4 0.4 0.2 0.6 0.8 0.9 1.0

Al 0.8 0.5 0.9 0.7 0.4 0.2 0.4 0.9 0.7 0.8 0.7 1.0

S 0.4 0.0 0.3 0.4 0.0 0.2 � 0.1 0.5 0.5 0.6 0.5 0.5 1.0

OC 0.4 � 0.2 0.5 0.5 � 0.3 0.3 0.2 0.6 0.5 0.4 0.3 0.5 0.7 1.0

C-horizon

V 1.0

Cr 0.9 1.0

Mo 0.4 0.5 1.0

Co 0.2 0.3 0.0 1.0

Ni 0.4 0.4 0.2 0.8 1.0

Cu 0.6 0.5 0.0 0.1 0.3 1.0

Zn 0.6 0.6 0.2 0.4 0.5 0.5 1.0

Pb 0.6 0.6 0.4 0.3 0.3 0.3 0.5 1.0

As 0.5 0.5 0.2 � 0.1 0.2 0.7 0.3 � 0.1 1.0

Fe 0.5 0.5 0.1 0.5 0.6 0.4 0.3 0.0 0.5 1.0

Mn 0.4 0.5 0.2 0.5 0.7 0.6 0.4 0.0 0.4 0.7 1.0

Al 0.9 0.8 0.4 0.3 0.4 0.5 0.7 0.8 0.3 0.3 0.3 1.0

S 0.1 0.1 0.0 0.5 0.7 0.4 0.4 0.0 0.1 0.5 0.8 0.1 1.0

OC 0.1 0.1 � 0.2 0.2 0.3 0.4 0.5 0.2 � 0.1 0.0 0.3 0.3 0.5 1.0

Indices indicate the distinct change of main scavengers over metal distribution as the soil horizons change.

T. Liaghati et al. / Environment International 29 (2003) 935–948944

tration in this horizon. Trace metals such as V, Pb and As

continue to be strongly correlated with clays, Fe and Mn

oxides, as this horizon is where the clays, Fe and Mn oxides

leached from the above horizons tend to accumulate. How-

ever, Zn, Cu and Cr showed only weak, little or no association

with the main scavengers. This may be due to the fact that

these metals tend to be adsorbed by clay minerals such as

smectite and illite (e.g. Alloway, 1995). In the majority of

samples, kaolinite was the dominant clay while smectite only

occurred in very low concentration. In the C-horizon the

dominant scavenger was Al, which showed a strong correla-

tion with all elements except for Cu (r= 0.5), which showed a

moderate correlation. This distribution is because the C-

horizon is comprised mainly of unconsolidated rock material

and is less affected by the leaching processes, which occur in

the upper horizons.

4.3. Approach to normalisation

As the samples analysed in this project were from a wide

range of unconsolidated material taken from various loca-

tions within the catchment, the data produced represent a

Page 11: Heavy metal distribution and controlling factors …dzumenvis.nic.in/Microbes and Metals Interaction/pdf...Heavy metal distribution and controlling factors within coastal plain sediments,

T. Liaghati et al. / Environment International 29 (2003) 935–948 945

heterogeneous geochemical data set controlled by a number

of sediment characteristics. Therefore, normalisation was

required to interpret data in more detail and to describe

patterns of distribution.

After testing Fe and OC, Al was chosen as the most

suitable normalizing element. Extractable Al is likely to be

associated with finer particles and is also one of the

conservative metals which is not affected by anthropogenic

activities. Therefore, aluminium has been used successfully

by several workers to account for grain size effect (Ryan and

Windom, 1988; Sinex and Wright, 1988; Balls et al., 1997;

Rubio et al., 2000). After choosing the best normalizing

factor, the next step was deciding on the most appropriate

sample as a reference to use in calculating EF. A mean of

five samples from the B-horizon (130–280 cm deep),

containing older sediments, most likely deposited under

preindustrial conditions, was chosen as reference. Further-

more, unlike A- and C-horizons, the B-horizon is more

structured and consistent (neither too close to surface nor

bedrock), and therefore, is more representative of sedimen-

tary environment (Murphy, 1991). The reference samples

were also located in the lower section of the catchment

within native vegetation, which has never been disturbed by

modern activities and, therefore, represented the background

values of the metals for the local area. Due to natural

mineralogical differences of the sediments and analytical

uncertainty, only sediments with an EF greater than 2 were

considered as enriched (e.g. Angelidis and Aloupi, 1997).

The EFs calculated for sediments revealed that they were

only enriched in V. The enrichment factor varied from 2 to 9.5

in enriched sites (B3, BN1, BN3, BS1, BS4, H1 andH2) (Fig.

2). All sites were located in native vegetation (Melaleuca and

mangroves) and the fact that they have not been disturbed

suggests that enrichment was in situ. Vanadium in sediment

solutions occurs predominantly in the + 5 and + 4 oxidation

states as the vanadate forms and as the vanadyl cation, VO2 +.

Under oxidised conditions with pH from neutral to alkaline,

V (e.g. V3 +) has high mobility and bioavailability. Under

reduced conditions, humus promotes easy reduction of vana-

date to vanadyl and causes the immobilisation of V when it is

bound to organic material (McBride, 1994). The immobiliz-

ing factors mentioned above are likely to be present in the

various enriched sites analysed. In samples B3, BN1 and

BS1, organicmatter probably was the main factor influencing

in immobilisation of V as they have 4.7%, 4.3%, and 5.5%

total organic carbon, respectively (the highest OC for this

sample set is 5.5%). Unstable forms of V may substitute

readily for Fe3 + in minerals such as Fe oxides and/or layer

silicate clays. This may be the case for some sites where Fe

oxides are abundant such as samples BN3 and BS4 with

hematite 2% and 0.6%, respectively. Pyrite and goethite

(0.8% and 0.5%, respectively) were found in sample H1,

while sample H2 contained extremely high goethite (in order

of tens of percent).

Soil samples (Fig. 3) were found to be enriched in Cr, Zn

and V. Chromium enrichment was found at two sites (542

and 509) and varied from 2.2 to 5. This element may be

present here as Cr3 + (chromic form), a very immobile cation

whose solubility decreases above pH 4 while above pH 5.5

complete precipitation occurs (pH in these sites varied

between 4.7 and 5.7) (McBride, 1994; McGrath and Smith,

1995). Mineralogically, while site 542 contained up to 3%

smectite, which is a strong adsorbent for trace elements, site

509 was very sandy. The high Cr at both sites (542 and 509)

may have originated from agricultural material (e.g. nitrate

fertilizers), as these sites are located close to turf and sugar

cane farms. Whan (2002) reported that the concentration of

nutrients (e.g. nitrate) for sites 542 and 509 was 784 and

1038 mg/kg, respectively (nitrate concentration was

between 61 and 1718 mg/kg for 21 sites and the above

sites were among those with highest nitrate).

Only three sites (509, 652 and 655) of surficial samples

(0–80 cm) were enriched in Zn with the EFs that varied from

4.6 to 6. In acid, aerobic soils, Zn has medium mobility as it

is held in exchangeable forms on clays and organic matter. At

higher pH (as is the case with enriched samples), chemi-

sorption on oxides and aluminium silicates, and complex-

ation with humus can decrease the solubility and mobility of

Zn2 +. The enriched sites here are all very sandy (up to 99%

quartz) and located in the lower section of the catchment; due

to intense leaching and weathering, these samples contain

relatively low amount of clays and other secondary minerals.

Cause of enrichment may be related to current land use in the

area. Sites 509 and 652 are adjacent to a sugar cane farm, and

655 is located within a golf course, thus fertilizer impurities

may be the reason for increased Zn concentration at the

above sites (e.g. Alloway, 1995; McMurtry et al., 1995).

Sites 562 and 651 were enriched in V (EF from 2 to 3.5).

Both cores are very clayey (55–60% kaolinite and 3–6% of

mixed layers of illite–smectite). Therefore, V enrichment

may be due to the high clay abundance in these cores.

Mineralogy of surficial sediments is an effective way to

explain the metal distribution among these samples. Samples

are typically recent, less reworked and contained significant

amounts of secondary minerals and organic carbon. More-

over, as there were not any point sources of pollution close to

the enriched sites, in situ enrichment due to significant

occurrence of organic and inorganic metal adsorbents was

the best explanation. For soil samples, however, land use in

the area is an additional means of explaining the enrichment

and of showing that there are several point sources for metals

such as Zn, Cr and V. Therefore, the limited amount of

scavenger materials at the sandy sites of the lower catchment

suggests that the enrichment cannot be natural and most

likely related to local land use.

To obtain a better understanding of enrichment factors, a

comparison between absolute and normalized metal con-

tents in soil cores with respect to Al content was made.

Values of Al (Fig. 8) demonstrate an irregular distribution

with core depth, indicating a lack of pattern in clay content

within the profile. While in cores with higher clay content

(high Al) absolute and normalized values covary, in sandier

Page 12: Heavy metal distribution and controlling factors …dzumenvis.nic.in/Microbes and Metals Interaction/pdf...Heavy metal distribution and controlling factors within coastal plain sediments,

Fig. 8. A comparison between absolute and normalised metal (V and Zn) content for cores 652, 562 and 651, and a comparison between absolute and

normalised values for Zn, V and Al.

T. Liaghati et al. / Environment International 29 (2003) 935–948946

cores with lower clay content, the positive correlation

becomes negative. Therefore, in a sandy core as the metal

concentration decreases, EF increases and a lower absolute

metal content does not necessarily demonstrate lower values

for EF and vice versa (Fig. 8). Overall, this comparison

showed that in a highly heterogeneous environment, such as

this setting, interpretations based on absolute concentration

of different elements alone could be misleading. Therefore,

normalizing the data is the best approach for sensible

interpretation of geochemical data.

5. Conclusions

The chemical analyses of three types of material (sedi-

ments, soils and bedrock) in this study helped to explain the

spatial and vertical distribution of trace elements within the

area. Analysing bedrock samples established the background

values for elements which is essential when assessing levels

occurring in the adjacent unconsolidated sediments. Overall,

mineral and chemical characteristics of the three categories of

samples analysed were systematic. However, varying degrees

of weathering have resulted in the heterogeneous nature of

the material, which in turn governs the trace metal distribu-

tion. Quartz was an important component in all three cate-

gories analysed; feldspars, however, were in more abundance

in surficial sediments compared to soils and some weathered

bedrock samples. The dominant trace metals were Zn and V,

followed by Cr and Pb. This overall pattern is preserved

throughout the study area, except for Cr, which was occa-

sionally present in higher concentrations than V, presumably

due to its immobility.

Page 13: Heavy metal distribution and controlling factors …dzumenvis.nic.in/Microbes and Metals Interaction/pdf...Heavy metal distribution and controlling factors within coastal plain sediments,

T. Liaghati et al. / Environment International 29 (2003) 935–948 947

In comparing sediment and soil samples, the correlation

matrix showed that while inorganic scavengers such as clay

minerals are dominant factors controlling trace metal dis-

tribution in sediments, in soil cores the relationships

between metals were different in each horizon. Both organic

and inorganic metal scavengers are present in the A-horizon;

in the B-horizon, OC and S occur in lower concentrations

and the Al is the strongest scavenger here. Finally, in the C-

horizon, Al is almost the only scavenger present. This is

probably because this horizon had been less affected by the

leaching processes, which were dominant in other horizons.

Data produced in this study represent a heterogeneous

geochemical data set controlled by various processes. The

enrichment factors calculated for sediments showed that

they had been enriched only in V. All enriched sites were

located within natural undisturbed vegetation so enrichment

was most likely due to the presence of organic material and/

or secondary minerals such as clays, hematite and goethite.

In soil samples, however, the enriched metals were Cr, Zn

and V. As most enriched sites contained relatively low

amount of clays and were located close to agricultural lands,

the current land use in the area is more likely to be the cause

of the enrichment, e.g. fertilizer impurities.

Thus, the main findings of this study were:

(1) Metal elevation was many times due to natural enrich-

ment and not to contamination.

(2) When analysing a large data set consisting of samples

from a variety of settings, it is essential to apply norma-

lisation as a tool to transform the heterogeneous geo-

chemical data and compensate for grain size effect.

(3) Natural sediment/soil characteristics such as mineralogy

are more important in controlling metal occurrence and

spatial/vertical distribution compared to anthropogenic

activities such as local land-use practices in the Bells

Creek catchment.

Acknowledgements

This study was funded by Lensworth Group. Authors

would like to thank Hayden McDonald from Mipela for the

GIS database as well as Queensland Acid Sulfate Soils

Investigation Team (QASSIT) for providing soil and

bedrock samples. We also wish to thank Bill Kwiecien,

Whatsala Kumar and Tony Raftery for practical assistance

with chemical and mineralogical analyses. Our colleague,

Tim Ezzy, is also thanked for assisting with fieldwork and

additional mapping information. Graham Kimber is greatly

appreciated for his inputs regarding data quality control

procedures.

References

Alloway BJ. Heavy metals in soils. Glasgow, UK: Blackie Academic &

Professional, an imprint of Chapman & Hall; 1995. p. 368.

Angelidis MO, Aloupi M. Assessment of metal contamination in shallow

coastal sediments around Mytilene, Greece. Int J Environ Anal Chem

1997;68(2):281–93.

Angelidis MO, Aloupi M. Geochemical study of coastal sediments influ-

enced by river transported pollution: southern Evoikos Gulf, Greece.

Mar Pollut Bull 2000;40(1):77–82.

Arakel AV, Hongjun T. Heavy metal geochemistry and dispersion pattern in

coastal sediments, soil and water of Kedron Brook floodplain area,

Brisbane. Environ Geol Water Sci 1992;20:219–31.

Balls PW, Hull S, Miller BS, Pirie JM, Proctor W. Trace metal in Scottish

estuarine and coastal sediments. Mar Pollut Bull 1997;34:42–50.

Chakrapani GJ, Subramanian V. Heavy metals distribution and fractiona-

tion in sediments of the Mahanadi River basin, India. Environ Geol

1993;22:80–7.

Chamley H. Clay sedimentology. Berlin: Springer-Verlag; 1989.

Chao TT. Use of partial dissolution techniques in geochemical exploration.

J Geochem Explor 1984;20:101–35.

Chen Z, Kostaschuk R, Yang M. Heavy metals on tidal flats in the Yangtze

Estuary, China. Environ Geol 2001;40(6):742–9.

Cox ME, Preda M, Brooke B. General features of the geo-setting of the

Pumicestone region. In: Cox ME, editor. PASSCON 2000. Brisbane:

School of Natural Resource Sciences, Queensland University of Tech-

nology; 2000. p. 7–12.

Daskalakis KD, O’Connor TP. Normalization and elemental sediment con-

tamination in the coastal United States. Environ Sci Technol 1995;

29(2):470–7.

Data DK, Subramanian V. Distribution and fractionation of heavy metals in

the surface sediments of the Ganges-Brahmaputra-Meghna river system

in the Bengal basin. Environ Geol 1998;36(1–2):93–101.

De Carlo EH, Anthony SS. Spatial and temporal variability of trace element

concentrations in an urban subtropical watershed, Honolulu, Hawaii.

Appl Geochem 2002;17:475–92.

Ellaway M, Hart BT, Beckett R. Trace metals in sediments from the Yarra

river. Aust J Mar Freshw Res 1982;33:761–78.

Fang T, Hong E. Mechanisms influencing the spatial distribution of trace

metals in surficial sediments off the Southwestern Taiwan. Mar Pollut

Bull 1999;38(11):1026–37.

Flood PG. Carbon-14 dates from the coastal plains of Deception Bay,

Southeastern Queensland. Queensl Gov Min J 1981;82:19–23.

Forstner U, Salomons W. Trace metal analysis on polluted sediments: Part

I. Assessment of sources and intensities. Environ Technol Lett 1980;1:

494–506.

Forstner U, Calmano W, Schoer J. Metals in sediments from the Elbe,

Weser and Ems estuaries and from the German Bight: grain size effects

and chemical forms. Thalass Jugosl 1982;18(1–4):97–122.

Forstner U, Wolfgang A, Wolfgang C, Kersten M, Salomons W. Mobility

of heavy metals in dredged harbour sediments. In: Sly PG, editor. Sedi-

ment and water interaction: proceedings of the third international sym-

posium on interactions between sediments and water. New York:

Springer-Verlag; 1984. p. 371–80.

Fung YS. Analysis and assessment of heavy metal pollution in Hong Kong’s

marine environment. Mar Biol South China Sea 1993;261–72.

Kelly AG, Wells DE, Fryer RJ. Sampling strategy to detect a change in

concentration of trace organic contaminants in marine sediment. Sci

Total Environ 1994;144:217–30.

Koljonen T, Malisa E. Solubility in aqua regia of selected chemical ele-

ments occurring in the fine-grained fraction of till. In: Pulkkinen E,

editor. Environmental geochemistry in northern Europe, special paper

9. Helsinki: Geological Survey of Finland; 1991. p. 49–52.

Krauskopf KB. Introduction to geochemistry. New York: McGraw-Hill;

1967.

Leckie JO, Parks GA. Geochemistry and environmental impact. In: Ho-

ward AD, Remson I, editors. Geology in environmental planning. New

York: McGraw-Hill; 1978. p. 276–88.

Lester J. Geomorphology, sedimentology and shoreline processes impact-

ing on the stability of the Bribie Island Spit. Honours thesis, Queens-

land University of Technology; 2000.

Page 14: Heavy metal distribution and controlling factors …dzumenvis.nic.in/Microbes and Metals Interaction/pdf...Heavy metal distribution and controlling factors within coastal plain sediments,

T. Liaghati et al. / Environment International 29 (2003) 935–948948

Loring DH. Lithium—a new approach for the granulometric normalization

of trace metal data. Mar Chem 1990;29:155–68.

Loring DH. Normalization of heavy-metal data from estuarine and coastal

sediments. J Mar Sci 1991;48:101–15.

Loring DH, Rantala RTT. Manual for geochemical analyses of marine

sediments and suspended particulate matter. Earth-Sci Rev 1992;32:

235–83.

McBride MB. Environmental chemistry of soils. New York: Oxford Univ

Press; 1994. 406 pp.

McConchie DM,MannAW, LinternMJ, LongmanD, Talbot V, GabelishAJ,

Gabelish MJ. Heavy metals in marine biota, sediments and waters from

the Shark Bay area, Western Australia. J Coast Res 1988;4(1):37–58.

McGrath SP, Smith S. Chromium and nickel. In: Alloway BJ, editor. Heavy

metals in soils. New York: Blackie, Wiley; 1995.

McMurtry GM, Wiltshire JC, Kauahikaua JP. Heavy metal anomalies in

coastal sediments of Oahu, Hawaii. Pac Sci 1995;49:452–70.

Murphy BW. The soil profile. In: Charman PEV, Murphy BW, editors.

Soils their properties and management. Sydney: Sydney University

Press in association with Oxford University Press Australia, 1991.

Murphy PR, Trezise DL, Hutton LJ, Cranfield LC. 1:100,000 Geological

map commentary. Caboolture, sheet 9443, Queensland. Department of

Mines, Geological Survey of Queensland: Brisbane; 1987.

Niencheski LF, Windom HL, Smith R. Distribution of particulate trace

metal in Patos Lagoon estuary (Brazil). Mar Pollut Bull 1994;28(2):

96–102.

NSR Environmental Consultants. Caloundra Downs informal land use in-

vestigation: Pumicestone Passage study. Report prepared for Lensworth

Group, Brisbane, Queensland; 1999a.

NSR Environmental Consultants. Caloundra Downs informal land use in-

vestigation: consolidated data report, main volume. Report prepared for

Lensworth Group, Brisbane, Queensland; 1999b.

Osterholm P, Astrom M. Spatial trends and losses of major and trace

elements in agricultural acid sulphate soils disturbed in the artificially

drained Rintala area, W. Finland. Appl Geochem 2002;17:1209–18.

Owen RB, Sandhu N. Heavymetal accumulation and anthropogenic impacts

on Tolo Harbour, Hong Kong. Mar Pollut Bull 2000;40(2):174–80.

Power M, Attrill MJ, Thomas RM. Heavy metal concentration trends in the

Thames estuary. Water Res 1999;33(7):1672–80.

Preda M, Cox ME. Sediment–water interaction, acidity and other water

quality parameters in a subtropical setting, Pimpama River, Southeast

Queensland. Environ Geol 2000;39(3–4):319–29.

Preda M, Cox ME. Trace metals is acid sediments and waters, Pimpama

River, Southeast Queensland. Environ Geol 2001;40(6):755–68.

Preda M, Price S, Cox ME. Trace metal occurrence in estuarine sediments

and mangroves, Pumicestone Passage. In: Cox ME, editor. PASSCON.

Brisbane, Australia: Queensland University of Technology; 2000. p. 161.

Raisanen ML. Geochemistry of podzolized tills and the implications for

aluminium mobility near industrial sites: a study in Kuopio, eastern

Finland. PhD dissertation, University of Abo, Finland; 1996.

Rose AW, Hawkes HE, Webb JS. Geochemistry in mineral exploration.

London: Academic Press; 1979.

Rubio B, Nombela MA, Vilas F. Geochemistry of major and trace elements

in sediments of Ria de Vigo (NW Spain): an assessment of metal

pollution. Mar Pollut Bul. 2000;40(11):968–80.

Ruiz F. Trace metals in estuarine sediments from southwestern Spanish

coast. Mar Pollut Bull 2001;42(6):482–90.

Ryan JD, Windom HL. A geochemical and statistical approach for

assessing metal pollution in coastal sediments. In: Seeliger U, editor.

Metals in coastal environments of Latin America. Berlin: Springer;

1988. p. 47–58.

Schneider PM, Davey SB. Sediment contaminants off the coast of Sydney,

Australia: a model for their distribution. Mar Pollut Bull 1995;31(4–12):

262–72.

Sinex SA, Wright DA. Distribution of trace metals in the sediments and

biota of Chesapake Bay. Mar Pollut Bull 1988;19:425–31.

Tam NFY, Yao MWY. Normalization and heavy metal contamination in

mangrove sediments. Sci Total Environ 1998;216:33–9.

Taylor SR. Abundance of chemical elements in the continental crust: a new

table. Geochim Cosmochim Acta 1964;28:1273–85.

Thornton I. Geochemical aspects of the distribution and forms of heavy

metals in soils. In: Lepp NW, editor. Effects of heavy metal pollution on

plants. London and New Jersey: Applied Science Publishers, 1981.

Trimble CA, Hoenstine RW. Baseline investigation of estuarine sediment

metals for the Steinhatchee River area of the Florida Big Bend. Environ

Geosci 1997;4(2):95–103.

Whan E. Nutrient chemistry of the surface drainage system and sediments

in Bells Creek catchment: sources, controls and relation to land use.

Honours, Queensland University of Technology; 2002.

Williams M, Dunkerley D, De Deckker P, Kershaw P, Chappell J. Quater-

nary environments. London: Arnold, Hodder Headline Group; 1998.

Windom HL, Schropp SJ, Calder FD, Ryan JD, Smith RG, Burney LC,

Lewis FG, Rawlinson CH. Natural trace metal concentrations in estuar-

ine and coastal marine sediments of the southern United States. Environ

Sci Technol 1989;23(3):314–20.